Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

RESEARCH ARTICLE New GNSS and Geological Data From the Indo-Burman

10.1029/2022JB025550
Subduction Zone Indicate Active Convergence on Both a
Key Points:
• T he Indo-Burma subduction zone is
Locked Megathrust and the Kabaw Fault
capable of hosting large megathrust Bar Oryan1,2 , Paul M. Betka3 , Michael S. Steckler1 , Scott L. Nooner4 , Eric O. Lindsey5 ,
earthquakes
Dhiman Mondal6, Austin M. Mathews3, Syed Humayun Akhter7, Sanju Singha8, and Oo Than9
• the Kabaw Fault accommodates
strike-slip and convergence motion 1
Lamont-Doherty Earth Observatory, Palisades, NY, USA, 2Laboratoire de Géologie, Département de Géosciences, École
previously ascribed to the Indo-Burma
megathrust Normale Supérieure, PSL Université, CNRS UMR, Paris, France, 3Atmospheric, Oceanic, and Earth Sciences, George
• We reveal a previously unrecognized Mason University, Fairfax, VA, USA, 4Department of Earth and Ocean Sciences, University of North Carolina Wilmington,
seismic hazard associated with the Wilmington, NC, USA, 5Department of Earth and Planetary Sciences, University of New Mexico, Albuquerque, NM, USA,
Kabaw Fault 6
MIT Haystack Observatory, Westford, MA, USA, 7Bangladesh Open University, Gazipur, Bangladesh, 8Department of
Geology, Dhaka University, Dhaka, Bangladesh, 9Division of Seismological, Department of Meteorology and Hydrology,
Supporting Information: Nay Pyi Taw, Myanmar
Supporting Information may be found in
the online version of this article.
Abstract The Indo-Burma subduction zone is a highly oblique subduction system where the Indian plate
Correspondence to:
is converging with the Eurasian plate. How strain is partitioned between the Indo-Burma interface and upper
B. Oryan, plate Kabaw Fault, and whether the megathrust is a locked and active zone of convergence that can generate
bar.oryan@columbia.edu great earthquakes are ongoing debates. Here, we use data from a total of 68 Global Navigation Satellite System
(GNSS) stations, including newly installed stations across the Kabaw Fault and compute an updated horizontal
Citation: and vertical GNSS velocity field. We correct vertical rates for fluctuating seasonal signals by accounting for
Oryan, B., Betka, P. M., Steckler, M. S., the elastic response of monsoon water on the crust. We model the geodetic data by inverting for 11,000 planar
Nooner, S. L., Lindsey, E. O., Mondal, D., and non-planar megathrust fault geometries and two geologically viable structural interpretations of the Kabaw
et al. (2023). New GNSS and geological
data from the Indo-Burman subduction Fault that we construct from field geological data, considering a basin-scale wedge-fault and a crustal-scale
zone indicate active convergence on both reverse fault. We demonstrate that the Indo-Burma megathrust is locked, converging at a rate
𝐴𝐴 of 11.6 ± 5.4
a locked megathrust and the Kabaw Fault. mm/yr, and capable of hosting >8.2 Mw megathrust events. We also show that the Kabaw Fault is locked and
Journal of Geophysical Research: Solid
Earth, 128, e2022JB025550. https://doi. accommodating strike-slip motion at a rate
𝐴𝐴 of 8.4 ± 3.0 mm/yr and converging at a rate
𝐴𝐴 of 5.7 ± 4.1 mm/yr. Our
org/10.1029/2022JB025550 interpretation of the geological, geophysical, and geodetic datasets indicates the Kabaw Fault is a crustal-scale
structure that actively absorbs a portion of the convergence previously ascribed to the Indo-Burma megathrust.
Received 31 AUG 2022 This reveals a previously unrecognized seismic hazard associated with the Kabaw Fault and slightly reduces the
Accepted 6 APR 2023
Corrected 15 MAY 2023 estimated hazard posed by megathrust earthquakes in the region.

This article was corrected on 15 MAY Plain Language Summary Subduction zones are plate boundaries where two tectonic plates
2023. See the end of the full text for converge and can generate large earthquake along its main fault, the megathrust. Earthquakes rupture on
details.
faults that remain locked and accumulate strain during the interseismic period, the period of time between
great earthquakes. The Indo-Burma subduction zone, where the Indian plate is converging with the Eurasia
Author Contributions:
plate, is home to more than 200 million people. In spite of its enormous population, it is unclear whether the
Conceptualization: Bar Oryan, Paul
M. Betka, Michael S. Steckler, Syed Indo-Burma subduction zone can generate large megathrust earthquakes. To evaluate its potential to host
Humayun Akhter great earthquakes, we combine new GPS datasets that record plate motion during the interseismic period with
Data curation: Bar Oryan, Paul M. geological analysis of the Kabaw Fault, one of the secondary faults in the Indo-Burma subduction zone. We
Betka, Michael S. Steckler, Eric O.
Lindsey, Dhiman Mondal, Syed Humayun demonstrate that Indo-Burma megathrust and Kabaw Fault are locked and can generate significant earthquakes.
Akhter, Sanju Singha, Oo Than For the first time, we show that the Kabaw Fault absorbs a portion of the convergence previously ascribed to the
Indo-Burma megathrust which could explain the long repeat time between great earthquakes on the megathrust.

1. Introduction
© 2023. The Authors.
This is an open access article under The Indo-Burma subduction zone lies on the eastern flank of the Indian plate north of the Sumatra-Andaman
the terms of the Creative Commons
Attribution-NonCommercial-NoDerivs subduction zone. It is a highly oblique subduction system where the Indian plate is obliquely converging with
License, which permits use and the Eurasian plate (Figure 1). Whether this megathrust is still locked and an active zone of convergence is an
distribution in any medium, provided the ongoing debate (Gahalaut et al., 2013; Kundu & Gahalaut, 2012; Le Dain et al., 1984; R. Mallick et al., 2019;
original work is properly cited, the use is
non-commercial and no modifications or Mukhopadhyay & Dasgupta, 1988; Ni et al., 1989; Panda et al., 2020; Panda & Kundu, 2022; Purnachandra Rao
adaptations are made. & Kumar, 1999; Rangin et al., 2013; Satyabala, 1998; Steckler et al., 2016; Stork et al., 2008). The existence of

ORYAN ET AL. 1 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Formal analysis: Bar Oryan, Paul M. active folds, a broad accretionary prism, gravity anomalies, deep seismic activity, subduction metamorphism, and
Betka, Scott L. Nooner, Eric O. Lindsey,
an andesitic volcanic arc are indicative of past and current subduction (Le Dain et al., 1984; Mukhopadhyay &
Austin M. Mathews
Funding acquisition: Paul M. Betka, Dasgupta, 1988; Purnachandra Rao & Kumar, 1999; Satyabala, 1998; Stork et al., 2008). However, seismic activ-
Michael S. Steckler, Syed Humayun ity in the slab is dominated by north-south trending P axes suggesting that motion absorbed within the subducting
Akhter
plate may be occurring solely on dextral faults (Gahalaut et al., 2013; Kundu & Gahalaut, 2012; Ni et al., 1989;
Investigation: Bar Oryan, Paul M. Betka,
Michael S. Steckler, Scott L. Nooner, Eric Purnachandra Rao & Kumar, 1999; Rangin et al., 2013).
O. Lindsey, Dhiman Mondal
Methodology: Bar Oryan, Paul M. Betka, Geodetic surveys, which present the most reliable way of determining whether convergence on a megathrust
Michael S. Steckler, Scott L. Nooner, Eric is ongoing, suggest that east-west trending shortening is continuing to occur (Figure 2; R. Mallick et al., 2019;
O. Lindsey, Dhiman Mondal
Supervision: Paul M. Betka, Michael S. Panda et al., 2020; Steckler et al., 2016). However, how this motion is being accommodated by upper plate defor-
Steckler mation, such as the Kabaw Fault system (Pivnik et al., 1998; Wang et al., 2014), and whether this megathrust is
Validation: Bar Oryan, Paul M. Betka, capable of hosting great earthquakes is still highly controversial.
Michael S. Steckler, Dhiman Mondal,
Syed Humayun Akhter
Using horizontal Global Navigation Satellite System (GNSS) velocities from 45 sites obtained from 2003 to 2011,
Steckler et al. (2016) suggested that convergence is ongoing at a rate of ∼12–18 mm/yr along a shallow-dipping
locked decollement capable of generating an Mw 8.2–9 megathrust earthquake. R. Mallick et al. (2019) used
block modeling and updated GNSS horizontal velocities including data from 13 additional stations. Their results
supported Steckler et al. (2016) conclusions and yielded 12–24 mm/yr of motion across a locked megathrust,
decreasing from south to north. On the other hand, Panda et al. (2020) used GNSS horizontal velocities that
excluded stations on or near the Shan Plateau, which is internally deforming, and instead referenced convergent
motion to the Sunda block to the south. As a result, they determined significantly lower convergence rates than
the previous estimates. Furthermore, they demonstrated that this lower convergence could be accommodated
either through shallow creep or in a stick-slip manner, thus leading to uncertainty in the seismic hazard in this
region.

Upper plate fault systems such as the Sagaing Fault (SF), Churachandpur-Mao Fault (CMF), and Kabaw Fault
(KF) also accommodate motion. The SF and CMF absorb most of the dextral motion between the plates (Figures 1
and 2). However, previous studies could not determine whether the Kabaw Fault accommodates active strike-slip
and/or convergence due to a lack of geodetic observations and ambiguity regarding the structure of the Kabaw
Fault (Oryan et al., 2020; Steckler et al., 2016, 2021). The Kabaw Fault generally separates the IBR accretionary
prism on the Indian Plate from forearc basin deposits of the Salin and Chindwin Basins on the Burma Plate.
The map trace of the fault follows along the trend of a belt of ultramafic rocks, known as the Western Ophiolite
Belt, that define the Tethyan suture between India and Burma (Figures 1 and 2; Liu, Chung, et al., 2016; Liu,
Zhang, et al., 2016; Morley et al., 2020; Westerweel et al., 2020). Prior structural analyses of the Kabaw Fault
offer conflicting interpretations of the fault structure, including both east-dipping (Böker et al., 2019; Pivnik
et al., 1998; Wang et al., 2014) and west-dipping fault geometries (Maurin & Rangin, 2009). This uncertainty
in the geometry of Kabaw Fault makes it difficult to assess its seismic potential, its role in the strain partition-
ing of the Indo-Burma subduction zone, and its part in the accretion and exhumation of the Western Ophiolite
Belt. Earlier geodetic analysis, which did not explicitly account for the Kabaw Fault system structure, suggested
that it may accommodate a significant portion of the motion between the Indian and Eurasian plates (Steckler
et al., 2016). If true, this additional partitioning of India-Eurasia convergence by the Kabaw Fault may pose a new
unrecognized seismic risk, but also lower the convergence rate taken up by the subduction megathrust. The latter
would increase the repeat time for large megathrust events in the region and could partially explain why no clear
evidence for a significant recent megathrust earthquake in the Bengal Basin has been found.

In this study, we use data from 68 GNSS stations to compute an updated horizontal and vertical GNSS velocity
field for Bangladesh, Myanmar, and adjacent regions using the latest available data set. We correct vertical rates
for the oscillating seasonal signal by accounting for the elastic response of monsoon waters on the lithosphere.
In addition, for the first time, we include GNSS rates from newly installed stations across the Kabaw Fault zone
and the Indo-Burman Ranges (IBRs), and present two geologically constrained models of the Kabaw Fault zone
to help estimate its structure and seismic potential. The combination of these datasets together may help illumi-
nate the megathrust geometry, and determine whether convergence is accommodated along a freely slipping or a
locked megathrust. It may also help determine how strain is distributed across the Indo-Burma subduction zone
system, providing a better estimation of the seismic hazard in the region. With more than 200 million inhabitants,
the Bengal Basin is one of the most densely populated regions in the world (Figure 1), and accurately assessing
the seismic hazard in the region is of utmost importance.

ORYAN ET AL. 2 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

2. Regional Setting
The Indian and Eurasian plates started colliding ca. 50 Ma (Bouilhol
et al., 2013), forming the Himalayas and the Tibetan Plateau, two of Earth's
most prominent topographic features. The current boundary between these
two plates stretches from northwestern Pakistan to eastern India, where it
bends around an eastern syntaxis in Arunachal into the Naga thrust belt
(Wang et al., 2014). Then, it continues southwards, transitioning from a
continental collision zone to an oceanic subduction zone manifested in the
Andaman-Sumatra subduction zone (Figure 1). The Indian-Eurasia motion
is almost entirely convergent along the Himalayas, but is highly oblique
(∼70°) in the east at the Indo-Burma subduction zone. The obliquity of
Indo-Burman subduction is partitioned among multiple upper plate fault
systems that include several N-striking right-lateral faults. About half of this
motion is taken up on the Sagaing Fault (∼20 mm/yr) with smaller portions
accommodated by the CMF (∼10 mm/yr) and, potentially, the Kabaw Fault
(Gahalaut et al., 2013; R. Mallick et al., 2019; Maurin et al., 2010; Nielsen
et al., 2004; Sloan et al., 2017; Steckler et al., 2016; Vigny, 2003; Wang
et al., 2011, 2014).

This unique geological setting is home to the Ganges-Brahmaputra Delta


(GBD). The Ganges and Brahmaputra rivers drain 75% of the Himalayas and
currently carry ∼6% of the world's riverine sediments to the Bengal Basin,
making it the largest delta in the world (Milliman & Farnsworth, 2011;
Steckler et al., 2008). Below this 16- to 20-km thick pile of sediments lies the
extended crust of the Indian Shield or modified oceanic crust (Curray, 1991;
Figure 1. Regional map showing population density (CIESIN, 2018), large Mitra et al., 2008, 2018; Singh et al., 2016). The Indian Shield crust thins
earthquakes, and fault traces. Major thrust faults and dextral faults are denoted southeastward across the hinge zone of an early Cretaceous continental
by thick white lines with triangles and white lines, respectively. Yellow margin. Further east, at a location that is highly debated, the continental
triangles mark volcanoes. Maron and blue lines mark national boundaries crust eventually transitions into oceanic crust (Mitra et al., 2008; Rangin &
and major rivers, respectively. The rupture extent of the 1934 (Sapkota
Sibuet, 2017; Talwani et al., 2016, 2017). The heavily sedimented GBD is
et al., 2013), 1950 (Coudurier-Curveur et al., 2020), 1548 (Gait, 1906; Iyengar
et al., 1999), 1897 (Subedi & Hetényi, 2021), 1762 (Cummins, 2007; Mondal also being overthrust from the north by the Shillong Massif, a 2-km high
et al., 2018; Wang et al., 2013), 2005 (Ishii et al., 2005; Lay et al., 2005), and basement-cored anticlinorium exposing Indian Shield (Clark & Bilham, 2008;
events along the Sagaing fault (Hurukawa & Maung Maung, 2011) are marked R. Mallick et al., 2020; Najman et al., 2016).
by colored sections. Lower Inset: Regional map with the red box shows the
location of the main map. The IBRs bound the GBD on the east and extend 1,400 km from the Assam
syntaxis in the north to the Bay of Bengal in southwestern Myanmar. These
ranges are the surface expression of the forearc and a 375-km wide accre-
tionary prism (Betka, Seeber, Thomson, Steckler, et al., 2018) that deforms Triassic-Pliocene deposits of the
Tethys ocean and ancestral Brahmaputra Delta (Alam et al., 2003; Betka et al., 2021; Govin et al., 2018; Morley
et al., 2020; Sincavage et al., 2020). The accretionary prism is the world's flattest (surface slope of 0.1°), indicat-
ing a very weak detachment (Betka, Seeber, Thomson, Steckler, et al., 2018; Steckler et al., 2008). The outermost
western folds of the accretionary prism extend above the detachment and are blind, buried by the rapid sedi-
mentation of the GBD (Sikder & Alam, 2003). The anticlines overlie a detachment that is as shallow as 3–4 km
in Tripura, India (Betka, Seeber, Thomson, Steckler, et al., 2018), but deepens to the north and south of that
culmination (Abdullah et al., 2022; Bürgi et al., 2021; Najman et al., 2016). Farther east in the inner foldbelt, the
surface slope increases to 0.5° and structures that overlie the detachment indicate that it dips ∼7° east and deepens
to ∼10 km (Betka, Seeber, Thomson, Sincavage, et al., 2018). Geomorphic analysis of the IBR indicates that the
most active regions of tectonic uplift are located in the inner belt, with uplift particularly concentrated between
the CMF and Kabaw Faults (Maneerat & Bürgmann, 2022; Maneerat, Bürgmann, et al., 2022).

The Kabaw Fault system lies west of the Sagaing Fault, in Myanmar, and juxtaposes ultramafic rocks of the
Naga-Kalemyo (Western) Ophiolite Belt with Cretaceous to Pleistocene sedimentary rocks of the Salin and
Chindwin Basins (Figures 2 and 3). Structural analysis and reflection seismic interpretations from the Salin Basin
recognize an east-dipping listric thrust fault that uplifts and tilts Cretaceous-Pleistocene Salin Basin strata in the
hanging wall and has been named the Kabaw Fault (Pivnik et al., 1998; Wang et al., 2014). In contrast, Maurin

ORYAN ET AL. 3 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 2. Horizontal Global Navigation Satellite System (GNSS) velocities and major faults and geological units. Blue,
Green, magenta, black, and red arrows show horizontal GNSS velocities in the Indian frame of reference for stations
installed by Lamont Doherty Earth Observatory (LDEO), Earth Observatory of Singapore, the Myanmar Department of
Meteorology and Hydrology, the Myanmar Earthquake Committee, the Geological Survey of Bangladesh (EOS; R. Mallick
et al., 2019), Gahalaut et al. (2013), Band-Aid (Shum et al., 2014), and Maurin et al. (2010). Arrow base marks the position
of the GNSS sites with three stations discussed in the main text denoted by their four-letter code. See Figure S1 in Supporting
Information S1 for the entire four-letter code list of GNSS sites. Yellow region denotes the Chindwin (CB) and Salin (SB)
sedimentary basins and purple color marks the location of the Western Ophiolite Belt after Morley et al. (2020). Blind, outer,
and inner indicate the position of folds. Blue area marks the surface of the Kaptai reservoir (KR). Black lines with black
triangles denote major thrust faults and black lines without ornaments mark strike-slip faults. The approximate position of
the deformation and thrust fronts are marked as dashed curve and black line with triangles, respectively (Betka, Seeber,
Thomson, Steckler, et al., 2018). Red lines mark international borders.

and Rangin (2009) interpret the Kabaw Fault as a steeply west-dipping crustal-scale reverse fault that uplifts the
core of the IBR, including ultramafic rocks of the Western Ophiolite Belt, in its hanging wall. The Western Ophi-
olites are part of a suture zone that records Mesozoic closure of Tethyan-realm ocean basins and defines the India-
Asia suture in Myanmar, continuous with the Indus-Yarlung suture zone of Tibet (cf. Liu, Chung, et al., 2016;
Liu, Zhang, et al., 2016; Searle et al., 2017). Importantly, Morley et al. (2020) review several published interpre-
tations of the Kabaw Fault and identify geologic evidence for west-vergent right-transpressional shearing along
its central and southern segments. This latter point suggests that any recent deformation along the Kabaw Fault
may be partly controlled by the geometry of preexisting crustal-scale structures in the suture zone.

3. Geology and Kinematic Modeling of the Kabaw Fault


3.1. Field and Kinematic Modeling Methods

A geologic strip map was constructed near latitude 23.2°N to help constrain the geometry of the Kabaw Fault.
Bedding dips (n = 725) and stratigraphic units were mapped from road-cut outcrops along a transect that parallels
the Myittha (Manipur) River during a field excursion in 2017 (Figure 3a; Table S5). Geologic units were extrap-
olated in the study area from the Geologic Map of Myanmar (Tun et al., 2014). To help inform the kinematic
models, a population of minor faults (n = 27) were measured (where preserved) in the study area. Brittle fault
slip indicators were used to determine the sense of slip (Petit, 1987). Kinematic axes (incremental shortening, P;
and extension, T axes) were calculated for each datum using the software FaultKin (Allmendinger et al., 2012;
Marrett & Allmendinger, 1990). Principal shortening and extension axes for fault populations within each struc-
tural domain were determined by calculating directional maxima (e1—extension, e2—intermediate, and e3—

ORYAN ET AL. 4 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 3. (a) Geologic map of the Chindwin Basin and eastern Chin Hills showing bedding strike and dip data across the
Kabaw Fault zone. Black strike-and-dip symbols indicate outcrop measurements, white symbols near the eastern outcrops
indicate orientations calculated from the hillshade digital elevation model using the three-point method. Global Navigation
Satellite System (GNSS) stations used in this study are shown with blue triangles. Map abbreviations: KF-A, Model A Kabaw
Fault; KF-B Model B Kabaw Fault. (b) Chin Hills (left) and Chindwin Basin (right) stereograms show poles to bedding,
Kamb contour intervals are plotted at the three-sigma level. The cylindrical best fit and pole are shown with a great circle
and blue square, respectively. The Kabaw Fault Zone stereograms (middle) present minor fault slip datasets collected from
the easternmost Chin Hills. Great circles and arrows show orientations of fault planes and slickenlines, respectively, with
the arrow giving the motion of the hanging wall. Blue and red circles show the orientation of incremental shortening and
extension axes, respectively (Marrett & Allmendinger, 1990). Bulk strain axes (e1, e2, and e3) were calculated with Bingham
Statistics (Allmendinger et al., 2012).

shortening) for clusters of P and T axes using Bingham statistics (Figure 3; Allmendinger et al., 2012; Marrett &
Allmendinger, 1990).

Cross sections A and B were constructed perpendicular to the fold axis of the Kabaw Hills monocline (Figure 3).
Geologic map data were projected to determine the locations of fault ramps and axial surfaces of fault-bend folds.
Forward modeling with the open-source structural geology software FaultFold (Richard W. Allmendinger, 1998;
Zehnder & Allmendinger, 2000) and fbfFor (Connors et al., 2021) and manual cross-section construction tech-
niques were used to interpret the kinematic evolution, tectonic shortening magnitudes, and ramp depths for each
Kabaw Fault model. A structural model was first constructed manually from the map data. Then, the kine-
matic parameters for each forward model (e.g., slip and ramp dip) were adjusted until the model limb dips, axial
surfaces, and fault and contact locations matched the dip data and geologic map. Two viable kinematic models are
presented (Models A and B) and the geological and geodetic implications of each model are discussed (Figure 4).

3.2. Structural Analysis and Kinematic Models of the Kabaw Fault and Chindwin Basin

Structural analysis of the Chindwin basin constrains a broad north-trending syncline cored by the
Pliocene-Pleistocene Irrawaddy Formation. The western flank of the syncline, within the Kabaw Hills, exposes
a relatively continuous Cretaceous-Pliocene stratigraphic succession of marine and fluvial deposits that is 17 km
thick (Figure 3a; cf. Westerweel et al., 2020). A small thrust fault is interpreted at the base of the Cretaceous
section along the western flank of the syncline and a second small thrust is inferred to uplift the Eocene-Miocene
section along the eastern flank, explaining the topographic relief in each location. Asymmetric, west-vergent
minor folds were observed within shale horizons in the Cretaceous and Eocene units along the western flank.
The fold vergence is consistent with the sense of slip along the thrust fault at the base of the section and probably
records flexural shear within these units.

West of the Kabaw Hills, the topography is covered by alluvial deposits from the Nayintaya River near Kale
(Figure 3a). Jurassic to Early Cretaceous Ultramafic massifs (peridotite, ca. 127 Ma; Liu, Chung, et al., 2016;
Liu, Zhang, et al., 2016) crop out along with western margin of the Kale valley and are tectonically juxtaposed
with Triassic marine deposits of the Pane Chaung Formation as well and Jurassic-Eocene age Indo-Burman
flysch deposits in the Chin Hills. The contact between the ultramafic rocks and Indo-Burman flysch is defined

ORYAN ET AL. 5 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 4. (a) Kinematic model of the Kabaw Fault showing the wedge-thrust interpretation (Model A). An interpretation
of a reflection seismic profile from the Chindwin Basin north of the study area near latitude 23.7°N is presented below at
the same scale as the geologic cross sections in A and B (Böker et al., 2019). (b) Kinematic model showing the deep listric
Kabaw Fault interpretation (Model B). The projections of bedding dips are shown on the topographic profile with thin black
lines. Right-transpressional shearing within the Triassic Pane Chaung Fm. and Jurassic-Cretaceous Indo-Burma Flysch
is shown schematically. The slip magnitude for each fault segment is annotated. The cross sections overlay a crustal-scale
tomography profile (>5 km depth, Zhang et al., 2021) showing local seismicity (white circles, Engdahl et al., 1998; Mon
et al., 2020). In B, the Moho depth is shown after CRUST 1.0 (Laske et al., 2013). Models A and B are presented at the same
scale as the geologic map in Figure 3.

by a several kilometer wide shear zone. The Triassic-Eocene strata are intensely folded by north-trending, gener-
ally west-vergent, asymmetric, refolded tight to isoclinal folds (Chin Hills stereogram in Figure 3b; cf. Morley
et al., 2020; Yao et al., 2021). Minor fault-slip data collected from this shear zone form two kinematically distinct
populations (Marrett & Allmendinger, 1990). North striking minor thrust faults dip both east and west and record
east-trending subhorizontal shortening (n = 15; Figure 3b). A younger population consists of northwest strik-
ing right-reverse faults that dip southwest and contain southeast plunging slickenlines (n = 12; Figure 3b). The
right-reverse faults cut the earlier folds and thrust faults and record south-southwest trending subhorizontal short-
ening, consistent with the modern India-Shan convergence direction.

A viable structural model of the Kabaw Fault must explain several geological observations: (a) uplift of
Jurassic-Cretaceous ultramafic rocks (peridotite) of the Western Ophiolite Belt between Kale and the Chin Hills

ORYAN ET AL. 6 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

(Liu, Chung, et al., 2016; Liu, Zhang, et al., 2016; Morley et al., 2020); (b) uplift and eastward tilting of the
Cretaceous (Kabaw Formation) to Pliocene (Irrawaddy Formation) stratigraphic section exposed in the Kabaw
Hills as well as uplift and westward tilting of the Eocene-Miocene stratigraphic section near the eastern outcrops
of the Chindwin Basin; (c) the presence of basinward-dipping thrust faults imaged in reflection seismic data that
bound the east and west sides of the Chindwin Basin, and (d) pervasive right-transpressional shearing observed
within the Triassic Pane Chaung Fm. and Jurassic-Cretaceous Indo-Burman Flysch in the Chin Hills, structurally
below the Western Ophiolite Belt. The kinematic models may also attempt to explain: (e) patterns of crustal seis-
micity beneath the Chindwin Basin (cf. Mon et al., 2020), and f) the structure of the Burma crust as constrained
by seismic wave velocity profiles (Zhang et al., 2021). We present two geologically viable kinematic models of
the Kabaw Fault and Chindwin Basin that attempt to satisfy these constraints and we discuss the geological and
geodetic implications of each.

Model A (Figure 4a) considers the Kabaw Fault as a wedge thrust. East dipping strata in the Kabaw Hills are modeled
as the forelimb of a deeper fault-bend fold that forms above a west-dipping ramp. The deep ramp uplifts ultramafic
rocks of the Western Ophiolite Belt and transfers slip into the Chindwin Basin at the base of the Cretaceous section.
The dip of the deep ramp is constrained by the dip of the fault-bend fold forelimb (Suppe, 1983), which partially
tilts the Cretaceous-Pliocene strata to the east. The Kabaw Fault is modeled as a back-thrust (i.e., wedge thrust) that
splays off the upper flat of the fault bend fold, uplifting and further tilting the Cretaceous-Pliocene section. A small
hanging wall splay thrust is interpreted to explain the steepening of the bedding dip in the Eocene part of the section.

Some of the slip in Model A is also absorbed by a forethrust located at the eastern margin of the Chindwin Basin.
The forethrust thrust is required to uplift Eocene-Miocene section that crops out in this location (Tun et al., 2014)
and has been imaged in reflection seismic data near latitude 23.7°N, to the north of our study area (i.e., the Mattaung
thrust in Figure 2b and Böker et al., 2019). The central fault-related anticline imaged in the seismic data (Incaw
anticline), however, does not appear to extend southward to the latitude of our cross-section based on map relation-
ships and it is not recorded by surface data along our section. The minimum slip required to uplift the ultramafic
rocks to the surface above the deep ramp is 45 km, with 21 km absorbed by the Kabaw back-thrust, 4.5 km by the
splay fault, 12 km by the eastern forethrust, and the remaining 7.5 km is taken up by the fault-bend fold.

Model A (Figure 4a) produces a good fit to the Chindwin Basin bedding data and explains the uplift and tilting of
the strata on either side of the basin. The shallow ramps closely resemble the basin-bounding thrust faults imaged
in reflection seismic data to the north (Figure 4). Deformed mélange units that have been documented near the
contact between the ultramafic rocks and Cretaceous Kabaw shale (Morley et al., 2020) are also consistent with
shearing along this contact.

Model B considers the Kabaw Fault to be an east-dipping crustal scale listric reverse fault that uplifts the ultramafic
rocks of the Western Ophiolite Belt in its hanging wall (cf. Morley et al., 2020). In this interpretation, the Kabaw
Fault is defined as the western contact between the Western Ophiolite Belt and the Triassic Pane Chaung Fm. The
ultramafic rocks are thrust westward in the hanging wall of the Kabaw Fault and juxtaposed structurally above the
sheared Triassic–Cretaceous deposits (Tun et al., 2014) of the Chin Hills in the footwall. The deep listric thrust
geometry of the Kabaw Fault in Model B is projected from surface bedding dip data to approximate patterns
of crustal seismicity that have been located using double-difference methods (Mon et al., 2020) and match the
crustal structure of the Chindwin Basin that is constrained by earthquake tomography (Vp; Zhang et al., 2021)
near latitude 23°N (Figure 4b). The eastern thrust was also constructed from surface bedding dip data to match
the geophysical data below the eastern part of the Chindwin Basin. Axial surfaces projected from the surface data
constrain the shape of the listric faults and indicate that they extend to ∼30 km depth, near the base of the Burma
crust (Mon et al., 2020). The modeled eastern thrust projects ∼14 km east of that imaged by reflection seismic
below the eastern outcrops of the Chindwin Basin to the north (Böker et al., 2019), although it does not crop out
in the central Burma Basin which is buried by Quaternary alluvium. This location for the eastern thrust is taken
from Mon et al. (2020), who show that seismicity associated with the eastern thrust steps ∼15–20 km eastward
toward the Ye-U fault system near latitude 23°N. The forelimb axial surface on the Model B Kabaw Fault requires
a minimum slip of ∼50 km to uplift the Cretaceous section to the correct location in the Kabaw Hills. The eastern
fault requires at least ∼13 km of slip to uplift the Eocene-Miocene section to the surface.

Model B accounts for the juxtaposition of the Western Ophiolite Belt above the deformed Triassic-Cretaceous
strata of the Chin Hills and west-vergent transpressional shearing within the Triassic-Eocene Pane Chaung Fm.
and Indo-Burman flysch in the immediate footwall (cf. Betka, Seeber, Thomson, Sincavage, et al., 2018; Morley

ORYAN ET AL. 7 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

et al., 2020). Model B is also a good match to the crustal velocity profile and seismicity below the Chindwin
Basin, which are both independent datasets from our geologically constrained fault model (Figure 4b; Zhang
et al., 2021). Model B, however, is not as close a match as Model A to the structures that have been identified
in reflection seismic images within the Chindwin Basin, although Model B does not necessarily preclude these
basin scale structures, which are located in a different part of the Chindwin Basin than our study area. In Model B,
shallow thrusting within the Chindwin Basin is not required; however, evidence for flexural slip shearing within
shale deposits in the Cretaceous-Eocene section was observed and may explain the topographic relief at the
western front of the Kabaw Hills, as well as the thrust faults imaged at the margins of the Chindwin Basin farther
north (Böker et al., 2019). Deformed mélange units located at the top of the Western Ophiolite Belts are also
consistent with shearing along this contact, as in Model A (Morley et al., 2020). Both models yield a similar total
slip (45 and 63 km for Models A and B, respectively) with only ∼12–13 km of slip absorbed by the eastern fault,
indicating that significant tectonic shortening (∼50 km) has been absorbed by the Kabaw Fault. In Section 5, we
present the best fits to the GNSS data (presented in Section 4) for both Model A and B Kabaw Fault geometries.

4. GNSS Rates in the Indo-Burma Subduction Zone


4.1. GNSS Data

The Global Navigation Satellite System (GNSS) enables precise measurements of motion of geodetic networks
over an extended period of years to estimate rates of tectonic deformation as well as subsidence or uplift on the
order of ±1 mm/yr. We processed all available GNSS data in the region of the Indo-Burma subduction from
continuous stations deployed jointly by the Lamont-Doherty Earth Observatory (LDEO) and Dhaka Univer-
sity between 2003 and 2019 (Figure 2) and by the LDEO and the Myanmar Department of Meteorology and
Hydrology (DMH) in 2018. We also processed data from sites that were installed and maintained by the Earth
Observatory of Singapore, the Myanmar DMH, the Myanmar Earthquake Committee, the Geological Survey of
Bangladesh (R. Mallick et al., 2019), and the French IRD through the Belmont Forum BanD-Aid project (Shum
et al., 2014).

Daily continuous RINEX (Receiver Independent Exchange Format) files from these stations and stable Interna-
tional GNSS Service stations available in the region of SE Asia were processed using GAMIT/GLOBK (Herring
et al., 2010a, 2010b) to produce time series in the IRTF2014 reference frame (Altamimi et al., 2016). We removed
the horizontal seasonal signal from all stations' time series and generated velocities in the north and east direc-
tions employing a correlated noise approach using a first-order Gauss-Markov extrapolation model with white
noise (Herring, 2003). In addition to these data, we also included previously published GNSS velocities from
stations in Myanmar (Maurin et al., 2010) and India (Gahalaut et al., 2013). Similar to Steckler et al. (2016), we
stabilized our network with the same set of stations used by Gahalaut et al. (2013) to minimize reference bias
between our rates and those that were recorded by Gahalaut et al. (2013) and Maurin et al. (2010) and converted
their rates to ITRF2014. Moreover, we utilized a Helmert Transformation to align the realizations. We note that
the difference between realizations for these stations is minor. All in all, we employ displacement rates from 68
stations across the Indo-Burma subduction zone.

We converted these velocities (Table S1) to the Indian plate reference frame using the Steckler et al. (2016)
Euler pole (Figure 2 and Table S3 in Supporting Information S1). We then projected the horizontal velocities to
the direction perpendicular to convergence, assuming east-west trending convergence (e.g., Panda et al., 2020).
We did not include data from stations affected by local tectonics, such as uplift of the Shillong Massif along
the Dauki Fault (e.g., Steckler et al., 2016). We also did not include data from the HAKA site installed in the
IBR. This station is considerably south of the transect of GNSS stations installed near the Kabaw Fault, and its
recorded displacement rates are quite different from other rates along the profile. We address this behavior in the
discussion section.

4.2. Horizontal and Vertical GNSS Rates

Typically, determining reliable horizontal rates requires at least 2.5 years of GNSS data (Blewitt & Lavallée, 2002).
The Indo-Burma region is characterized by considerable seasonal variation due to the monsoon climate, which
could increase the minimum duration for consistent displacement rates (e.g., Chaumillon et al., 2017; Chiu &
Small, 2016; B. Mallick et al., 2011). Our stations across the Kabaw Valley and IBR were deployed in late 2018, so

ORYAN ET AL. 8 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 5. Horizontal and vertical Global Navigation Satellite System (GNSS) time series and rates. (a) Vertical GNSS time series and corrected rates for water loading.
Colored dots indicate daily relative vertical GNSS position. Curves denote long-term rates and seasonal trends from water loading estimates. (b) Variation in calculated
displacement rates with time for recently deployed stations. Panels show the difference between the final displacement rate to those computed with data available only
up to a given time before the present. Upper and lower sections denote variation in the east and north directions, respectively. Displacement rate estimations include
annual and semiannual seasonal fluctuations. Three small plots show 18 years of water level in few wells (red) and river gauge (black) sites used in the water loading
correction. Map: Circles and rectangles show the location of horizontal and vertical GNSS sites in the figure, respectively. Red and black dots mark the location of
water wells and river gauges used to estimate the effect of monsoon waters on the crust (see Figure S2 in Supporting Information S1 for the position of all gauges). Blue
area marks the surface of the Kaptai reservoir. Red curve marks the national border. Gray lines show fault traces.

it is conceivable that their time series are not extensive enough. We tested if the long-term tectonic rates observed
by these stations varied significantly with additional data and concluded that all our new stations contain a suffi-
ciently long time series to estimate the horizontal displacement rates (Figure 5b). Horizontal displacement rates
in the Indian reference frame for all stations that we used in our analysis are displayed in Figure 2.

Due to the higher uncertainty in the vertical component of GNSS data, the minimum time to get a reliable
vertical rate from a GNSS-derived time series is longer (e.g., Blewitt & Lavallée, 2002). Moreover, the vertical
component is more sensitive to seasonal fluctuations and affected by non-tectonic subsidence effects such as
water extraction and sediment compaction (e.g., Grall et al., 2018; Steckler et al., 2022). Due to these reasons, we
exclude vertical rates from stations deployed in the subsiding Ganges-Brahmaputra Delta (Steckler et al., 2022)
and near large cities where there is significant groundwater pumping and ultimately employ vertical rates from
8 stations only. Fortunately, the transition from a fully locked to a freely creeping subduction interface generates
a strong vertical signal, so even a few stations above where the interface transitions from full creep to interseis-
mic standstill can provide a valuable indication of the detachment locking position and displacement mechanism
(e.g., Lamb, 2021). Because of the importance of the vertical signal, we removed seasonal oscillations from the
GNSS time series by computing the expected elastic loading effect of monsoon waters on the crust, as described
in Section 4.3 (Figure 5).

4.3. Vertical GNSS Corrections

GNSS vertical time series show a significant seasonal component with an amplitude of up to a few cm/yr in the
region (Figure 5a and Table S2 in Supporting Information S1; Steckler et al., 2010). Downward motion during
the summer monsoon is due to the loading by seasonal flooding and recharge of groundwater during the intense
summer monsoon, representing lithospheric-scale elastic deformation from an average load of ∼100 × 10 9 tons
of water (e.g., Materna et al., 2021). Upward motion during the dry season represents the opposite phenomenon.

To correct this, we used data from 304 river gauges and 1,221 groundwater well measurements of the water table
to create a gridded surface of daily water levels in Bangladesh. We then calculated the surface deformation caused

ORYAN ET AL. 9 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

by the weight of the water using a model that simulates the response of a uniform elastic half-space to a rectangu-
lar load (Becker & Bevis, 2004). We adjusted the Young's modulus parameter to find the best fit for the vertical
motion at GNSS sites used to determine vertical rates. The residual of these fits represents the long-term vertical
tectonic motion (e.g., Steckler et al., 2010, 2022). Figure 5a illustrates the differences between the calculated and
observed vertical motions at these sites, which were used to determine these trends.

Lake Kaptai is a large water reservoir with seasonal fluctuations in water level that is located near several of the
GNSS stations used in this study (Figure 5). Unfortunately, we do not have water level records from the lake itself,
only from a few wells nearby. However, much of the lake seasonal variation is accounted for when computing
water level surface described above. Calculations of surface deflection from the Lake assuming an 8 m seasonal
water level change (Nuruddin & Hai, 2003) suggest a maximum annual contribution to the seasonal vertical elas-
tic deformation at the nearby stations is less than 3 mm. Any unaccounted-for water load could result in a small
decrease in the best fit Young's modulus for the closest GNSS stations, but would have a negligible effect on the
long-term vertical rate estimations.

Finally, our analysis did not include vertical rates measured near large concentrations of population, such as
Dhaka, and within the GBD as water pumping and subsidence due to sediment compaction, among other reasons,
overshadow the vertical tectonic signature (e.g., Grall et al., 2018; Morris et al., 2003; Steckler et al., 2022).

5. Geodetic Modeling of the Indo-Burma Subduction Zone


5.1. Geodetic Modeling of Fault Slip Rates

We computed the horizontal and vertical surface displacement rates associated with locked thrust faults using
the finite element numerical code PyLith (Aagaard et al., 2017) with triangular meshes generated by CUBIT
(Blacker et al., 1994). PyLith is an open-source code capable of simulating crustal deformation in a variety of
settings ranging from meters to hundreds of kilometers and milliseconds to thousands of years in purely elastic
and elastic-plastic rheologies. We refer readers who would like to use PyLith to its insightful manual and provide
an example mesh, a script to generate it, and Pylith input files to compute the surface deformation associated with
the fault defined in the mesh (Files S1 and S2).

To model the surface deformation associated with interseismic motion along a fault, we employed a purely elastic
rheology with a sufficiently wide (1,800 km) and deep (1,000 km) 2D domain with a uniform Poisson's ratio of
0.25. We used PyLith's fault interface module to model dislocation across fault surfaces in which only the deep
portion of the fault accommodates motion during the interseismic period. The megathrust fault was seeded in
the center of the domain and its displacement was computed for eleven thousand planar and non-planar faults
with varying curvature, steepness, locking depth, and down-going plate thickness (Kanda & Simons, 2010). We
assumed that the shallower portion of the interface of each fault is fully coupled, and its deeper portion is creeping
steadily at a constant rate (Almeida et al., 2018; Lindsey et al., 2021). For more details, please see Text S1 and
Figures S3–S5 in Supporting Information S1.

We model the Kabaw Fault using our two geologically viable structural interpretations of the Kabaw Fault and
the surrounding region (Section 3). The first structural model posits that the Kabaw Fault is an upper crustal
listric wedge-thrust that is rooted in the Cretaceous section at the base of the Chindwin Basin. The second model
considers the possibility that the Kabaw Fault is a larger structure that cuts and uplifts the Burma crust and is
creeping near the Moho (Figure 4). To account for the structure of the fault and its motion on both a north-south
(strike-slip) and east-west (convergence) orientations, we constructed a 2.5D domain using Pylith and CUBIT
with various locking depths assuming only structures below the locking depth are active during the interseismic
period (Figure S3 in Supporting Information S1; Aagaard et al., 2017; Blacker et al., 1994).

5.2. GNSS Fit Between Predicted Displacement Rates and Measured GNSS Rates

We found the best-fitting fault geometry and displacement rate using a grid-search approach to minimize the
misfit between the measured and modeled surface convergence rates. To estimate the total surface deformation,
we considered dip-slip motion on both the modeled subduction interface and the Kabaw Fault and computed the
goodness𝐴𝐴of fit, 𝐴𝐴 2 , for all fault configurations. We assumed normal distribution and𝐴𝐴used 𝐴𝐴 2 to compute the prob-
ability density function and uncertainties associated with the inversion. We examined the best fitting scenario as

ORYAN ET AL. 10 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

well as two subsets of solutions where we limit the possible outcome to (a) a single detachment so convergence
on the Kabaw Fault is negligible (reference model 1) and (b) to a locking depth shallower than 3 km. The latter
corresponds to a creeping detachment (reference model 2) in accordance with the upper limit of the geologically
constrained depth for the interface (Betka, Seeber, Thomson, Steckler, et al., 2018).

It should be noted that we did not attempt to fit the behavior of faults on the eastern side of the Chindwin Basin
due to the scarcity of geodetic data there. We also excluded the behavior of the CMF and Sagaing faults which
previous work have covered with little remaining controversy (Gahalaut et al., 2013; R. Mallick et al., 2019;
Maurin et al., 2010; Maurin & Rangin, 2009; Nielsen et al., 2004; Socquet et al., 2006; Steckler et al., 2016;
Vigny, 2003; Wang et al., 2011).

Our inversion utilizing GNSS horizontal rates indicates a megathrust locking depth𝐴𝐴 of 14 ± 10 km with conver-
gence rates
𝐴𝐴 of 5.7 ± 4.1
𝐴𝐴 and 11.6 ± 5.4 mm/yr on the Kabaw Fault and detachment, respectively. The best fitting
solution suggests a megathrust geometry that follows the top of the intra-slab seismicity (Figure 6), the expected
position of the interface separating the down-going and upper plates. Moreover, it agrees with the position of the
top interface boundary that is constrained by receiver functions and the anomaly in P-wave velocities derived by
a tomography survey and interpreted to indicate the location of the down-going Indian plate (Figure 6; Zhang
et al., 2021; Zheng et al., 2020).

The best fitting solution for the subset of results where shortening in the Indo-Burma subduction is limited to a
single detachment (reference model 1) indicates a megathrust locking depth 𝐴𝐴 of 30 ± 12 km with a convergence
rate
𝐴𝐴 of 17.7 ± 4.8 mm/yr. This solution offers inferior fit to the GNSS data and is 290 time less likely than the
best fitting scenario discussed above. The inconsistency between the projected megathrust fault geometry and
the observed intraslab seismicity distribution and interface position, as indicated by receiver functions and tomo-
graphic surveys (Figure 6), underscores the poor fit of this model.

The solution found most probable for a creeping megathrust (reference model 2) suggests a convergence rate of
9 mm/yr, which is slightly higher than the value suggested by Panda et al. (2020). The inconsistency between
the measured and predicted creep rates is quite substantial (Figure 6), leading to a solution 𝐴𝐴
that is 106 less likely
than the best fitting solution. It is worth noting that when we do not consider the Kabaw Fault, the likelihood of a
creeping detachment drops𝐴𝐴 to 10−8 in comparison with the most probable fault configuration. Moreover, the shape
of the freely slipping detachment does not match the intraslab seismicity and other seismic constraints (Figure 6).

Solutions within one standard deviation from the best fitting result point to a similar trend as those discussed
above (Figure 7). One noticeable feature is that none of the creeping solutions align with the seismic constraints.
Moreover, the steeper fault configurations within one standard deviation from the best fitting solution show
similarities with reference model 1 and do not agree with the seismic signature of the slab. In contrast, the gently
sloping fault geometries linked with the best solution suggest shallower locking depth and lower megathrust slip
rate. These faults are in better agreement with the position of the slab (Figure 7) and indicate that the locking
depth of the most accurate fault configuration is probably on the lower end of our uncertainties and is possibly
even shallower than 14 km. This is a reasonable estimate in comparison with geological constraints suggesting
a detachment depth of ∼7 km at that location (92.4 E; Betka, Seeber, Thomson, Sincavage, et al., 2018; Betka,

𝐴𝐴
Seeber, Thomson, Steckler, et al., 2018). It is also a significant improvement compared to the single detachment
model (ref. model 1), and previous geodetic work, which used a planar megathrust and predicted a locking depth
>25 km (R. Mallick et al., 2019; Steckler et al., 2016).

We note that the inversion combining both horizontal and vertical rates produces similar results to these yielded
by the inversion relying on horizontal rates only (Figures S7 and S8 in Supporting Information S1); however,
our preferred inversion is based on the latter. Our preference is based on two factors: (a) low signal-to-noise
ratio associated with vertical GNSS data and (b) a peak in vertical displacement rates appear only when the
locked-to-creeping transition is a step function. Wide zone of transition in coupling will generate a gentle and
broad uplift signal (Kanda & Simons, 2010). Nonetheless, the significance of the vertical velocities can be
observed in Figure 6, where the mismatch of reference model one is particularly apparent considering the vertical
field.

Our network of GNSS stations encompasses a region approximately 1,000 km wide and extends to latitudes
between 22° and 25° N. Our analysis is based on the premise that the slab geometry remains constant along strike
and that the Earth's curvature does not result in rotational motion. To verify that our 2-D analysis is not skewed

ORYAN ET AL. 11 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 6. Best fitting solutions. (a) Solutions for horizontal fits in three directions. Colored circles, triangles, and squares denoted Global Navigation Satellite System
(GNSS) velocities in perpendicular, parallel, and vertical orientations, respectively. Blue, Green, magenta, black, and red velocities show horizontal GNSS stations
installed by LDEO, EOS (R. Mallick et al., 2019), Gahalaut et al. (2013), Band-Aid (Shum et al., 2014), and Maurin et al. (2010). All velocities are in the frame
reference of the Indian plate with positive perpendicular and parallel velocities directed in west and south directions, respectively. Empty symbols mark stations that
were not included in the analysis. Error bars show a confidence level of 95%. Colored curves show solutions for various fits with the red curve showing model B for
the Kabaw Fault. (b) Vectors of motion (parallel rate, perpendicular rate) across the Indian-Shan faults. Dextral motion on the Sagaing and Churachandpur-Mao Fault
(CMF) faults were modeled by previous studies (Gahalaut et al., 2013; R. Mallick et al., 2019; Maurin et al., 2010; Nielsen et al., 2004; Socquet et al., 2006; Steckler
et al., 2016; Vigny, 2003; Wang et al., 2011). (c) Schematic cross-section of Indo-Burma convergence zone𝐴𝐴along 23.2 N. Colored curves denote modeled faults. Thick

dashed line marks the position of CMF, which we did not simulate. Thin dashed curves mark the possible position of the locked portions of our modeled faults. Green
(slow) and yellow (fast) colors (scale in Figure 4) denote P-wave velocities derived by tomography survey𝐴𝐴along 23 N with LV1 mark the interpreted position of the

subducting Indian crust (Zhang et al., 2021). The surface of the topography cross-section is overlain by colored dots marking the stream-scale normalized steepness
index with higher values (red), suggesting faster uplift rates (Maneerat & Bürgmann, 2022). The yellow triangle marks the location of the volcanic arc. Locking
positions are denoted by black circles. Black empty circles denote local seismicity near the profile (Engdahl et al., 1998; Mon et al., 2020). Thick dashed orange lines
mark the position of the interface derived by receiver functions𝐴𝐴along 22 N (Zheng et al., 2020). IBR, Indo-Burman Fold-belt; KF, Kabaw Fault; SF, Sagaing Fault;

MT, Indo-Burma megathrust; BARK, HAKA, and TEDM are GNSS stations discussed in the main text. Modified after Murphy (1988), Steckler et al. (2008, 2016),
and Wang et al. (2014).

by rotational motion, we created a simple 3-D block model (Meade & Loveless, 2009) accounting for the large
faults in the region of our analysis and five different microplates. We used the GNSS horizontal data described
in Section 4 in the 2014 IRTF projection and inverted for motion on the faults. The 3-D block model generates
velocities similar to those described above and thus verifies our 2-D analysis. For more details, please see Text
S2 and Figure S9 in Supporting Information S1.

Finally, it is interesting to note that site BARK in Barkal, Bangladesh (Figure 2) is recording faster uplift rates in
comparison to other sites and our preferred estimates of uplift (Figure 6). This station is located east of Kapati

ORYAN ET AL. 12 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Figure 7. Solutions within one standard deviation from the best fitting results. Panels (a–c) show the horizontal and vertical
fits as well as predicted geometry for the best fitting model including loading on the megathrust and Kabaw Fault. Panels d
and e show the predicted geometries for reference models 1 and 2, respectively. Solutions are color coded by their distance
from the most probable solution, with black indicating high probability. Black circles show locking point. Yellow circles
denote local seismicity (Engdahl et al., 1998; Mon et al., 2020) and thick dashed orange lines mark the position of the
interface derived by receiver functions𝐴𝐴along 22 N (Zheng et al., 2020).

Lake (Figures 2 and 5), a hydroelectric power reservoir created by filling two synclinal valleys between anticlinal
ridges (e.g., BPDB, 1985; Karmakar et al., 2011). This region has been recording enhanced seismicity with fairly
large events (Mw 5.6 Barkhol, 2003; Mw 4.9 Barkhol, 2007) that resulted in extensive damage and a few fatalities
(Alam et al., 2006; Steckler et al., 2008). The continuing seismicity is possibly reservoir-induced (e.g., Simpson
et al., 1988), giving rise to increased slip on the thrust fault rooting the anticline, which could accelerate the
inelastic growth of the anticline (e.g., R. Mallick et al., 2021) and account for the rates recorded.

5.3. Geodetic Fit to Geological Constrained Kabaw Fault Models

We used two geologically derived fault models (A and B; Section 3, Figure 4) to fit the GNSS data and describe
the motion on the Kabaw Fault. In addition to the inversion associated with dip-slip motion (Section 5.2), we also

ORYAN ET AL. 13 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

inverted for north-south (strike-slip) motion on the Kabaw Fault using GNSS stations in its proximity (Figure
S1 in Supporting Information S1). The convergence and strike-slip inversions indicate that Model B is 2.8 times
more likely than Model A, partly due to the slip required on the deep, western-dipping fault-bend fold ramp
needed for Model A (Figure S10 in Supporting Information S1 and Figure 4a). The best fit for Model A suggests
a locking depth of 12 𝐴𝐴
km (∼95.65 E) and a right-lateral strike-slip and convergence rate
𝐴𝐴 of 7.8 and 3.3 mm/yr,

respectively (Figure S10 in Supporting Information S1). The geodetic fit for Model B (Figure 4b) produces a
locking depth of 22 𝐴𝐴
km (∼95.65 E), and a right-lateral strike-slip and convergence rate
𝐴𝐴 of 8.4𝐴𝐴 and 5.71 mm/yr,

respectively (Figure 6). We note that the inversion for the strike-slip motion was more sensitive to the Kabaw
Fault locking depth, and yet we purposely did not ascribe a locking depth uncertainty as it depends on the kine-
matic models of the fault. Moreover, we assign the uncertainty associated with the block modeling (3 mm/yr;
Text S2 in Supporting Information S1) to the strike-slip component of the Kabaw Fault as it accounts for motion
on the Sagaing and CMF faults. Finally, based on the geodetic results presented here, Model B is a better fit to
the Kabaw Fault GNSS data.

6. Discussion
6.1. Revised Megathrust Geometry and Loading Rate

Using updated horizontal and vertical GNSS rates, including recently installed sites in the IBR and near the
Kabaw Fault, we are able to demonstrate how convergent strain is distributed between the megathrust and the
Kabaw Fault. Our megathrust analysis employs interseismic geodetic data sets that predominantly records deep
aseismic slip that cannot shed light on the geometry of the locked portion of the megathrust. Nevertheless, our
results suggest that the shift from a fully coupled to freely slipping megathrust detachment occurs at a depth of
10–20 km and a longitude of ∼92.4° E. While uncertainty remains as to the exact locking depth, our work spec-
ifies the shallowest part of the megathrust that is detectable with our interseismic data set is in reasonable agree-
ment to the depth estimates stemming from geological and seismic constraints (∼7 km; Betka, Seeber, Thomson,
Steckler, et al., 2018; Bürgi et al., 2021; Abdullah et al., 2022). This is an improvement from previous geodetic
work that determined locking depths of 25–35 km in a similar position (e.g., R. Mallick et al., 2019; Steckler
et al., 2016). We attribute this improvement to the addition of vertical rates and introducing non-planar faults in
our geodetic analysis. Although we cannot determine the shape of the locked portion of the megathrust, we note
that the curved geometries determined here are in good agreement with the geologically and geophysically esti-
mated structure of the megathrust (Abdullah et al., 2022; Betka, Seeber, Thomson, Steckler, et al., 2018; Bürgi
et al., 2021). Our preferred convergence rate (∼12 mm/yr) is on the lower end of values suggested by previous
geodetic work (R. Mallick et al., 2019; Steckler et al., 2016) as our models show that a portion of the Indian-Shan
shortening is also absorbed by the Kabaw Fault, reducing the convergence rate on the megathrust.

6.2. Geologic Significance and Cenozoic History of the Kabaw Fault

The map trace of the Kabaw Fault extends for >500 km from the northern ranges of the IBR to the Rakhine Hills of
southwest Myanmar (Morley et al., 2020). Along its entire length, the Kabaw Fault defines a prominent geomor-
phic boundary between the relatively high elevation, high-relief, and highly deformed Mesozoic-Paleogene stratig-
raphy of eastern ranges of the IndoBurman accretionary prism to the west, and the low-lying Mesozoic-Cenozoic
sub-basins of the central Burma basin to the east (Figure 2). The along-strike continuity and significance of
this fault system are recognized in active tectonic maps (Maurin & Rangin, 2009; Steckler et al., 2016, 2021;
Wang et al., 2014) and geomorphic indices of active deformation in the IBR (Maneerat & Bürgmann, 2022).
Our kinematic models predict ∼45–50 km of slip near the center of the Kabaw Fault, indicating that the fault
length is greater than the maximum displacement by a factor of roughly 10. This result is consistent with global
compilations of fault length-displacement trends (Cowie & Scholz, 1992; Schultz & Fossen, 2002) and suggests
that our geologic shortening estimates are reasonable.

The timing of Cenozoic slip along the Kabaw fault is partly constrained by growth strata imaged in seismic data
and low-temperature thermochronology. Interpretations of reflection seismic data in the Salin Basin (Figure 2)
indicate that eastward-tilted Miocene strata in the hanging wall of the Kabaw Fault are pre-kinematic deposits
with respect to thrusting along the Kabaw Fault (Figure 2; Pivnik et al., 1998). Assuming that the Miocene
deposits in the Chindwin Basin are also pre-kinematic with respect to thrusting on the Kabaw Fault, we estimate

ORYAN ET AL. 14 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

a geologic slip rate on the order of 10 km/Myr (10 mm/yr) to absorb ∼50 km of shortening since ∼5 Ma. This
estimated long-term rate is of the same order of magnitude as our modeled geodetic convergence rate across the
Kabaw Fault (∼6–7 mm/yr) and suggests that the Kabaw Fault has been an important part of the Indo-Burman
subduction zone since the Miocene.

Thermochronological (Najman et al., 2020, 2022) and stratigraphic (Westerweel et al., 2020) studies of the Chin
Hills and Chindwin Basin document two major regional unconformities and periods of exhumation during the
middle Eocene to Early Oligocene and again from the Late Oligocene to Early Miocene. The Eocene-Oligocene
event is recorded by exhumation of the Triassic Pane Chuang Formation that we infer may be explained by uplift
of the suture zone units in the hanging wall of the Kabaw Fault (Model B) during the accretion of the Burma
Terrane to India (Najman et al., 2022). The later Oligocene-Miocene event records the collision of the Burma
Terrane with the Eastern Himalaya (Najman et al., 2022; Westerweel et al., 2020) and a change in the dynamics of
the accretionary prism as more Himalayan sediment is delivered to the Bengal Basin (Betka et al., 2021; Najman
et al., 2022). Thus, it is likely that deformation along the Kabaw Fault Zone occurred in episodic tectonic events
throughout the Cenozoic (cf. Morley et al., 2020).

6.3. Evaluation of Kabaw Fault Kinematic Models

Kabaw Fault Model A aligns with regional tectonic interpretations which assume the Western Ophiolites are
thrust eastward onto the Burma margin above a west-dipping crustal-scale fault (Maurin & Rangin, 2009). The
deeper, west-dipping ramp in Model A may help to explain the higher relief and rates of vertical uplift in the
eastern ranges of the IBR (i.e., the Chin Hills) documented by stream-scale geomorphic analysis (Maneerat
& Bürgmann, 2022; Figure 6). Model A, however, does not adequately explain the deeper crustal structure
constrained by seismicity and tomography below the Chindwin Basin (Mon et al., 2020; Zhang et al., 2021).
Furthermore, Model A cannot readily explain the west-vergent transpressional shearing (Morley et al., 2020)
within the Triassic Pane Chaung Fm. which is located on the back-limb (west dipping limb) of the fault-bend
fold (Figure 4). Model A is also a poorer fit to geodetic data (i.e., Section 5.3) than Model B. The geodetic fit
for Model A can be improved compared to that of Model B, if we similarly consider that the deep, west-dipping
ramp is converging at a negligible rate in comparison with the IBR megathrust (cf. back-thrust at Chilean subduc-
tion zone; Lamb, 2022). Such a model partitions nearly all of the Kabaw Fault convergence onto the shallow
east-dipping Kabaw thrust that uplifts the Chindwin Basin. This kinematic scenario would, however, require a
second shallow detachment that extends eastward from Chindwin Basin, below the central Burma Basin, and
unrelated to the structure uplifting the rocks of the Western Ophiolite Belt. Such a regional structure has not
been previously proposed within Myanmar and seems less likely than our preferred geological interpretation
(Model B).

Model B considers the Kabaw Fault as an east-dipping structure that separates the suture zone from Mesozoic to
Cenozoic sedimentary rocks of the accretionary prism. This model implies that the Western Ophiolite Belt is part
of the Burman crust, which is underthrust by the accretionary prism. Model B is consistent with the crustal-scale
geophysical data and regional tectonic interpretations which infer that seismicity below the Kabaw Hills reflects
sediment subduction within the accretionary wedge (e.g., Bai et al., 2021; Mon et al., 2020; Wu et al., 2021).
While the eastward dip of the Kabaw Fault in Model B allows for sediment subduction, it cannot solely explain
observed high rates of uplift in the eastern ranges of the IBR which are located in the footwall of the fault (i.e.,
the Chin Hills; Maneerat & Bürgmann, 2022; Figure 6); however, other explanations are possible. For example,
Morley et al. (2020) describe a stronger component of right lateral shearing along the southern segment of the
Kabaw Fault than on its northern segment. This observation suggests that right-lateral slip along the CMF, to
the northeast of the study area (Figures 1 and 2), may be transferred across the Chin Hills to the southern Kabaw
Fault, forming a contractional left-stepover between the two right-lateral fault systems. This posited left stepover
may explain the ongoing uplift and deformation of the Chin Hills and account for the higher IBR perpendicular
rates recorded in TEDM and HAKA stations located within the Chin Hills (Figure 6) than the Indian GNSS sites
farther north. Thus, Model B provides the most simple explanation for most of first-order tectonic features in
the region, including the shape of the Chindwin Basin syncline, west-vergent right-transpressional shearing of
Triassic-Cretaceous strata in the Chin Hills, and the crustal structure of Burma as constrained by crustal geophys-
ics. Model B also produces a significantly better geodetic fit to the GNSS data without requiring novel geologic
interpretations, it is therefore our preferred interpretation for the Kabaw Fault.

ORYAN ET AL. 15 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

6.4. Seismic Potential of the Megathrust and Kabaw Fault

The addition of GNSS stations near the Kabaw Fault and its structural reconstruction helped us determine how
strain is distributed across the Indo-Burma plate boundary zone and demonstrate that the megathrust accumu-
lates strain at a slower rate than indicated before (R. Mallick et al., 2019; Steckler et al., 2016). Nonetheless,
a megathrust event could rupture from the poorly constrained location of the northern edge of the 1762 event
to the Shillong massif (∼300 km), and updip from the locking depth to the thrust front (∼100 km). Assuming
this event will release strain accumulated during the last 400–1,000 years in highly sedimented settings (shear
modulus ∼20 GPA) at a constant rate of ∼12 mm/yr, a megathrust earthquake of Mw 8.2–8.5 is plausible (Figure
S11 in Supporting Information S1). This reduces the hazard associated with a deeper locking depth previously
suggested (e.g., R. Mallick et al., 2019; Steckler et al., 2016) and yet, our geodetic analysis cannot resolve whether
such an event will rupture all at once, comparable to the 1700 Cascadia event, or if anticline splay faults will be
activated, as was as has been observed for various continental and oceanic events (Cummins & Kaneda, 2000;
Hubbard et al., 2010; Melnick et al., 2012; Satake et al., 1996; Wang et al., 2014; Yue et al., 2005). Moreover, we
cannot determine if this event will rupture past the thrust front to the deformation front like the 2011 Tōhoku-oki
and other tsunami earthquakes (e.g., Oryan & Buck, 2020) or will be confined solely to the deeper part of
the interface (Lay et al., 2012). Finally, a portion of the elastic interseismic deformation might be permanently
stored in the upper plate, reducing the overall strain available to drive earthquakes (e.g., Jolivet et al., 2020;
Madella & Ehlers, 2021; Meade, 2010; Oryan et al., 2021, 2022). Thus, a megathrust rupture could be smaller
or larger depending on the rupture area and time passed since the last earthquake (Figure S11 in Supporting
Information S1).

Our results also demonstrate the capacity of the Kabaw Fault to host large events. If we consider the along-strike
extent of the Kabaw Fault (>500 km) (Maurin & Rangin, 2009; Morley et al., 2020; Wang et al., 2014; Figure 2),
our fitted locking depth (22 km) and the fact that the region has not experienced a significant event at least in the
past century (Kundu & Gahalaut, 2012) we conclude that it is capable of hosting Mw 7–7.5 events. Our interpre-
tation aligns with stress inversion from local seismicity suggesting that the Kabaw Fault is actively shortening
(Earnest et al., 2021) and with oblique-thrust focal mechanisms in the area which are distinct from other crustal
and intraslab earthquakes (Fadil et al., 2023). In any case, lack of moderate interplate seismic activity and our
combined geodetic and structural analysis points to the potential of the Indo-Shan plate boundary to host large
earthquakes both on the Indo-Burma megathrust and the Kabaw Fault.

6.5. Comparison With Previous Work Indicating Indo-Burma Subduction Zone Is Aseismic

In contrast to our results, some previous work indicated that it is unclear whether the detachment is fully coupled
or freely slipping (Panda et al., 2020). This line of reasoning follows previous arguments claiming the Indo-Burma
subduction zone is incapable of hosting large megathrust events (Gahalaut et al., 2013; Kundu & Gahalaut, 2012;
Le Dain et al., 1984; Ni et al., 1989; Panda et al., 2022). These arguments lean on three lines of evidence
regarding the Indo-Burma interface: (a) it is considered to be a shallow dipping and weak structure (Abdullah
et al., 2022; Betka, Seeber, Thomson, Steckler, et al., 2018; Bürgi et al., 2021; Steckler et al., 2008), (b) it lacks
evidence for moderate interseismic interplate activity (Kundu & Gahalaut, 2012), and (c) seismic activity in the
slab is dominated by North-South P axes (Purnachandra Rao & Kumar, 1999). We address all of these arguments
in the following paragraphs.

First, as presented in Section 4.2, we do not find geodetic evidence to support the argument that the Indo-Burma
detachment is creeping as we do not observe a sharp offset in displacement rates in our profile (Figure 6). More-
over, the strength of faults is not associated with their capacity to host earthquakes. Extremely weak detach-
ments and continental faults (Gao & Wang, 2014; Hass & Harris, 2016; Lachenbruch & Sass, 1980; Oryan &
Savage, 2021), such as the south segment of the Chile subduction, have generated great earthquakes. Second, the
lack of recorded interseismic moderate interplate events is not necessarily an indication of shortening accommo-
dated through creep. The Cascadia subduction zone is a converging locked megathrust characterized by a high
sedimentation rate that shows very little moderate interplate activity but produced a M9 megathrust event in the
past (Satake et al., 1996). An additional example of this behavior lies much closer to our region of interest. The
M > 8 great Shillong earthquake ruptured on the Oldham fault, an area that is characterized by limited seismic
activity (Subedi & Hetényi, 2021; Figure 1). Additionally, other subduction zones that have hosted great megath-
rust events, such as the Andamans, Kyushu, and Cascadia, display similar slab and upper plate seismic signatures

ORYAN ET AL. 16 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

to those produced by the Indo-Burma subduction zone (R. Mallick et al., 2017, 2019; Seno & Yoshida, 2004).
Lastly, recent stress inversions based on earthquakes from the Indo-Burma plate boundary suggest the megath-
rust can produce large earthquakes in spite of the north-south stress orientation (Earnest et al., 2021; Maneerat,
Dreger, et al., 2022).
It has also been argued that the Indo-Burma subduction zone is creeping (Panda et al., 2020, 2022). Panda
et al. (2020) used GNSS rates from the Indo-Burma subduction zone employing a new reference frame and indi-
cated that the Indo-Burma detachment is potentially creeping. In contrast, our corrected vertical rates, which are
insensitive to the choice of frame of reference, suggest that convergence on a freely slipping interface is insuf-
ficient (Figure 6). Moreover, our analysis shows that the best fitting solution for a freely creeping detachment
does not fit the GNSS displacement rates and intra-slab seismicity. Modeling efforts also suggest the Indo-Burma
subduction zone is capable of hosting significant events. Vorobieva et al. (2021) modeled the seismic behavior
of the Indo-Burma subduction zone using block-and-fault dynamic simulations that allow motion on faults to be
accommodated both in stick-slip and slow creep fashion. They showed that a locked detachment model repro-
duces tectonic velocities, earthquake size distribution, and seismicity rate that are consistent with observations
from the Indo-Burma subduction zone. They concluded that the locked megathrust could generate great Mw 8+
events with an average repeat time of over 1,000 years.
Finally, the historical record might be the strongest evidence that the Indo-Burma region can produce large
earthquakes. Great earthquakes have ruptured the Indo-Burma subduction zone (Figure 1) with the 1762
Arakan megathrust event rupturing along the southern section of our study area with a tsunami affecting Dhaka
(Akhter, 2010; Steckler et al., 2008). The event caused relative sea-level changes, caused mud volcano eruptions,
and raised corals and marine terraces along the Myanmar and Bangladesh coasts (Akhter, 2010; Cummins, 2007;
Mondal et al., 2018; Steckler et al., 2008; Wang et al., 2013). The event also caused 1–2 m of subsidence of
salt kilns in the GBD (Hanebuth et al., 2021). Similar evidence for a 900 CE event has been observed as well
(Hanebuth et al., 2021; Mondal et al., 2018). The Cascadia subduction zone is converging about 2.5 times faster
than the Indo-Burma subduction and in the last 6000 years ruptured 12 times, indicating an average inter-event
of ∼500 years (Nelson et al., 2006; Schellart et al., 2011). If once again, we adopt the Cascadia subduction as
an analogous case, an inter-event time of >1,000 years for the Indo-Burma subduction zone is not unreasonable.

7. Conclusions
We present new GNSS horizontal and vertical displacement rates from the Indo-Burma region, including from
newly installed stations in the IBR and Myanmar. The vertical rates are corrected for seasonal hydrologic loading
using a dense network of well and river gauge stations in Bangladesh. We also present two viable structural inter-
pretations of the Kabaw Fault and surrounding region based on new field observations and existing seismic data.
Using these datasets, we demonstrate that the convergent strain is distributed between both the active and locked
Kabaw Fault, and the Indo-Burma megathrust. Our results show the potential of the Indo-Burma subduction zone
to host great earthquakes with the best fitting solution suggesting a locking depth
𝐴𝐴 of 14 ± 10 km and convergence
rate
𝐴𝐴 of 11.6 ± 5.4 mm/yr along a curved interface. The geometry and locking depth of our fitted Indo-Burma
interface agrees with geological evidence, the distribution of slab seismicity, and the position of the slab derived
by tomography and receiver functions. Our preferred model for the Kabaw Fault (Model B) indicates that it is an
important and active right-oblique reverse fault that cuts the Burma crust and absorbs
𝐴𝐴 8.4 ± 3.0 mm/yr of dextral
shear
𝐴𝐴 and 5.7 ± 4.1 mm/yr of convergence, and it is capable of hosting Mw 7–7.5 earthquakes.

Data Availability Statement


GNSS Rinex files can be found in the public UNAVCO repository (https://www.unavco.org/data/gps-gnss/
data-access-methods/gnss-data-access-notebook/gnss-data-access-notebook.html) and in R. Mallick et al. (2019).
Published GNSS rates can be found in Maurin et al. (2010) and Gahalaut et al. (2013). Rinex files were processed
using the free software GAMIT/GLOBK (http://geoweb.mit.edu/gg/) and tsview (http://wwwgpsg.mit.edu/∼tah/
GGMatlab/). Elastic solutions were computed using Pylith (https://geodynamics.org/resources/pylith) and CUBIT
(https://coreform.com/products/downloads/). GNSS rates were fitted to the elastic solutions using NumPy (https://
numpy.org/) and SciPy (https://scipy.org/). Structural forward modeling was calculated using the open-source
software FaultFold (https://www.rickallmendinger.net/faultfold). Figures were compiled using GMT (Wessel
et al., 2019), matplotlib (Caswell et al., 2021), and illustrator (https://www.adobe. com/products/illustrator.html).

ORYAN ET AL. 17 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Acknowledgments References
The authors are grateful for constructive
and helpful reviews by Rishav Mallick, Aagaard, B., Knepley, M., & Williams, C. (2017). PyLith v2.2.1. Computational Infrastructure for Geodynamics. https://doi.org/10.5281/
the Associate editor, and one anonymous zenodo.886600
reviewer whose comments have signif- Abdullah, R., Aurthy, M. R., Khanam, F., Hossain, M. M., & Sayem, A. S. M. (2022). Structural development and tectonostratigraphic evolution
icantly improved this manuscript. The of the Sylhet Trough (northeastern Bengal Basin) in the context of Cenozoic Himalayan Orogeny: Insights from geophysical data interpreta-
authors would also like to acknowledge tion. Marine and Petroleum Geology, 138, 105544. https://doi.org/10.1016/j.marpetgeo.2022.105544
prolific discussion with Roger Buck, Akhter, S. H. (2010). Earthquakes of dhaka. In Environment of Capital Dhaka—Plants wildlife gardens parks air water and earthquake
Romain Jolivet, and Arthur Olive, which (pp. 401–426). Asiatic Society of Bangladesh.
contributed to ideas presented here. The Alam, M., Alam, M. M., Curray, J. R., Chowdhury, M. L. R., & Gani, M. R. (2003). An overview of the sedimentary geology of the Bengal
authors would like to thank UNAVCO Basin in relation to the regional tectonic framework and basin-fill history. Sedimentary Geology, 155(3), 179–208. https://doi.org/10.1016/
(now Earthscope Consortium), Keith S0037-0738(02)00180-X
Williams, and Zaw Min of the DMH for Alam, M., Bhuiyan, A., & Islam, M. (2006). Seismic structural assessment of damaged chittagong public library building during 27 July 2003
supporting the installation of the new earthquake.
GNSS stations in Bangladesh and Myan- Allmendinger, R. W. (1998). Inverse and forward numerical modeling of trishear fault-propagation folds. Tectonics, 17(4), 640–656. https://doi.
mar. The authors also thank the schools, org/10.1029/98TC01907
DMH field offices, and Buddhist monas- Allmendinger, R. W., Cardozo, N. C., & Fisher, D. (2012). Structural geology algorithms: Vectors & tensors (p. 289). Cambridge University
teries hosting the GNSS sites. This work Press.
was supported by NSF EAR 17-14892. Almeida, R., Lindsey, E. O., Bradley, K., Hubbard, J., Mallick, R., & Hill, E. M. (2018). Can the updip limit of frictional locking on megathrusts
be detected geodetically? Quantifying the effect of stress shadows on near-trench coupling. Geophysical Research Letters, 45(10), 4754–4763.
https://doi.org/10.1029/2018GL077785
Altamimi, Z., Rebischung, P., Métivier, L., & Collilieux, X. (2016). ITRF2014: A new release of the International Terrestrial Reference Frame
modeling nonlinear station motions. Journal of Geophysical Research: Solid Earth, 121(8), 6109–6131. https://doi.org/10.1002/2016JB013098
Bai, Y., He, Y., Yuan, X., Tilmann, F., Ai, Y., Jiang, M., et al. (2021). Seismic structure across central Myanmar from joint inversion of receiver
functions and Rayleigh wave dispersion. Tectonophysics, 818, 229068. https://doi.org/10.1016/j.tecto.2021.229068
Becker, J. M., & Bevis, M. (2004). Love’s problem. Geophysical Journal International, 156(2), 171–178. https://doi.
org/10.1111/j.1365-246X.2003.02150.x
Betka, P. M., Seeber, L., Thomson, S. N., Sincavage, R., Steckler, M. S., Zoramthara, C., & Gahalaut, V. K. (2018). New geologic map, cross
sections, and structural models of the Indo-Burman accretionary prism. Paper presented at the AGU Fall Meeting 2018. AGU. Retrieved from
https://agu.confex.com/agu/fm18/meetingapp.cgi/Paper/447205
Betka, P. M., Seeber, L., Thomson, S. N., Steckler, M. S., Sincavage, R., & Zoramthara, C. (2018). Slip-partitioning above a shallow, weak
décollement beneath the Indo-Burman accretionary prism. Earth and Planetary Science Letters, 503, 17–28. https://doi.org/10.1016/j.
epsl.2018.09.003
Betka, P. M., Thomson, S. N., Sincavage, R., Zoramthara, C., Lalremruatfela, C., Lang, K. A., et al. (2021). Provenance shifts during Neogene
Brahmaputra delta progradation tied to coupled climate and tectonic change in the eastern Himalaya. Geochemistry, Geophysics, Geosystems,
22(12), e2021GC010026. https://doi.org/10.1029/2021GC010026
Blacker, T. D., Bohnhoff, W. J., & Edwards, T. L. (1994). CUBIT mesh generation environment. Volume 1: Users manual (No. SAND-94-1100).
Sandia National Laboratories. https://doi.org/10.2172/10176386
Blewitt, G., & Lavallée, D. (2002). Effect of annual signals on geodetic velocity. Journal of Geophysical Research: Solid Earth, 107(B7),
ETG9-1–ETG9-11. https://doi.org/10.1029/2001JB000570
Böker, U., Leo, C., & Sone, P. (2019). Structural and stratigraphic review of the Southern Chindwin Basin. MAPG Myanmar Petroleum Tech-
nology Conference.
Bouilhol, P., Jagoutz, O., Hanchar, J. M., & Dudas, F. O. (2013). Dating the India-Eurasia collision through arc magmatic records. Earth and
Planetary Science Letters, 366, 163–175. https://doi.org/10.1016/j.epsl.2013.01.023
BPDB. (1985). Karnaphuli hydro station. (p. 20). Bangladesh Power Development Board.
Bürgi, P., Hubbard, J., Akhter, S. H., & Peterson, D. E. (2021). Geometry of the décollement below eastern Bangladesh and implications for
seismic hazard. Journal of Geophysical Research: Solid Earth, 126(8), e2020JB021519. https://doi.org/10.1029/2020JB021519
Caswell, T. A., Droettboom, M., Lee, A., Andrade, E. S. D., Hoffmann, T., Hunter, J., et al. (2021). matplotlib/matplotlib: REL: V3.4.3. Zenodo.
https://doi.org/10.5281/zenodo.5194481
Chaumillon, E., Bertin, X., Fortunato, A. B., Bajo, M., Schneider, J.-L., Dezileau, L., et al. (2017). Storm-induced marine flooding: Lessons from
a multidisciplinary approach. Earth-Science Reviews, 165, 151–184. https://doi.org/10.1016/j.earscirev.2016.12.005
Chiu, S., & Small, C. (2016). Observations of cyclone-induced storm surge in coastal Bangladesh. Journal of Coastal Research, 32(5), 1149–
1161. https://doi.org/10.2112/JCOASTRES-D-15-00030.1
CIESIN (2018). Documentation for the Gridded Population of the World, Version 4 (GPWv4), revision 11 data sets. NASA Socioeconomic Data
and Applications Center (SEDAC), Center for International Earth Science Information Network, Columbia University. https://doi.org/10.7927/
H45Q4T5F
Clark, M. K., & Bilham, R. (2008). Miocene rise of the Shillong Plateau and the beginning of the end for the Eastern Himalaya. Earth and Plan-
etary Science Letters, 269(3), 337–351. https://doi.org/10.1016/j.epsl.2008.01.045
Connors, C. D., Hughes, A. N., & Ball, S. M. (2021). Forward kinematic modeling of fault-bend folding. Journal of Structural Geology, 143,
104252. https://doi.org/10.1016/j.jsg.2020.104252
Coudurier-Curveur, A., Tapponnier, P., Okal, E., Van der Woerd, J., Kali, E., Choudhury, S., et al. (2020). A composite rupture model for the
great 1950 Assam earthquake across the cusp of the East Himalayan Syntaxis. Earth and Planetary Science Letters, 531, 115928. https://doi.
org/10.1016/j.epsl.2019.115928
Cowie, P. A., & Scholz, C. H. (1992). Displacement-length scaling relationship for faults: Data synthesis and discussion. Journal of Structural
Geology, 14(10), 1149–1156. https://doi.org/10.1016/0191-8141(92)90066-6
Cummins, P. R. (2007). The potential for giant tsunamigenic earthquakes in the northern Bay of Bengal. Nature, 449(7158), 75–78. https://doi.
org/10.1038/nature06088
Cummins, P. R., & Kaneda, Y. (2000). Possible splay fault slip during the 1946 Nankai earthquake. Geophysical Research Letters, 27(17),
2725–2728. https://doi.org/10.1029/1999GL011139
Curray, J. R. (1991). Geological history of the Bengal geosyncline. Journal of Association of Exploration Geophysicists.
Earnest, A., Sunilkumar, T. C., & Silpa, K. (2021). Sinking slab stress and seismo-tectonics of the Indo-Burmese arc: A reappraisal. Tectonics,
40(8), e2021TC006827. https://doi.org/10.1029/2021TC006827

ORYAN ET AL. 18 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Engdahl, E. R., van der Hilst, R., & Buland, R. (1998). Global teleseismic earthquake relocation with improved travel times and procedures for
depth determination. Bulletin of the Seismological Society of America, 88(3), 722–743. https://doi.org/10.1785/BSSA0880030722
Fadil, W., Wei, S., Bradley, K., Wang, Y., He, Y., Sandvol, E., et al. (2023). Active faults revealed and new constraints on their seismogenic depth
from a high-resolution regional focal mechanism catalog in Myanmar (2016–2021). Bulletin of the Seismological Society of America, 113(2),
613–635. https://doi.org/10.1785/0120220195
Gahalaut, V. K., Kundu, B., Laishram, S. S., Catherine, J., Kumar, A., Devchandra Singh, M., et al. (2013). Aseismic plate boundary in the
Indo-Burmese wedge, northwest Sunda Arc. Geology, 41(2), 235–238. https://doi.org/10.1130/G33771.1
Gait, E. (1906). A history of Assam. Thacker, Spink & Company.
Gao, X., & Wang, K. (2014). Strength of stick-slip and creeping subduction megathrusts from heat flow observations. Science, 345(6200),
1038–1041. https://doi.org/10.1126/science.1255487
Govin, G., Najman, Y., Copley, A., Millar, I., van der Beek, P., Huyghe, P., et al. (2018). Timing and mechanism of the rise of the Shillong Plateau
in the Himalayan foreland. Geology, 46(3), 279–282. https://doi.org/10.1130/G39864.1
Grall, C., Steckler, M. S., Pickering, J. L., Goodbred, S., Sincavage, R., Paola, C., et al. (2018). A base-level stratigraphic approach to deter-
mining Holocene subsidence of the Ganges–Meghna–Brahmaputra Delta plain. Earth and Planetary Science Letters, 499, 23–36. https://doi.
org/10.1016/j.epsl.2018.07.008
Hanebuth, T. J. J., Kudrass, H. R., Zander, A. M., Akhter, H. S., Neumann-Denzau, G., & Zahid, A. (2021). Stepwise, earthquake-driven coastal
subsidence in the Ganges–Brahmaputra Delta (Sundarbans) since the eighth century deduced from submerged in situ kiln and mangrove
remnants. Natural Hazards, 111(1), 163–190. https://doi.org/10.1007/s11069-021-05048-2
Hass, B., & Harris, R. N. (2016). Heat flow along the Costa Rica Seismogenesis Project drilling transect: Implications for hydrothermal and
seismic processes. Geochemistry, Geophysics, Geosystems, 17(6), 2110–2127. https://doi.org/10.1002/2016GC006314
Herring, T. (2003). MATLAB Tools for viewing GPS velocities and time series. GPS Solutions, 7(3), 194–199. https://doi.org/10.1007/
s10291-003-0068-0
Herring, T. A., King, R. W., & McClusky, S. C. (2010a). GAMIT reference manual, release 10.4. Massachusetts Institute of Technology.
Herring, T. A., King, R. W., & McClusky, S. C. (2010b). GLOBK reference manual, global Kalman Filter VLBI and GPS analysis program,
release 10.4. Massachusetts Institute ofTechnology, Department of Earth, Atmospheric, and Planetary Sciences.
Hubbard, J., Shaw, J. H., & Klinger, Y. (2010). Structural setting of the 2008 Mw 7.9 Wenchuan, China, Earthquake. Bulletin of the Seismological
Society of America, 100(5B), 2713–2735. https://doi.org/10.1785/0120090341
Hurukawa, N., & Maung Maung, P. (2011). Two seismic gaps on the Sagaing Fault, Myanmar, derived from relocation of historical earthquakes
since 1918. Geophysical Research Letters, 38(1), L01310. https://doi.org/10.1029/2010GL046099
Ishii, M., Shearer, P. M., Houston, H., & Vidale, J. E. (2005). Extent, duration and speed of the 2004 Sumatra-Andaman earthquake imaged by
the Hi-Net array. Nature, 435(7044), 933–936. https://doi.org/10.1038/nature03675
Iyengar, R. N., Sharma, D., & Siddiqui, J. M. (1999). Earthquake history of India (Vol. 58). Medieval Times.
Jolivet, R., Simons, M., Duputel, Z., Olive, J., Bhat, H. S., & Bletery, Q. (2020). Interseismic loading of subduction megathrust drives long term
uplift in northern Chile, 1–21. https://doi.org/10.1029/2019GL085377
Kanda, R. V. S., & Simons, M. (2010). An elastic plate model for interseismic deformation in subduction zones. Journal of Geophysical Research,
115(B3), B03405. https://doi.org/10.1029/2009JB006611
Karmakar, S., Sirajul Haque, S. M., Mozaffar Hossain, M., & Shafiq, M. (2011). Water quality of Kaptai reservoir in Chittagong Hill Tracts of
Bangladesh. Journal of Forestry Research, 22(1), 87–92. https://doi.org/10.1007/s11676-011-0131-6
Kundu, B., & Gahalaut, V. K. (2012). Earthquake occurrence processes in the Indo-Burmese wedge and Sagaing fault region. Tectonophysics,
524–525, 135–146. https://doi.org/10.1016/j.tecto.2011.12.031
Lachenbruch, A. H., & Sass, J. H. (1980). Heat flow and energetics of the San Andreas Fault Zone. Journal of Geophysical Research: Solid Earth,
85(B11), 6185–6222. https://doi.org/10.1029/JB085iB11p06185
Lamb, S. (2021). The relation between short- and long-term deformation in actively deforming plate boundary zones. Philosophical Transactions
of the Royal Society A: Mathematical, Physical & Engineering Sciences, 379(2193), 20190414. https://doi.org/10.1098/rsta.2019.0414
Laske, G., Masters, G., Ma, Z., & Pasyanos, M. (2013). Update on CRUST1. 0—A 1-degree global model of Earth’s crust. Geophysical Research
abstracts, 15(15), 2658.
Lay, T., Kanamori, H., Ammon, C. J., Koper, K. D., Hutko, A. R., Ye, L., et al. (2012). Depth-varying rupture properties of subduction zone
megathrust faults. Journal of Geophysical Research: Solid Earth, 117(4), 1–21. https://doi.org/10.1029/2011JB009133
Lay, T., Kanamori, H., Ammon, C. J., Nettles, M., Ward, S. N., Aster, R. C., et al. (2005). The great Sumatra-Andaman earthquake of 26 Decem-
ber 2004. Science, 308(5725), 1127–1133. https://doi.org/10.1126/science.1112250
Le Dain, A. Y., Tapponnier, P., & Molnar, P. (1984). Active faulting and tectonics of Burma and surrounding regions. Journal of Geophysical
Research, 89(B1), 453. https://doi.org/10.1029/JB089iB01p00453
Lindsey, E. O., Mallick, R., Hubbard, J. A., Bradley, K. E., Almeida, R. V., Moore, J. D. P., et al. (2021). Slip rate deficit and earthquake potential
on shallow megathrusts. Nature Geoscience, 14(5), 321–326. https://doi.org/10.1038/s41561-021-00736-x
Liu, C.-Z., Chung, S.-L., Wu, F.-Y., Zhang, C., Xu, Y., Wang, J.-G., et al. (2016). Tethyan suturing in Southeast Asia: Zircon U-Pb and Hf-O
isotopic constraints from Myanmar ophiolites. Geology, 44(4), 311–314. https://doi.org/10.1130/G37342.1
Liu, C.-Z., Zhang, C., Xu, Y., Wang, J.-G., Chen, Y., Guo, S., et al. (2016). Petrology and geochemistry of mantle peridotites from the Kalaymyo
and Myitkyina ophiolites (Myanmar): Implications for tectonic settings. Lithos, 264, 495–508. https://doi.org/10.1016/j.lithos.2016.09.013
Madella, A., & Ehlers, T. A. (2021). Contribution of background seismicity to forearc uplift. Nature Geoscience, 14(8), 1–6. https://doi.
org/10.1038/s41561-021-00779-0
Mallick, B., Rahaman, K. R., & Vogt, J. (2011). Coastal livelihood and physical infrastructure in Bangladesh after cyclone Aila. Mitigation and
Adaptation Strategies for Global Change, 16(6), 629–648. https://doi.org/10.1007/s11027-011-9285-y
Mallick, R., Bürgmann, R., Johnson, K., & Hubbard, J. (2021). A unified framework for earthquake sequences and the growth of geological
structure in fold-thrust belts. Journal of Geophysical Research: Solid Earth, 126(9), e2021JB022045. https://doi.org/10.1029/2021JB022045
Mallick, R., Hubbard, J. A., Lindsey, E. O., Bradley, K. E., Moore, J. D. P., Ahsan, A., et al. (2020). Subduction initiation and the rise of the
Shillong Plateau. Earth and Planetary Science Letters, 543, 116351. https://doi.org/10.1016/j.epsl.2020.116351
Mallick, R., Lindsey, E. O., Feng, L., Hubbard, J., Banerjee, P., & Hill, E. M. (2019). Active convergence of the India-Burma-Sunda plates revealed
by a new continuous GPS network. Journal of Geophysical Research: Solid Earth, 124(3), 3155–3171. https://doi.org/10.1029/2018JB016480
Mallick, R., Parameswaran, R. M., & Rajendran, K. (2017). The 2005 and 2010 earthquakes on the Sumatra–Andaman trench: Evidence
for post-2004 megathrust intraplate rejuvenation. Bulletin of the Seismological Society of America, 107(4), 1569–1581. https://doi.
org/10.1785/0120160147

ORYAN ET AL. 19 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Maneerat, P., & Bürgmann, R. (2022). Geomorphic expressions of active tectonics across the Indo-Burma Range. Journal of Asian Earth
Sciences, 223, 105008. https://doi.org/10.1016/j.jseaes.2021.105008
Maneerat, P., Bürgmann, R., & Betka, P. M. (2022). Thrust sequence in the Western fold-and-thrust belt of the Indo-Burma Range determined
from fluvial profile analysis and dynamic landform modeling. Tectonophysics, 845, 229638. https://doi.org/10.1016/j.tecto.2022.229638
Maneerat, P., Dreger, D. S., & Bürgmann, R. (2022). Stress orientations and driving forces in the Indo-Burma plate boundary zone. Bulletin of
the Seismological Society of America, 112(3), 1323–1335. https://doi.org/10.1785/0120210303
Marrett, R. A., & Allmendinger, R. W. (1990). Kinematic analysis of fault-slip data. Journal of Structural Geology, 12(8), 973–986. https://doi.
org/10.1016/0191-8141(90)90093-e
Materna, K., Feng, L., Lindsey, E. O., Hill, E. M., Ahsan, A., Alam, A. K. M. K., et al. (2021). GNSS characterization of hydrological loading in
South and Southeast Asia. Geophysical Journal International, 224(3), 1742–1752. https://doi.org/10.1093/gji/ggaa500
Maurin, T., Masson, F., Rangin, C., Min, U. T., & Collard, P. (2010). First global positioning system results in northern Myanmar: Constant and
localized slip rate along the Sagaing fault. Geology, 38(7), 591–594. https://doi.org/10.1130/G30872.1
Maurin, T., & Rangin, C. (2009). Structure and kinematics of the Indo-Burmese Wedge: Recent and fast growth of the outer wedge. Tectonics,
28(2), TC2010. https://doi.org/10.1029/2008TC002276
Meade, B. J. (2010). The signature of an unbalanced earthquake cycle in Himalayan topography? Geology, 38(11), 987–990. https://doi.
org/10.1130/G31439.1
Meade, B. J., & Loveless, J. P. (2009). Block modeling with connected fault-network geometries and a linear elastic coupling estimator in spher-
ical coordinates. Bulletin of the Seismological Society of America, 99(6), 3124–3139. https://doi.org/10.1785/0120090088
Melnick, D., Moreno, M., Motagh, M., Cisternas, M., & Wesson, R. L. (2012). Splay fault slip during the Mw 8.8 2010 Maule Chile earthquake.
Geology, 40(3), 251–254. https://doi.org/10.1130/G32712.1
Milliman, J. D., & Farnsworth, K. L. (2011). River discharge to the coastal ocean: A global synthesis. Cambridge University Press.
Mitra, S., Bhattacharya, S. N., & Nath, S. K. (2008). Crustal structure of the Western Bengal basin from joint analysis of teleseismic receiver func-
tions and Rayleigh-wave dispersion. Bulletin of the Seismological Society of America, 98(6), 2715–2723. https://doi.org/10.1785/0120080141
Mitra, S., Priestley, K. F., Borah, K., & Gaur, V. K. (2018). Crustal structure and evolution of the eastern Himalayan plate boundary system,
northeast India. Journal of Geophysical Research: Solid Earth, 123(1), 621–640. https://doi.org/10.1002/2017JB014714
Mon, C. T., Gong, X., Wen, Y., Jiang, M., Chen, Q.-F., Zhang, M., et al. (2020). Insight into major active faults in Central Myanmar and the related
geodynamic sources. Geophysical Research Letters, 47(8), e2019GL086236. https://doi.org/10.1029/2019GL086236
Mondal, D. R., McHugh, C. M., Mortlock, R. A., Steckler, M. S., Mustaque, S., & Akhter, S. H. (2018). Microatolls document the 1762 and prior
earthquakes along the southeast coast of Bangladesh. Tectonophysics, 745, 196–213. https://doi.org/10.1016/j.tecto.2018.07.020
Morley, C. K., Tin Tin, N., Searle, M., & Robinson, S. A. (2020). Structural and tectonic development of the Indo-Burma ranges. Earth-Science
Reviews, 200, 102992. https://doi.org/10.1016/j.earscirev.2019.102992
Morris, B. L., Seddique, A. A., & Ahmed, K. M. (2003). Response of the Dupi Tila aquifer to intensive pumping in Dhaka, Bangladesh. Hydro-
geology Journal, 11(4), 496–503. https://doi.org/10.1007/s10040-003-0274-4
Mukhopadhyay, M., & Dasgupta, S. (1988). Deep structure and tectonics of the Burmese arc: Constraints from earthquake and gravity data.
Tectonophysics, 149(3–4), 299–322. https://doi.org/10.1016/0040-1951(88)90180-1
Murphy, R. W. (1988). Bangladesh enters the oil era. Bangladesh Enters the Oil Era, 86(9), 76–82.
Najman, Y., Bracciali, L., Parrish, R. R., Chisty, E., & Copley, A. (2016). Evolving strain partitioning in the eastern Himalaya: The growth of the
Shillong Plateau. Earth and Planetary Science Letters, 433, 1–9. https://doi.org/10.1016/j.epsl.2015.10.017
Najman, Y., Sobel, E. R., Millar, I., Luan, X., Garzanti, E., Parra, M., et al. (2022). The timing of collision between Asia and the West Burma
Terrane, and the development of the Indo-Burman Ranges. Tectonics, 41(7), e2021TC007057. https://doi.org/10.1029/2021TC007057
Najman, Y., Sobel, E. R., Millar, I., Stockli, D. F., Govin, G., Lisker, F., et al. (2020). The exhumation of the Indo-Burman ranges, Myanmar.
Earth and Planetary Science Letters, 530, 115948. https://doi.org/10.1016/j.epsl.2019.115948
Nelson, A. R., Kelsey, H. M., & Witter, R. C. (2006). Great earthquakes of variable magnitude at the Cascadia subduction zone. Quaternary
Research, 65(3), 354–365. https://doi.org/10.1016/j.yqres.2006.02.009
Ni, J. F., Guzman-Speziale, M., Bevis, M., Holt, W. E., Wallace, T. C., & Seager, W. R. (1989). Accretionary tectonics of Burma and the three-dimensional
geometry of the Burma subduction zone. Geology, 17(1), 68. https://doi.org/10.1130/0091-7613(1989)017<0068:ATOBAT>2.3.CO;2
Nielsen, C., Chamot-Rooke, N., & Rangin, C. (2004). From partial to full strain partitioning along the Indo-Burmese hyper-oblique subduction.
Marine Geology, 209(1–4), 303–327. https://doi.org/10.1016/j.margeo.2004.05.001
Nuruddin, M., & Hai, M. A. (2003). Kaptai Lake. Banglapedia - Natioal Encycylopedia of Bangladesh.
Oryan, B., & Buck, W. R. (2020). Larger tsunamis from megathrust earthquakes where slab dip is reduced. Nature Geoscience, 13(4), 319–324.
https://doi.org/10.1038/s41561-020-0553-x
Oryan, B., Olive, J.-A., Jolivet, R., Bruhat, L., & Malatesta, L. (2021). Non-recoverable deformation during the interseismic phase of the subduc-
tion earthquake cycle. Paper presented at the AGU Fall Meeting Abstracts, 2021, T45F-07.
Oryan, B., Olive, J.-A., Jolivet, R., Bruhat, L., & Malatesta, L. (2022). Long-term coastal uplift due to non-recoverable forearc deformation during
the interseismic phase of the subduction earthquake cycle (No. EGU22-1557). Paper presented at the EGU22, Copernicus Meetings. https://
doi.org/10.5194/egusphere-egu22-1557
Oryan, B., & Savage, H. (2021). Regional heat flow analysis reveals frictionally weak Dead Sea fault. Geochemistry, Geophysics, Geosystems,
22(12), e2021GC010115. https://doi.org/10.1029/2021GC010115
Oryan, B., Steckler, M. S., Mondal, D. R., Akhter, S. H., Singha, S., Grall, C., & Lindsey, E. O. (2020). The Indo-Burma detachment geometry
constrained by an updated vertical and horizontal GPS velocity field in Bangladesh. Paper presented at the AGU Fall Meeting Abstracts, 2020,
T048-0021.
Panda, D., & Kundu, B. (2022). Geodynamic complexity of the Indo-Burmese arc region and its interaction with Northeast Himalaya.
Earth-Science Reviews, 226, 103959. https://doi.org/10.1016/j.earscirev.2022.103959
Panda, D., Kundu, B., Gahalaut, V. K., & Rangin, C. (2020). India-Sunda plate motion, crustal deformation, and seismic hazard in the
Indo-Burmese arc. Tectonics, 39(8), e2019TC006034. https://doi.org/10.1029/2019TC006034
Panda, D., Samanta, S. K., Singh, M. D., Gahalaut, V. K., & Kundu, B. (2022). Low-effective fault strength of a blind detachment beneath
the Indo-Burmese Arc (NE-India) induced by frictional–viscous flow. Journal of Earth System Science, 131(1), 9. https://doi.org/10.1007/
s12040-021-01754-4
Petit, J. P. (1987). Criteria for the sense of movement on fault surfaces in brittle rocks. Journal of Structural Geology, 9(5–6), 597–608. https://
doi.org/10.1016/0191-8141(87)90145-3
Pivnik, D. A., Nahm, J., & Tuck, R. S. (1998). Polyphase deformation in a fore-arc/back-arc basin, Salin Subbasin, Myanmar (Burma). AAPG
Bulletin, 82(10), 1837–1856. https://doi.org/10.1306/1D9BD15F-172D-11D7-8645000102C1865D

ORYAN ET AL. 20 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Purnachandra Rao, N., & Kumar, M. R. (1999). Evidences for cessation of Indian plate subduction in the Burmese arc region. Geophysical
Research Letters, 26(20), 3149–3152. https://doi.org/10.1029/1999GL005396
Rangin, C., Maurin, T., & Masson, F. (2013). Combined effects of Eurasia/Sunda oblique convergence and East-Tibetan crustal flow on the active
tectonics of Burma. Journal of Asian Earth Sciences, 76, 185–194. https://doi.org/10.1016/j.jseaes.2013.05.018
Rangin, C., & Sibuet, J.-C. (2017). Structure of the northern Bay of Bengal offshore Bangladesh: Evidences from new multi-channel seismic data.
Marine and Petroleum Geology, 84, 64–75. https://doi.org/10.1016/j.marpetgeo.2017.03.020
Sapkota, S. N., Bollinger, L., Klinger, Y., Tapponnier, P., Gaudemer, Y., & Tiwari, D. (2013). Primary surface ruptures of the great Himalayan
earthquakes in 1934 and 1255. Nature Geoscience, 6(1), 71–76. https://doi.org/10.1038/ngeo1669
Satake, K., Shimazaki, K., Tsuji, Y., & Ueda, K. (1996). Time and size of a giant earthquake in Cascadia inferred from Japanese tsunami records
of January 1700. Nature, 379(6562), 246–249. https://doi.org/10.1038/379246a0
Satyabala, S. P. (1998). Subduction in the Indo-Burma region: Is it still active? Geophysical Research Letters, 25(16), 3189–3192. https://doi.
org/10.1029/98GL02256
Schellart, W. P., Stegman, D. R., Farrington, R. J., & Moresi, L. (2011). Influence of lateral slab edge distance on plate velocity, trench velocity,
and subduction partitioning. Journal of Geophysical Research: Solid Earth, 116(B10), B10408. https://doi.org/10.1029/2011JB008535
Schultz, R. A., & Fossen, H. (2002). Displacement–length scaling in three dimensions: The importance of aspect ratio and application to defor-
mation bands. Journal of Structural Geology, 24(9), 1389–1411. https://doi.org/10.1016/S0191-8141(01)00146-8
Searle, M. P., Morley, C. K., Waters, D. J., Gardiner, N. J., Htun, U. K., Nu, T. T., & Robb, L. J. (2017). Chapter 12 – Tectonic and metamor-
phic evolution of the Mogok Metamorphic and Jade Mines belts and ophiolitic terranes of Burma (Myanmar). Geological Society, London,
Memoirs, 48(1), 261–293. https://doi.org/10.1144/M48.12
Seno, T., & Yoshida, M. (2004). Where and why do large shallow intraslab earthquakes occur? Physics of the Earth and Planetary Interiors,
141(3), 183–206. https://doi.org/10.1016/j.pepi.2003.11.002
Shum, C. K., Kuo, C. Y., Guo, J., Shang, K., Tseng, K. H., Wan, J., et al. (2014). Quantifying and projecting relative seasea-level rise at the
regional scale: The Bangladesh Sea-Level Project (BanD-AID). Paper presented at the AGU Fall Meeting Abstracts, 2014, G21B-0439.
Sibuet, J.-C., Klingelhoefer, F., Huang, Y.-P., Yeh, Y.-C., Rangin, C., Lee, C.-S., & Hsu, S.-K. (2016). Thinned continental crust intruded by
volcanics beneath the northern Bay of Bengal. Marine and Petroleum Geology, 77, 471–486. https://doi.org/10.1016/j.marpetgeo.2016.07.006
Sibuet, J.-C., Klingelhoefer, F., Yeh, Y.-C., Rangin, C., & Lee, C.-S. (2017). Reply to the comment of Talwani et al. (2017) on the Sibuet et al.
(2016) paper entitled “Thinned continental crust intruded by volcanics beneath the northern Bay of Bengal. Marine and Petroleum Geology,
88, 1126–1129. https://doi.org/10.1016/j.marpetgeo.2017.07.023
Sikder, A. M., & Alam, M. M. (2003). 2-D modelling of the anticlinal structures and structural development of the eastern fold belt of the Bengal
Basin, Bangladesh. Sedimentary Geology, 155(3–4), 209–226. https://doi.org/10.1016/S0037-0738(02)00181-1
Simpson, D. W., Leith, W. S., & Scholz, C. H. (1988). Two types of reservoir-induced seismicity. Bulletin of the Seismological Society of America,
78(6), 2025–2040. https://doi.org/10.1785/BSSA0780062025
Sincavage, R., Betka, P. M., Thomson, S. N., Seeber, L., Steckler, M., & Zoramthara, C. (2020). Neogene shallow-marine and fluvial sediment
dispersal, burial, and exhumation in the ancestral Brahmaputra delta: Indo-Burman Ranges, India. Journal of Sedimentary Research, 90(9),
1244–1263. https://doi.org/10.2110/jsr.2020.60
Singh, A., Bhushan, K., Singh, C., Steckler, M. S., Akhter, S. H., Seeber, L., et al. (2016). Crustal structure and tectonics of Bangladesh: New
constraints from inversion of receiver functions. Tectonophysics, 680, 99–112. https://doi.org/10.1016/j.tecto.2016.04.046
Sloan, R. A., Elliott, J. R., Searle, M. P., & Morley, C. K. (2017). Chapter 2 – Active tectonics of Myanmar and the Andaman Sea. Geological
Society, London, Memoirs, 48(1), 19–52. https://doi.org/10.1144/M48.2
Socquet, A., Vigny, C., Chamot-Rooke, N., Simons, W., Rangin, C., & Ambrosius, B. (2006). India and Sunda plates motion and defor-
mation along their boundary in Myanmar determined by GPS. Journal of Geophysical Research: Solid Earth, 111(B5). https://doi.
org/10.1029/2005JB003877
Steckler, M. S., Akhter, S. H., & Seeber, L. (2008). Collision of the Ganges-Brahmaputra Delta with the Burma Arc: Implications for earthquake
hazard. Earth and Planetary Science Letters, 273(3–4), 367–378. https://doi.org/10.1016/j.epsl.2008.07.009
Steckler, M. S., Mondal, D. R., Akhter, S. H., Seeber, L., Feng, L., Gale, J., et al. (2016). Locked and loading megathrust linked to active subduc-
tion beneath the Indo-Burman Ranges. Nature Geoscience, 9(8), 615–618. https://doi.org/10.1038/ngeo2760
Steckler, M. S., Nooner, S. L., Akhter, S. H., Chowdhury, S. K., Bettadpur, S., Seeber, L., & Kogan, M. G. (2010). Modeling Earth deformation
from monsoonal flooding in Bangladesh using hydrographic, GPS, and Gravity Recovery and Climate Experiment (GRACE) data. Journal of
Geophysical Research: Solid Earth, 115, B08407. https://doi.org/10.1029/2009jb007018
Steckler, M. S., Oryan, B., Betka, P. M., Seeber, L., & Than, O. (2021). Strain partitioning in the IndoBurma Subduction Zone: GNSS constraints
on the Kabaw Fault in Myanmar. Paper presented at the AGU Fall Meeting 2021. AGU. Retrieved from https://agu.confex.com/agu/fm21/
meetingapp.cgi/Paper/962856
Steckler, M. S., Oryan, B., Wilson, C. A., Grall, C., Nooner, S. L., Mondal, D. R., et al. (2022). Synthesis of the distribution of subsidence of the
lower Ganges-Brahmaputra Delta, Bangladesh. Earth-Science Reviews, 224, 103887. https://doi.org/10.1016/j.earscirev.2021.103887
Stork, A. L., Selby, N. D., Heyburn, R., & Searle, M. P. (2008). Accurate relative earthquake hypocenters reveal structure of the Burma subduction
zone. Bulletin of the Seismological Society of America, 98(6), 2815–2827. https://doi.org/10.1785/0120080088
Subedi, S., & Hetényi, G. (2021). Precise locating of the great 1897 Shillong Plateau earthquake using teleseismic and regional seismic phase
data. The Seismic Record, 1(3), 135–144. https://doi.org/10.1785/0320210031
Suppe, J. (1983). Geometry and kinematics of fault-bend folding. American Journal of Science, 283, 684–721. https://doi.org/10.2475/
ajs.283.7.684
Talwani, M., Desa, M. A., Ismaiel, M., & Sree Krishna, K. (2016). The Tectonic origin of the Bay of Bengal and Bangladesh. Journal of Geophys-
ical Research: Solid Earth, 121(7), 4836–4851. https://doi.org/10.1002/2015JB012734
Talwani, M., Krishna, K. S., Ismaiel, M., & Desa, M. A. (2017). Comment on a paper by Sibuet et al. (2016) entitled “Thinned continental
crust intruded by volcanics beneath the northern Bay of Bengal”. Marine and Petroleum Geology, 88, 1123–1125. https://doi.org/10.1016/j.
marpetgeo.2016.12.009
Tun, S. T., Thein, M., Htay, N., & Sein, K. (2014). Geological map of Myanmar (1: 2,250,000). Myanmar Geosciences Society.
Vigny, C. (2003). Present-day crustal deformation around Sagaing fault, Myanmar. Journal of Geophysical Research, 108(B11), 2533. https://
doi.org/10.1029/2002jb001999
Vorobieva, I., Gorshkov, A., & Mandal, P. (2021). Modelling the seismic potential of the Indo-Burman megathrust. Scientific Reports, 11(1),
21200. https://doi.org/10.1038/s41598-021-00586-y

ORYAN ET AL. 21 of 22
Journal of Geophysical Research: Solid Earth 10.1029/2022JB025550

Wang, Y., Shyu, J. B. H., Sieh, K., Chiang, H.-W., Wang, C.-C., Aung, T., et al. (2013). Permanent upper plate deformation in western Myanmar
during the great 1762 earthquake: Implications for neotectonic behavior of the northern Sunda megathrust. Journal of Geophysical Research:
Solid Earth, 118(3), 1277–1303. https://doi.org/10.1002/jgrb.50121
Wang, Y., Sieh, K., Aung, T., Min, S., Khaing, S. N., & Tun, S. T. (2011). Earthquakes and slip rate of the southern Sagaing fault:
Insights from an offset ancient fort wall, lower Burma (Myanmar). Geophysical Journal International, 185(1), 49–64. https://doi.
org/10.1111/j.1365-246X.2010.04918.x
Wang, Y., Sieh, K., Tun, S. T., Lai, K.-Y., & Myint, T. (2014). Active tectonics and earthquake potential of the Myanmar region. Journal of
Geophysical Research: Solid Earth, 119(4), 3767–3822. https://doi.org/10.1002/2013JB010762
Wessel, P., Luis, J. F., Uieda, L., Scharroo, R., Wobbe, F., Smith, W. H. F., & Tian, D. (2019). The generic mapping tools version 6. Geochemistry,
Geophysics, Geosystems, 20(11), 5556–5564. https://doi.org/10.1029/2019GC008515
Westerweel, J., Licht, A., Cogné, N., Roperch, P., Dupont-Nivet, G., Kay Thi, M., et al. (2020). Burma Terrane collision and northward indenta-
tion in the Eastern Himalayas recorded in the Eocene-Miocene Chindwin Basin (Myanmar). Tectonics, 39(10), e2020TC006413. https://doi.
org/10.1029/2020tc006413
Wu, S., Yao, J., Wei, S., Hubbard, J., Wang, Y., Htwe, Y. M. M., et al. (2021). New insights into the structural heterogeneity and geodynamics of
the Indo-Burma subduction zone from ambient noise tomography. Earth and Planetary Science Letters, 562, 116856. https://doi.org/10.1016/j.
epsl.2021.116856
Yao, J., Liu, S., Wei, S., Hubbard, J., Huang, B.-S., Chen, M., & Tong, P. (2021). Slab models beneath central Myanmar revealed by a joint
inversion of regional and teleseismic traveltime data. Journal of Geophysical Research: Solid Earth, 126(2), e2020JB020164. https://doi.
org/10.1029/2020JB020164
Yue, L.-F., Suppe, J., & Hung, J.-H. (2005). Structural geology of a classic thrust belt earthquake: The 1999 Chi-Chi earthquake Taiwan
(Mw=7.6). Journal of Structural Geology, 27(11), 2058–2083. https://doi.org/10.1016/j.jsg.2005.05.020
Zehnder, A. T., & Allmendinger, R. W. (2000). Velocity field for the trishear model. Journal of Structural Geology, 22(8), 1009–1014. https://
doi.org/10.1016/S0191-8141(00)00037-7
Zhang, G., He, Y., Ai, Y., Jiang, M., Mon, C. T., Hou, G., et al. (2021). Indian continental lithosphere and related volcanism beneath Myanmar:
Constraints from local earthquake tomography. Earth and Planetary Science Letters, 567, 116987. https://doi.org/10.1016/j.epsl.2021.116987
Zheng, T., He, Y., Ding, L., Jiang, M., Ai, Y., Mon, C. T., et al. (2020). Direct structural evidence of Indian continental subduction beneath
Myanmar. Nature Communications, 11(1), 1944. https://doi.org/10.1038/s41467-020-15746-3

Erratum
The originally published version of this article contained some typographical errors. In the tenth sentence of the
fourth paragraph of Section 3.2, “6)” should be “f).” The sentence now reads: “f) the structure of the Burma crust
as constrained by seismic wave velocity profiles (Zhang et al., 2021).” Figures 6 and S10 included typograph-
ical errors involving missing circles and lines. In addition, the Author Contributions list contained typographi-
cal errors involving contributions directed to only a few authors. Conceptualization should be attributed to Bar
Oryan, Paul M. Betka, and Syed Humayun Akhter. Data curation should be attributed to Paul M. Betka, Michael
S. Steckler, Bar Oryan, Eric O. Lindsey, Syed Humayun Akhter, and Dhiman Mondal. Funding Acquisition
should be attributed to Syed Humayun Akhter and Paul M. Betka. Investigation should be attributed to Bar Oryan,
Paul M. Betka, Scott L. Nooner, Eric O. Lindsey, Dhiman Mondal, and Michael S. Steckler. Methodology should
be attributed to Scott L. Nooner, Eric O. Lindsey, Michael S. Steckler, and Dhiman Mondal. Supervision should
be attributed to Paul M. Betka. Validation should be attributed to Bar Oryan. The errors have been corrected, and
this may be considered the authoritative version of record.

ORYAN ET AL. 22 of 22

You might also like