You are on page 1of 23

GeoScienceWorld

Lithosphere
Volume 2022, Article ID 6058346, 23 pages
https://doi.org/10.2113/2022/6058346

Research Article
Understanding the Deformation Structures and Tectonics of the
Active Orogenic Fold-Thrust Belt: Insights from the Outer
Indo-Burman Ranges

Md. Sakawat Hossain ,1,2,3 Songjian Ao ,2,3,4 Tridib Kumar Mondal ,5 Arnab Sain ,6
Md. Sharif Hossain Khan ,1 Wenjiao Xiao ,3,4,7 and Pengpeng Zhang 2,3
1
Department of Geological Sciences, Jahangirnagar University, Dhaka 1342, Bangladesh
2
State Key Laboratory of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of Sciences,
Beijing 100029, China
3
Innovation Academy for Earth Science, CAS, China
4
China-Pakistan Joint Research Center on Earth Sciences, Chinese Academy of Sciences, Chengdu, China
5
Geological Studies Unit, Indian Statistical Institute, Kolkata 700108, India
6
Department of Geology, Presidency University, Kolkata 700073, India
7
Xinjiang Research Center for Mineral Resources, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences,
Urumqi 830011, China

Correspondence should be addressed to Songjian Ao; asj@mail.iggcas.ac.cn

Received 20 November 2021; Accepted 18 March 2022; Published 18 April 2022

Academic Editor: Feng Cheng

Copyright © 2022 Md. Sakawat Hossain et al. Exclusive Licensee GeoScienceWorld. Distributed under a Creative Commons
Attribution License (CC BY 4.0).

The tectonic deformation of the outer Indo-Burman Ranges (i.e., Chittagong Tripura Fold Belt, CTFB) is associated with the
oblique convergence of Indo-Burmese plates since the latest Miocene. This article presents detailed field evidence of
deformation structures and their kinematics in the exposed Tertiary successions in the CTFB. We combine observations made
in this study with the published structural, geodetic, and seismic data sets to present an overview of the active tectonic
framework of the region and its strain partitioning. To determine the kinematic evolution, décollement depth, and amount of
strain, we combined geologic field mapping, structural analysis of fifteen anticlines, fracture/lineament analysis, and paleostress
analysis of faults that define the ∼100 km wide CTFB. Structural data and kinematic analyses suggest subhorizontal plane strain
with approximately 10% east-west shortening (oriented ~65°) that is perpendicular to the axial plane (oriented ~155°) of the
CTFB anticlines. No evidence of significant transpression or strike-slip faulting has been observed in the CTFB and, therefore,
suggests that full slip-partitioning is normal to the outer belt and parallel to the inner belt of the IBR. Paleostress analysis
results are in good agreement with the present-day stress regime, and this implies that past and present deformation is
dynamically related with the normal component of India-Burma oblique vector velocity motion.

1. Introduction earthquake-prone area, where deformation and kinematics


are governed by oblique subduction of the Indian Plate
Understanding the structural evolution of the outer Indo- beneath the Burma plate [1–6]. This oblique convergence-
Burman Rages (IBR) is crucial to reconstructing the Ceno- related transpressional tectonics in orogenic belts is a com-
zoic geodynamics of the Bengal Basin and the complex mon phenomenon giving rise to complex deformation pat-
development of collisional events between the Indian and terns (e.g., [7–11]). Transpressional deformation involves
Burmese plates [1]. The plate boundary region between the both strike-slip and dip-slip components. However, the
India and Burma plates in the outer IBR region consists of strike-slip deformations can deviate from simple shear when
fold belts, and regional scale thrust (Figure 1) is a prominent shortening oblique to the deformation zone [10]. At the

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
2 Lithosphere

87°E 90°E 93°E

Himalayan Mountain Ranges


27°N

27°N
Main Frontal Thrust
.
Suture
gaF

Dhubri F.

Ko
Na
Thrust

pil
Sinistral
Dextral Mikir Hills

iF
Old

.
F. ham
Ti

F.
sa
ta
F.

.
Jamuna F

24°N
24°N

Bay of Bengal
21°N

21°N
N

Onshore digital elevation from SRTM (ver. 3)


Projection: WGS 1984 World Mercator SRTM Elevation
Model
50 0 100 200 8500
km
Reference Scale: 1:12,50,000
1250

255
Thrust Fault Deformation Front
Thrust/Strike-slip Undifferrentiated Fault 50
Thrust Front CTFB = Chittagong-Tripura Fold Belt
Strike-slip Fault IBR = Indo-Burman Ranges
18°N

18°N

0m
87°E 90°E 93°E

Figure 1: Simplified schematic tectonic map of the Bengal Basin and its surroundings. The background image is the digital elevation model
(DEM) of the area based on Shuttle Radar Topography Mission (SRTM) images. Compilations of tectonic features are based on Hossain
et al. [44]. The black rectangular box is the study area (CTFB). The major tectonic domains are shown in the index map (modified
after [103]).

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 3

obliquely convergent plate boundaries, deformation occurs ropy of the fracture system (i.e., fracture spacing and orien-
both locally and regionally and produces a range of defor- tation in a given direction) has been analyzed to understand
mation structures [9, 12]: folds, thrusts, and shear zones, the spatial distribution of the fractures and associated stress
which are essential for oil and gas exploration [13, 14] and field.
active tectonic studies [15, 16]. Understanding the evolution of geological structures and
Tectonic deformation of the western outer edge of the their dissemination with progressive deformation is the fun-
young and active IBR, known as the Chittagong Tripura damental aspect in inaugurating the tectonics of any region
Fold Belt (CTFB), is mainly attributed to the oblique [35, 36]. This leads to a better appraisal of the state of paleos-
India-Burmese convergence. This oblique convergence- tress conditions that led to the deformation on account of
related transpressional tectonics splits along different regional tectonics. Therefore, it becomes essential to study
tectono-geomorphic units separated by large-scale thrust the paleostress condition in accordance with the kinematics
and dextral strike-slip faults and produces a fold-thrust belt, of the structures. The previous studies on CTFB have only
in which the western limit is marked by the “Deformation been restricted to geochemistry and provenance studies
Front” (Figure 1) [2, 17]. The convergence is accommodated [37–41]; seismic hazards [2, 4–6, 19, 42–44]; and geody-
by different tectonic structures where the deformation inten- namics, tectonics, and geophysical investigations [1–3,
sity and age gradually decrease from IBR in the east to the 18–20, 42, 45, 46]. Except Betka et al.’s [17] slip-
CTFB in the west. According to Maurin and Rangin [18], a partitioning study that carried out locally in Tripura, no
total of about 11 km east-west shortening occurs in the last attempt has been made to comprehensively decipher the
2 Ma, indicating ~5 mm/year long-term shortening rate. deformation structures, related kinematics and stress regime,
Two regional scale active faults and thick-skinned tectonics and balanced geological cross-sections of the CTFB’s
of the CTFB constantly accommodate the stress associated Tertiary sediments and amount of strain in terms of field
with the Indo-Burma oblique subduction and, therefore, investigation. The present paper aims to fill this lacuna by
are responsible for the kinematic evolution of the CTFB presenting a complete structural map of the CTFB area of
[18, 19]. The two regional faults are the Chittagong Coastal the outer IBR (Figure 2) for the first time based on several
Fault (CCF) to the west and Kaladan Fault to the east, and comprehensive geological fieldwork carried out since 2008
these faults are running approximately N-S (Figure 1). The and other published maps and data sets ([2, 20, 47] and ref-
CCF is approximately a coast parallel structure where an erences there in). It also provides evidence for structural
overthrusting hanging wall concealed major parts of the deformations and interprets their kinematics in the exposed
western flank of the Sitakund, Inani, and Dakhin Nhila anti- Tertiary sedimentary rocks in the area. Structural interpreta-
clines (Figure 2). Recent geodetic measurements suggest the tions are then constrained by in-depth analysis of the folds,
tectonic convergence with overall E-W shortening in the faults, and fractures/lineaments geometry to understand
CTFB area occurred at the rate of ~5 mm/year [3, 5, 19]. the prevailing stress regime in the area.
This east-west shortening produces approximately a north- Three balanced geological cross-sections from the north-
south complex fold and overlapping thrust systems in the ern, central, and southern parts of the study area illustrate
CTFB area. The plunging fold system consists of a series of the geometry and kinematics of the structures developed in
NNW-SSE trending anticlines and synclines, where the the exposed Tertiary sediments and the total amount of
geometry is predominantly expressed by NNW-SSE striking strain. The fault-slip analysis reveals the paleostress direc-
reverse faulting [20]. Moreover, the presence of active tion that led to the formation of deformation structures in
macro-gas seepages in the CTFB area is reported and the CTFB. The tectonic evolution of the CTFB in response
encountered during the fieldwork. The presence of such to the oblique convergence along the Kaladan Fault and
seepages suggests hydrocarbon accumulation in the subsur- CCF is also discussed. This study helps characterize the fault,
face [21]. Therefore, the proper understanding of the defor- lineament/fracture disposition patterns in CTFB and pro-
mation structures (e.g., fold and fault) of the study area will vides a clearer understanding into the deformation kinemat-
facilitate having a clear insight regarding the hydrocarbon ics and tectonics of a young and active orogenic belt.
system of the area.
The folds, faults, joints, fractures, and their geometry can 2. Geological Settings
be used to interpret the regional deformation and kinematic
history. Fractal dimensional analyses of these structures over Extending from the northern end of the Arakan Trench in
2D space can reveal their formation processes and kinematic the south to the juncture of the Dauki, Halflong, Naga, and
evolution [22–25]. For example, the lineament/fracture net- Disag thrusts to the north, the CTFB has been recognized
work revealed from the Advanced Land Observing Satellite as the outer edge of the western Neogene belt of the Indo-
(ALOS) images is the imprint of two-dimensional features Burman Ranges (IBR), which separated from the eastern
in the ground surface [26, 27]. Rocks exhibit anisotropy Palaeogene belt of the IBR along the Kaladan Fault
when subjected to stress, and the resultant cracks/fractures [17–19]. Geologically, the CTFB is bounded by the Foredeep
are assumed to have preferential orientation [28–31]. The basin part of the Bengal Basin to the west and the IBR to the
study of such fracture networks, their size, and their spatial east and considered to be an active orogenic belt in the world
relationships are important to establish the regional stress (Figure 1). To the east, it is structurally separated from the
field, rock strength and its deformability [32], and kinematic IBR by the approximately N-S running Kaladan thrust (also
model [33, 34]. In this study, the fractal dimension anisot- known as Tut Fault). To the west, the boundary of the CTFB

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
4 Lithosphere

92°E 93°E

Barja
Sald
Aizawl

la
ana
di

Mualuaw
ti F.
Gum

Kho

m
Tichna

wai
C′
C

Tula
Gaja

mur
Fig.7d
lia

Da
kh
in

23°N
Nh
ila
N

Kaladan
Fig. 7d

F.
Chittagong
Sangu

22°N
Magnama

Lower Middle Bhuban and


Pre-Miocene Sediments
Mahesk

Upper Bhuban (Dominantly


sandstone with shale)
Mowdak
hali

Boka Bil (Shale and


sandstone alternation)

Tipam Sandstone

Dupi Tila Sandstone


Inan

45 30
Pin

20
i

10
cha

Quaternary Holocene
Deposits (Alluvium) N
F.

Big City
Thrust W E
Undifferentiated Fault (F.)
Dak

21°N

C′
Doubly plunging anticline C
hin

S
Pin

Boundary between
Nhila

Fig. 7f
cha

WFTZ and EHCFTZ


0 20 40 60 80
km

Figure 2: Geological map of the CTFB area of the Bengal Basin (structural data sets were complemented from [2, 20]; surface geology was
complemented from [47]). Abbreviations: CCF: Chittagong Coastal Fault; F: fault. AA ′ , BB ′ , and CC ′ lines are three balanced geological
cross-sections. The boundary between the eastern highly compressed fold-thrust zone (EHCFTZ) and the western fold-thrust zone
(WFTZ) is marked with the thick red dotted line on the map. The black rectangular box area shows the enlarge view of the CC ′ cross-
sectional area.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 5

Table 1: Major geotectonic events related to the convergence of the India-Burmese plates and subsequent development of the CTFB in the
eastern part of the Bengal Basin (after [2, 20]).

Geologic time Age in Ma Major geodynamic events Major tectonic events


The fold belt has been progressively
growing westward and actively
Late Pleistocene –recent ~0.125 developing the “Emerging Fold Belt”
between the Thrust Front and
Continued subduction – westward Deformation Front
migration of the Deformation Front
In the upper part (~ up to 5 km),
sediments deform above a decollement
Early Pleistocene ~2
level through thin-skinned and thick-
skinned tectonic processes
Compression Oblique subduction of the Indian Plate
beneath the Burmese Plate in an arc-
trench setting started to develop
Late Pliocene ~3.5-3
accretionary prism (i.e., CTFB) within the
Subduction (ocean-continent) – upper parts of the thick deltaic sequence
development of the IBR at the beginning in the eastern margin of the Bengal Basin
and later CTFB as the accretionary prism
Sediment contributions to the Bengal
Basin began to arrive from the rising IBR
Early Miocene ~20
due to the subduction of the Indian Plate
beneath the Burmese Plate
The northeastern corner of the Indian
plate making a glancing contact with the
Early Eocene ~45 Soft lateral convergence
Sumatra Block, followed by the Burmese
Plate

with the Foredeep basin is assumed to be along the edge of set, the CTFB area has been divided into two tectonic zones
the convex shape “Thrust Front” (Figure 1) [17, 20], repre- (Figure 2): (a) western fold-thrust zone and (b) eastern
sented by the CCF. The CCF is a regional thrust fault, run- highly compressed fold-thrust zone [20]. To the west,
ning NNW-SSE along the northeastern coastline of the Bay subduction-related deformation continues from the CTFB
of Bengal (Figures 1 and 2). According to Maurin and Ran- towards the current Foredeep basin, but the intensity of
gin [18] both the Kaladan and CCF have a dextral-slip com- the deformation gradually decreases from the east to the
ponent. In between the CCF and the Kaladan Fault, this west and ceases along the Deformation Front. The transi-
~100 km wide deformation/transpressional zone (i.e., the tional boundary of these two tectonic units is marked by
CTFB) is framed by en echelon plunging folds and the “Thrust Front” [17]. The continued deformation from
faults [20]. the Deformation Front to the Thrust Front produces an
Before the convergence of the India-Burma plates, the approximately 80 km wide young subsurface fold belt, which
CTFB area was an integral part of the Foredeep basin of is known as the Emerging Fold Belt [20].
the Bengal Basin [48]. The commencement of the CTFB The CTFB area consists of late Miocene–Quaternary
development in the eastern part of the Foredeep basin is accreted sediments of ~5 km thickness (Figure 2), where
mainly presumed to be started from the Pliocene as an the sedimentary facies ranges from shallow marine through
accretionary wedge due to the incessant subduction of the deltaic to fluvial environments [40, 48]. Within the CTFB,
Indian Plate’s oceanic crust beneath the Burmese Plate [4, the Miocene Surma Group consists of shallow marine shelfal
18, 39, 41]. With continued collision, the eastern part of and intertidal deposits and is divided into two formations:
the Foredeep basin evolved as a fold belt and form folded Bhuban at the bottom and Boka Bil at the top (Table 2).
flank of the Bengal Basin, currently known as the CTFB Overlying Boka Bil Formation, the Tipam Group comprised
(Table 1). Since last 2 Ma, the rapid development of the of Pliocene fluvial deposits and is divided into two forma-
CTFB and its westward propagation is facilitated from tions: Tipam Sandstone at the bottom and Girujan Clay at
upliftment of the Shillong Plateau and consequent deepen- the top. Above the Tipam Group, the Pliocene–Pleistocene
ing and sedimentary filling of the Sylhet Trough (Figure 1) Dupi Tila Group consists of meandering channels, flood-
[18]. Due to this rapid westward propagation since Pliocene, plain, and alluvial deposits [2, 40, 48].
thin-skinned tectonics of the CTFB is gradually engraved In the CTFB (i.e., the western prolongation of the Indo-
westward by thick-skinned tectonics of the IBR [18, 49]. Burman accretionary prism), the upper deltaic sediments of
Therefore, the deformation age of the structures gradually the Surma, Tipam, and Dupi Tila groups occur (~up to
increases eastward from the CTFB to the IBR. 5 km) above a major basal detachment fault and deformed
Based on the geometry and deformation intensity of the [42]. In the last 2 Ma, these deformations rise to a series of
folded structures, field observation, Shuttle Radar Topogra- elongate, ~NNW–SSE trending curvilinear plunging folds,
phy Mission (SRTM) image analysis, and published data indicating the combination of thick-skinned and thin-

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
6 Lithosphere

Table 2: Stratigraphic succession of the CTFB in the eastern part of the Bengal Basin (after [2, 48]).

Age Thickness
Group Formation Lithology Depositional environment
(approx.) Max (m)
Holocene Alluvium Fluvial
Coarse ferruginous sandstone with
layers of quartz pebbles and
Dupi Tila Fluvial
siltstone with lignitic fragments and
Plio-Pleistocene petrified wood 1,600
Girujan Clay Clay and siltstone
Tipam Coarse-grained, pebbly, cross- Alluvial
Tipam Sandstone
bedded sandstone
Fluvial
Late Dark grey pyrite-bearing shale,
Boka Bil 1,600 Tidal-deltaic
Miocene sandy shale, and sandstone
Estuarine
Miocene Surma
Sandstone and pebbly sandstone at
Middle
Bhuban the top and sandy shale at the 1,500 Shallow marine
Miocene
bottom

skinned tectonic processes [2, 17, 18]. These folds are Determination of the bedding to mean pole follows Right
arranged in a set of alternating low relief ridge-forming Hand Rule (RHR), and pi axis orientation (trend and
fault-cored anticlines and valley-forming synclines. The plunge) of each segment was determined using the same
anticlines are continuous along-strike for tens to more than software.
100 km. Moreover, the geometry of the CTFB fold ridges is
attributed to the free westward migration of fold, longitudi- 3.2. Lineament/Fracture Analysis. The primary data set used
nal thrust, and multi-phase deformation of the Tertiary sed- to detect fold-related fracture (i.e., lineaments) are ALOS,
iments [2, 46]. SRTM-based digital elevation models (DEMs), and Google
Earth. After initial processing of the ALOS images (e.g.,
3. Methods spatial enhancement and hillshade) in ArcGIS 10.3 environ-
ments (Figures 5(a) and 5(d)), lineament/fracture demar-
The study connects the outcrops-scale surface geology and cated, and digitized manually through visual interpretation
fault kinematics data with large-scale kinematics and tec- to avoid the manmade features (e.g., road and foot track).
tonics to determine the overall geometry and kinematics During manual extraction, the utmost care has been given
of the CTFB. The bedding attitude data (from fifteen anticli- to maintain the same scale/zooming level. Features more
nal structures), lineament/fracture data (from two struc- than 500 m in length and having structural significance such
tures), fault-slip data (from four structures), exposed as fracture, faults, abrupt changes of river courses, and linear
deformation structures, and stratigraphic and lithologic data scarp faces were digitized as lineament/fracture and stored in
from the CTFB were measured and recorded. A geologic a geographic information system (GIS) database. For the
map of the CTFB area that defines the outer belt was con- comparison and necessary corrections of lineaments
structed (Figure 2). Outcrops are sparsely exposed due to obtained by remote sensing, the resulting lineaments maps
ever-increasing settlement, heavy vegetation cover, and (Figures 5(b) and 5(e)) were cross-checked with the geolog-
rapid erosion. Field data were collected only from the ical map, topographic map, SRTM-based DEM, Google
road-cut, river, and lake outcrops along transects that cross Earth, and field observations to remove roads, lithological
each anticline during several fieldwork conducted in the last boundaries, and any other linear artifacts. After all these
decade. Lithologic and fault contacts detected along each processing and ground checks, a total of 483 and 345 linea-
transect were extrapolated manually on both sides using a ments/fractures were mapped for the Sitakund and Sitapa-
hillshade digital elevation model based on the SRTM images har structures, respectively (Figures 5(b) and 5(e)).
(ver. 3.0 with 1° × 1° tiles at 1 arc second, i.e., 30 m resolu- We used the AMOCADO software [51] to analyze the
tion). Secondary tectonic and structural data sets from the possible relationship between the extracted fracture/linea-
published maps are digitized, scaled, and integrated with ment (Figures 5(b) and 5(e)) anisotropy with the stress filed
the primary field data to get the total coverage of the study of the area. AMOCADO is based on the modified Cantor
area. dust method [52], where a set of scanlines (parallel) are
superimposed on the lineament/fracture pattern. The num-
3.1. Bedding Analysis. To quantify the fold belt geometry, we ber of segments Ns that intercept the lineament/fracture
have analyzed the bedding attitude data (total n = 1356) and segment-length s are automatically measured and plot-
individually for each anticlinal structure and all of them ted as a log cumulative Ns vs. s in the graph to determine
together (Figures 3 and 4). The orientations of the studied the slope m (i.e., fractal dimension). Consequently, the scan-
structure were determined using a cylindrical best fit to poles line set is rotated in 1° incremental steps up to 180° to pro-
to bedding using the software Stereonet 11 (Figure 3) [50]. duce the complete result. Here, the AMOCADO method

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 7

(a) (b) (c)

.8
88
2.0/
23
5.7
.0/8
064
.4 /86.4
067

Barkal n = 25 Belasari n = 18 Gobamura n = 54

(d) (e) (f)

4.7
253.7/8 89.
1
.8/
240

1.5
.3/8
057

Kasalong n = 31 Uttan Chatra n = 16 Bandarban n = 88

(g) (h) (i)

9.4 /89.7
259.1/8 076.4

Changotaung n = 11 Dakhin Nhila n = 202 Inani n = 40

(j) (k) (l)

.9 .8
/87 /87 89.
1
5.7 6.6 .8/
05 23 240

Matamuhuri n = 21 Patiya n = 28 Semutang n = 16

(m) (n) (o)

9.2
.7/8 7.3
068 88.
2
.0/8
.5, 067
245

Sitapahar n = 420 Sitakund n = 287 Olathang n = 77

Figure 3: Stereograms show the Kamb density contour of the poles to bedding (contour interval = 2σ) with a cylindrical best fit line
(magenta solid line), axial plane (magenta dashed line), and fold axis (hollow red circle) for each studied anticline of the CTFB area
(locations are marked in Figure 2). All stereograms are generated with Stereonet 11 through equal-area lower hemisphere projections
[50]. The orientation of the fold axis (hollow red circle) is labeled with trend and plunge. Note: n: number of bedding attitude data;
WFTZ: western fold-thrust zone; and EHCFTZ: eastern highly compressed fold-thrust zone.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
8 Lithosphere

180

33
5.6
160

Fold Interlimb Angle (°)


01
.9
140
2
R = 0.486
120

100

80 Gobamura
88.1 Kasalong
65 .6/ 60
Barkal
40
n = 1356 n = 1356 n = 1356 50 60 70 80 90
Axial Plane Dip (°)
(a) (b) (c) (d)

160
4

120 3
Mean Value

Mean value
2
80
1

40 0

Fold axial trend Interlimb angle Fold axial plunge Shear strain
Fold Attribute Fold Attribute

(e) (f)

CTFB Zone
EHCFTZ
WFTZ
CTFB Total

Figure 4: Combined plot of the bedding attitude data from fifteen anticlines of the CTFB area. (a) Stereograms show poles to bedding (blue
points), (b) the Kamb density contour of the poles to bedding (contour interval = 2σ) with a cylindrical best fit line (solid black), and
regional fold axis (red circle) of the CTFB area. (c) Rose diagram of strike of the bedding attitude data of the CTFB area plotted with 5°
class interval (petals parallel strike direction). All stereograms are lower hemisphere equal-area projections generated with Stereonet 11
[50]. The trend and plunge of the regional fold axis (red circle) is labeled. Note: n: number of bedding attitude data. (d) Fold interlimb
angle vs. an axial surface dip of the same folds. If Barkal, Kasalong, and Gobamura values excluded from linear regression calculation,
the plot reveals a weak linear relationship (R2 = 0:486). In general, gentle folds are more upright compared to the open folds. (e and f)
Mean value and confidence interval of the fold attributes of the WFTZ and EHCFTZ of the CTFB and CTFB as a whole. Error bars
represent 95% confidence interval.

[51] is used to quantify the fractal dimension anisotropy identified in this study and are presented in Figures 5 and
(insets in Figures 5(b) and 5(e)) of the extracted fracture/lin- 6, respectively. The angle of each set (i.e., axial, cross-axial,
eament system of the Sitakund and Sitapahar anticlines. The and oblique fracture sets) was carefully compared with the
FracPaQ, a MATLAB™ open-source toolbox [53], was used fold axial trace. Finally, the bisector of the two oblique frac-
to quantify fracture/lineament patterns (Figures 5(b) and 5 ture sets was regarded as parallel to subparallel of the maxi-
(e)), including their distributions and spatial variation mum shortening direction (i.e., σ1) [57, 58].
(Figures 5(c) and 5(f)).
The rose diagrams (Figure 6) of the extracted lineaments 3.3. Kinematic Analysis of Outcrop-Scale Structures. Kine-
were plotted and compared with the general trend of the matic analysis of the mesoscale deformation structures
CTFB. The total fractures obtained from the remote sensing (Figure 7) was performed based on field geological investiga-
technique were divided into different sets on the rose dia- tions, oriented field photographs, geometric measurements
grams according to their angle and direction relative to the of the bedding and fault-slip data (e.g., rake, displacement),
fold axis [54–56]. If a fracture set has a trend parallel with and their mutual relationship with respective anticlinal
the fold axial trend (i.e., within 0-15° of the fold axial trace), structures of the CTFB.
it is defined as the “axial fracture set.” On the other hand, if a
fracture set is perpendicular to the fold axial trend (i.e., 80- 3.4. Balanced Geological Cross-section. To evaluate the struc-
90° from the trend of the related fold axis), it is defined as tural pattern and shortening of the Tertiary sediments in the
a “cross-axial fracture set.” In this study, the axial fracture outer belt of the Indo-Burma Ranges, we constructed a bal-
set is detonated by S1 (Figure 6(e)). Fractures having greater anced cross-section based on our field data and seismic pro-
angles (i.e., >15° and <80°) with the trend of the fold axis files of the study area from the previous literature [18, 39, 42,
were called an “oblique fracture set.” Two oblique fracture 59]. Three probable balanced sections (Figure 8) are drawn
sets defined with S2 and S3 (Figures 6(d) and 6(f)) were in the northern, middle, and southern part of the study area

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 9

10 10
~3.08
22∘ 50′
b

y-data
0 a
20 20
~–3.08

~–3.57 0 ~3.57
x-data

30 30
22∘ 40′

n = 483 n = 483
0 0
40 40

Y, km

Y, km

S2
1 50

50 50
22∘ 30′

S3
2
100

N 60 3 60

S1
150

0 8 16 Lineaments Lineament
22∘ 20′

km 70 length, km 70 strike, degrees


91∘ 40′ 91∘ 50′ 30 40 50 30 40 50
X, km X, km
(a) (b) (c)

~2.71
22∘ 40′

b
y-data

15 0 15
a
~–2.71
20 ~–3.53 0 ~3.53
20
x-data n = 345
25 0 25 0
n = 345
S2

30 1 30 40
Y, km
Y, km
22∘ 30′

35 2 35 80
S3

40 3 40
120
N 45 45
4
S1

160
0 5 10
km 50 50
Lineaments Lineament
length, km strike, degrees
55 55
92∘ 05′ 92∘ 10′ 92∘ 15′ 10 15 20 25 30 10 15 20 25 30
X, km X, km
(d) (e) (f)

Figure 5: Lineaments extraction and analysis using remote sensing technique in the study area. (a and d) The Shaded Relief map of the
Sitakund and Sitapahar anticlines using Advanced Land Observing Satellite (ALOS) images for lineaments extraction. The location is
marked by the rectangular box on the inset CTFB figure. (b and e) The length-orientation map of the primary fracture system was
obtained using remote sensing techniques from the Sitakund and Sitapahar anticlines, respectively. Note: a and b are long and short axis,
respectively. The inset figures in (b) and (e) are the results of the anisotropy of fractal dimension analysis by AMOCADO of the
Sitakund and Sitapahar fractures, respectively. (c and f) Fracture/lineament strike orientation map of the Sitakund and Sitapahar
anticlines, respectively (prepared using FracPaq). Three distinct fracture/lineament sets (S1-S3) orientations are marked in the color bar.
Note: n: number of extracted lineaments.

along AA ′ , BB ′ , and CC ′ traverse, respectively. In these pro- Here, the cross-section lines of the formation layers and fault
file sections, the lines are drawn perpendicular to the general planes are primarily oriented perpendicular to the principal
trend of the fault planes along the ENE-WSW direction. direction of shortening. Apparent dips of each layer and

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
10 Lithosphere

0∘ 0∘

S2
S2

10%
𝛼 𝛼 5%
1%
S3
1% S3
5%
10%
20%

n = 483 n = 345

S1 S1
180∘ 180∘

0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
Fracture/lineament strike, degrees Fracture/lineament strike, degrees
(a) (b)

Kapti Lake

S2
Ka

:V
Fig (e) ert
rna

ica
l (N
phu

Fig. d, f Shale NE
)
li
Riv

10 km
er

NNW SSE
(c) (d)

S1: S70W/86 S3: Vertical

S0 S0

Shale

Shale

NNW SSE E W
(e) (f)

Figure 6: Rose diagram of the fold-related fracture/lineament trends of the study area. (a) Rose diagram of the fracture/lineaments trends of
the Sitakund Anticline (based on the fractures shown in Figure 5(b)). (b) Rose diagram of the fracture/lineament trends of the Sitapahar
Anticline (based on the fractures shown in Figure 5(e)). S1, S2, and S3 mark three distinct fracture/lineament sets (their mean
orientation marked with the blue lines). Note: n: number of analyzed fracture/lineament. (c) Panchromatic satellite image taken from the
Google Earth Pro showing the Sitapahar Anticline (structure marked with dotted white line). White rectangle represents the location of
the measured lineaments. (d) White arrow on the field photograph marked the S2 lineament (joint). The student in the photograph is
1.75 m height. (e) Smooth fracture surface represents S1 lineament oriented approximately parallel to the fold axis. The student in the
photograph is 1.70 m height. (f) Approximately EW oriented S3 lineament (joint). Bedding plane is marked by the dashed white line
(S0), and axial cleavage is marked by the dotted white line. The student in the photograph is 1.77 m height.

fault planes were calculated from the field data and projected are used, and the surface data is projected downward using
along with strikes on the balanced cross-section. To under- the kink methods [60, 61]. Basal detachment fault is identi-
stand the subsurface structure, available seismic sections fied from the different seismic sections [42]. To construct

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 11

(a) (b)
S0 S0
Fault

Sandstone

Shale
S0

S0 Shale
Sandstone ENE WSW Shale
Shale

WSW Sangu River ENE


S0
(c) CL
Fault
Sangu River WSW
ENE

(f) S0
S0
S0 S0
Fault

CL
Shale Cleavage
Shale
S0 E W
(d)
S0

S0 Shale
S0

S0
Fault

Fault

1m
N20 ∘W S20 ∘E ENE WSW

(e)

S0

Fault

Shale
N60 ∘E/2 ∘ N6
6∘
0 E/4
6∘

S0
3∘
/4
W
6∘
S6

Kaptai Lake
NE

Figure 7: Kinematics of the exposed faults in different anticlinal structures of the CTFB (all locations are marked in Figure 2). (a) Near
vertical thrust plane cross-cut the sandstone (Tipam Formation) and thinly bedded shale unit with calcareous sandstone band (Boka Bil
Formation) in the downstream part of the Sangu River section (east flank of the Bandarban structure). The inset figure in the upper
right corner is the zooming of the rectangular area in (a). The layer parallel broken calcareous sandstone fragment is marked by the
white outline. The length of the hammer in the picture is 0.20 m. Note: In all photos; the thrust plane is marked with the solid pink line;
the bedding plane (S0) is marked with the solid white line; half-arrow indicates the hanging-wall movement direction; and the
stereogram shows poles to the bedding (black), fault (pink), and cleavage (green) planes. (b) West verging and east-dipping antithetic
back thrust in the upstream part of the Sangu River (Bandarban structure). The length of the hammer in the picture is 0.20 m. (c) West
verging and east-dipping thrust-shear zone in the Debchara section (west flank of the Sitapahar Structure). The cleavage plane (CL) is
marked with the white dotted line. The length of the hammer in the picture is 0.20 m. (d) West verging and east-dipping thrust in the
Sahasradhara area of the Labanakhya Chara section (west flank of the Sitakund Structure). The in-sequence thrust plane cross-cuts the
thinly bedded shale unit with a calcareous sandstone band (Boka Bil Formation). (e) Meso-scale fault propagating kink fold in a thinly
bedded shale at the tip of the east verging and west-dipping thrust in the Kaptai Lake section near Shuvolong Bazar Army camp (east
flank of the Gobamura Structure). The green cap student is 1.83 m tall. (f) Small-scale ~northwest verging and ~southeast dipping thrust
in the west flank of the Dakhin Nhila Structure, just east of the marine drive. The in-sequence thrust with dextral strike-slip component
cross-cuts the thinly bedded shale unit (Boka Bil Formation). The inset figure in the upper right corner is the zooming of the rectangular
area in (f). The length of the hammer in the picture is 0.27 m.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
12 Lithosphere

N65∘E WFTZ EHCFTZ


A Sitakund A'
anticline Changotaung Gobamura Kasalong Shishuk
Semutang anticline anticline anticline anticline
CCF
anticline
0
2.5
5.0
7.5
km 0 20 40 60 80 90 km

(a)

WFTZ EHCFTZ
Sakudaung N60∘E Dakhin Nhila
B
N70∘E
Bandarban Anticline C’
anticline B’ C
anticline
Jadi Matamuhari
anticline 0
anticline
0 2.5
2.5 5.0
5.0
7.5
7.5 km 0 5
km 0 20 40 60 km (c)

Lower Middle Bhuban and


Tipam Sandstone
Pre-Miocene Sediments
Dupi Tila Sandstone and
Upper Bhuban Quaternary Holocene Deposits

Boka Bill

(b)

20
A A’
y = 0.146x + 2.76
R2 = 0.98
15
Shortening (km)

10 Shishuk
Kasalong
Sitakund Gobamura
Changotaung
5
Semutang
y = –0.0135x + 3.43
R2 = 0.27
0
0 10 20 30 40 50 60 70 80 90 100
Horizontal distance (km)
Individual shortening
Cumulative shortening
(d)

Figure 8: Deformed state geological cross-section of the CTFB, Bangladesh (i.e., the outer edge of the Indo-Burman Ranges). (a, b, and c)
Balanced section along the profile line AA ′ , BB ′ , and CC ′ shown in Figure 2. Note: Decollement has been marked with thick pink line and
the depth of the decollement adopted from Bürgi et al. [42]. (d) Individual and cumulative shortening of the CTFB anticlines along the
section AA ′ .

the cross-section, it is assumed that there was no transport of (Bandarban, Sitapahar, Gobamura, and Dakhin Nhila) were
material; hence, the volume was maintained during the used for paleostress determination (Figure 9). Common
deformation as the domain suffered major tectonic deforma- kinematic indicators such as Riedel shears, shear bands,
tion after the deposition of Tipam Formation. and bedding offset (generally well preserved ∼0.1–1 m of dis-
placement between fault blocks) were used to decipher the
3.5. Fault-Slip Analysis/Paleostress Analysis. All fault-slip sense of slip.
data (orientations of fault planes and slip vectors) recorded Following Angelier [62], a total of 46 fault data are col-
during the field works in the four different structures lected from the study area and classified as an oblique-slip

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 13

Schmidt Lower Schmidt Lower


R. Dihedron N Weight Mode 2 R. Dihedron N Weight Mode 2
n/nt: 20/22 n/nt: 8/11

σ1: 08/095 𝜎1: 07/256


σ2: 07/186 𝜎2: 04/347
σ3: 79/317 𝜎3: 82/103
R: 0.88 CD: 20.3 R: 0.5 CD: 22.2
QRw: D QRt: D QRw: D QRt: D
Counting deviation Counting deviation
60 60

30 30

0 0
0 Sum of Weights 51 0 Sum of Weights 11

(a) (b)

Schmidt Lower Schmidt Lower


R. Dihedron N R. Dihedron N Weight Mode 2
Weight Mode 2
n/nt: 4/4 n/nt: 11/11

𝜎1: 01/239 𝜎1: 05/249


𝜎2: 04/329 𝜎2: 13/340
𝜎3: 86/140 𝜎3: 76/137
R: 0.5 CD: 6.7 R: 0.64 CD: 19.3
QRw: E QRt: E QRw: C QRt: C
Counting deviation Counting deviation
60 60

30 30

0 0
0 Sum of Weights 13 0 Sum of Weights 21

(c) (d)

Figure 9: Results of paleostress analysis using the right dihedron method. (a, b, c, and d) Results of paleostress analysis of faults recorded
from Bandarban, Dakhin Nhila, Gobamura, and Sitapahar, respectively. Blue and green arrows are for maximum (σ1) and minimum (σ3)
principal compression, respectively. The histograms give the counting deviation. Note: n and nt are the numbers of data accepted for tensor
calculation and the total number of data, respectively. Stress ratio, R = ðσ2 − σ3 Þ/ðσ1 − σ3 Þ. QRw and QRt give a quality ranking index of
the data.

thrust fault. There are several methods suggested for paleos- subsets using the software program “Win_Tensor” (version
tress investigation using fault-slip data (e.g., [62–79]). The 5.8.6; [82, 83]). This process eliminates all probable hetero-
paleostress analysis using fault-slip data can be done using geneities from the initial data. All the data are presented in
kinematic or dynamic methods based on the following a single data set, and the initial separation has been com-
assumptions [67, 77, 80]: (1) The bulk state of stress is pleted using the Right Dihedron method. Further, the data
assumed to be constant, and any displacement on the fault is filtered based on stress ratio (R), stress axes orientation,
planes are independent; (2) the slip on the fault plane ensues and symmetry of the measured sets. Following Delvaux
the direction of resolved shear stress maximum under a cer- and Sperner [83], the best possible data are recognized with
tain state of stress condition (Wallace–Bott Hypothesis); and a low value of “counting deviation” and “nominal counting”
(3) the faults are believed to be homogeneous and formed by values of 0 and 100 for σ1 and σ3, respectively. The detailed
the same tectonic event [62, 67, 77, 81]. We resolved the result of paleostress analysis is presented in Section 4.5.
paleostress condition in this study, using fault-slip data col-
lected from the studied region by Right Dihedron methods. 4. Results
Although some of the fault planes show a wide range of
strike orientations, most of them are oriented along the 4.1. Bedding Geometry Analysis. Bedding attitude measure-
NNW-SSE (maxima) or NNE-SSW (submaxima). The col- ments (n = 1356 data points) of 15 anticlines show almost
lected fault data sets are arranged into homogeneous data consistent NNW strike with variable dip amounts

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
14 Lithosphere

Table 3: Geometric fold data from the fifteen anticlines of the CTFB area. Note: IA = fold interlimb angle; APD = fold axial plane dip;
APDD = axial plane dip direction.

Interlimb angle Average IA Axial plane dip, α (°); Average APD Shear strain,
CTFB fold zones Fold name
(°) (°) APDD (°) γ
Barkal 58.6 83.9; W 0.11
Belasari 96.6 75.8; W 0.25
Eastern highly compressed
Gobamura 91.6 96.82 83.5; W 83.70 0.11
fold-thrust zone
Kasalang 84.0 87.1; E 0.05
Utan Chatra 153.3 88.2; W 0.03
Bandarban 123.0 77.9; W 0.21
Changotaung 111.2 88.9; W 0.02
Dakhin
119.2 82.7; W 0.13
Nhila
Inani 161.0 86.6; W 0.06
Western fold-thrust zone Matamuhari 145.2 125.19 87.7; E 84.20 0.04
Patiya 147.6 87.0; E 0.05
Semutang 155.5 89.5; W 0.009
Sitapahar 100.4 87.6; E 0.04
Sitakund 75.4 67.6; E 0.41
Walayataung 113.4 86.5; E 0.06

(Figure 3). Only a few of these measurements are shown in zone (Utan Chatra, eastern zone; Sitakund, western zone)
Figure 2 due to limited space. Table 3 summarizes the main shows different results (Table 3).
geometrical features of the fifteen anticlines from the two A combined analysis of the entire bedding attitude mea-
zones of the CTFB area, and Figure 3 represents the density surements (n = 1356) of the CTFB area shows an axial plane
contour of the poles to bedding with a cylindrical best fit line orientation (Figures 4(a)–4(b)) and strike trend (Figure 4
axial plane and fold axis for each studied anticline of the (c)). A weak linear relationship has been revealed based on
CTFB area. Figure 4(d) presents plots of fold interlimb angle graphical analysis of the fold interlimb angle vs. an axial sur-
vs. axial plane dip. However, for the eastern highly com- face dip of the same folds (Figure 4(d)), except for Barkal,
pressed zone, plunges vary from 0.9° to 5.3°, and for the Kasalong, and Gobamura. Shear strain, γ, has been calcu-
western fold-thrust zone, plunges vary from 0.3° to 8.5°. lated for the fifteen studied folds of the CTFB area based
The standard deviation of the trend and plunge of the fold on the following equation: γ = ½tan ð90° − αÞ. Here, α is
axis is 7.93° and 2.21°, respectively, (Figures 4(e) and 4(f)). the dip of the axial plane [84]; the result is reported in
Out of fifteen anticlines (Figure 3), about a half (seven) Table 3 and shows a considerable variation of the shear
plunges to SSE, while another half (eight) plunges to strain from 0.009 to 0.41 (Supplement 1, Figure S1). The
NNW. highest value (0.41) has been calculated for Sitakund
The dip of the axial surface of the studied anticlines var- Anticline, whereas the lowest value is calculated for
ies from 68° to 89°. For the eastern highly compressed fold- Semutang Anticline, and interestingly, both are situated at
thrust zone, the dip of the axial plane varies from 76° to the western fold-thrust zone.
88°, and for the western fold-thrust zone, the axial plane
dip varies from 68° to 89°. The standard deviation of the 4.2. Fracture/lineament Pattern Analysis. Elongated folds
axial plane dip is 6.00. Out of fifteen studied anticlines, six dominate the structural geology of the CTFB with minimum
of their axial planes dip to the east, whereas the remaining aspect ratio (i.e., length of the hinge line divided by the
axial planes dip to the west (Figure 3, Table 3). Interestingly, width of the fold) of more than 3 (Figure 2). Fold-related
the axial plane of most of the anticlines of the eastern highly fractures/lineaments have been analyzed from the two anti-
compressed zone is dipping to the west except one, whereas clines within the western fold-thrust zone and had a geomet-
in the western fold-thrust zone, both west and east-dipping rical relation to the fold elements. For both anticlines, three
axial planes have been observed. distinct fracture/lineament sets (S1-S3) were identified.
The interlimb angles of the analyzed folds show a dis- These fracture/lineament sets can be attributed to syn-
tinct difference between the eastern highly compressed folding and/or post-folding regimes. The chronological and
fold-thrust zone and the western fold-thrust zone geometrical studies can reveal the relation between fractures
(Figures 4(e) and 4(f)). Generally, the folds in the western and the fold [85, 86]. The fold-related fractures are visual-
zone show a larger interlimb angle (average is 125°), whereas ized using the rose diagram (Figure 6).
folds in the eastern compressed zone have a smaller inter- Based on these diagrams and following the RHR, for
limb angle (average is 97°). However, one fold from each Sitakund anticline, S1 fractures include the NNW-SSE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 15

Table 4: The results of the AMOCADO analysis of the fracture system of Sitakund and Sitapahar anticlines of the CTFB area.

Axial ratio or azimuthal Angle between north and


Name of the structure Coefficient of determination (R2)
anisotropy (a/b) the long axis (a)
Sitakund 1.16 82° 0.934
Sitapahar 1.27 79° 0.890

fracture set with the mean direction of N155°±5, the S2 frac- underthrusted by the east-dipping and west verging CCF
tures include NE-SW fracture set with the mean direction of (Figure 2).
N45°±5 and S3 fractures with almost E-W orientation with
the mean direction of N95°±5 (Figures 5 and 6). For Sitapa- 4.4. Balanced Geological Cross-Section. The field data sets
har anticline, S1 fractures include the NNW-SSE fracture set that were used to generate the balanced cross sections are
with the mean direction of N155°±5. The S2 fractures presented on the map (Figure 2) and stereographic projec-
include NE-SW fracture set with the mean direction of tions (Figure 3), accompanied by a description of field obser-
N37°±5 and S3 fractures with almost E-W orientation with vations in Table 3 and a field photo showing the structural
the mean direction of N90°±5 (Figures 5 and 6). With relationship in Figure 7. Sections are drawn parallel to the
respect to the Sitakund and Sitapahar fold axial trace, these principal direction of shortening to estimate the shortening
fractures are divided into two major groups: (a) axial frac- of the Tertiary sediments. Three ENE-WSW oriented bal-
tures S1, which are parallel to the fold axial trace, and (b) anced cross-sections from the northern (AA ′ ), middle
oblique fractures S2 and S3 with an angle of 120° and 65°, (BB ′ ), and southern part (CC ′ ) of the study area
respectively, relative to the fold axial trace (Figure 6). The (Figure 2) were constructed to examine the tectonic features
acute angle (α) of the two intersecting S2 and S3 fracture sets on the surface and the subsurface across the CTFB (Figure 8
is 50°±5. and Figure S2 of Supplement 2). Sections are drawn parallel
The results of the AMOCADO analysis of the fracture to the principal direction of shortening to estimate the
system are presented in Figure 5 and Table 4. The axial shortening of the Tertiary sediments (Figure 8(d)). Here,
ratios (i.e., a/b) of the fractal dimension anisotropy ellipse, most of the anticlines are bounded by single thrust fault on
which represents anisotropy intensity for the fracture system both sides except the Sitakund anticline (Figure 8(a)). The
of the Sitakund and Sitapahar anticlines are found to be 1.16 western flank of the Sitakund anticline is affected by
and 1.27, respectively. The azimuthal trends of the major imbricate thrust faults. Therefore, the Sitakund anticline is
axis (a) for Sitakund and Sitapahar anticline fracture systems more tightly folded than the other folds in the western
are found to be 82° and 79°, respectively, which are approx- fold-thrust zone. All the anticlines show gentle to open
imately orthogonal to their respective anticline trend fold in the surface, and the interlimb angle decreases with
(Table 4). depth as well as towards the east. The shortening estimates
in the northern part of the study area along AA ′ are
~16 km for the 94 km section. In the central part along BB ′
4.3. Deformation Structures and Their Kinematics. Among
section, the shortening is ~4.2 km for the 68 km section. In
the two broad zones of the CTFB, brittle (fault, joint, and
the southern part along CC ′ section, the estimated
cleavage) and brittle-ductile (boudins and shear bands)
shortening is ~1.5 km for the 7.5 km section. The
deformations are dominant in the eastern highly compressed
percentages of shortening of the Tertiary sediments in AA ′ ,
fold-thrust zone compared to the western fold-thrust zone.
Only compressional/transpressional deformation has been BB ′ , and CC ′ sections are ~15%, ~6%, and ~16.7%,
observed in the two deformation zones of the CTFB. The respectively (Supplement 3, Figures S3 and S4).
observations from field and satellite images on different
scales suggest a fold parallel to subparallel thrust with dom- 4.5. Paleostress Analysis. The stress tensor obtained by pro-
inant NNW-SSE to NNE-SSW trend (Figure 7). cessing all fault data from CTFB gives a NE-SW to ENE-
Both the east- and west-dipping thrusts at different WSW directed compression. In Figure 9, a relative position
scales with highly variable dipping angles are omnipresent of the principal stress axes and stress ratio (R) is provided
in the study area (Figure 7). Observed displacement along (also see Supplement 4, Tables S1 and S2). It may be noted
the fault plane ranges from a few cm to tens of m. Evidence that the maximum principal stress axes in all locations are
of synthetic (Figure 7(a)) antithetic (Figure 7(b)) thrusts, found to be horizontal, while the minimum principal stress
thrust-shear zone (Figure 7(c)), fault propagating kink fold axes are vertical, consistent with a compressive
(Figure 7(e)), and thrust fault with dextral shear (Figure 7 environment. This suggests that all the faults recorded in
(f)) are present in both deformed zone of the CTFB, but the study area result from thrust due to NE-SW to ENE-
dominantly to the eastern zone. The age and relief of the WSW directed compression. The fault-slip analysis also
exposed Tertiary rocks increase towards the east, and the helps to analyze the tectonic stress regime, including any
oldest exposed formation in the study area is the Bhuban variation in the stress regime. The stress regime is
Formation, mainly exposed at the eastern anticlines of the determined with the help of the stress regime index (R ′ )
CTFB. However, the exceptions are the Sitakund, Inani, invoked in the program “Win_Tensor.” The R ′ is related
and Dakhin Nhila anticlines, whose western flanks are to R values depending on the relative orientations of the

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
16 Lithosphere

principal stress axes and includes a range of values. Delvaux the westward migration of fold, multiphase deformation of
et al. [87] have quantified the R ′ values in detail for various soft sedimentary rocks, and longitudinal thrust as well as
stress regimes. These stress regimes are generally recognized transverse faulting [2, 88]. In the western fold-thrust zone,
as radial extension, pure extension, transtension, pure the curvature of the anticlines is higher than the synclines,
strike-slip, transpression, pure compression, and radial whereas in the eastern highly compressed zone, the curva-
compression [72, 83]. The stress regimes for all faults have ture of the synclines and the anticlines are more or less sim-
been evaluated based on the R ′ value and the principal stress ilar (Figure 8). It indicates that the development of anticlines
orientations (see Supplement 4, Table S1). The stress regime in the western fold-thrust zone is more fault-controlled,
indicates that in all the locations, the faults lie within the while the anticlines in the eastern highly compressed zone
pure compressional domain. are superposed in nature and result from buckling as well
as fault propagation. This is a typical structural style in
5. Discussion fold-and-thrust belts [89–92].

5.1. Spatial Variation of Anticlines in Chittagong Tripura 5.2. Competence of the Folded Rock Units and Folding
Fold Belt. The western Indo-Burman Ranges (i.e., CTFB) Mechanism. At low to very low metamorphic levels, Ramsay
has developed an exemplary linear NNW-SSE oriented dou- [93] classifies the common rock units from the highest (with
bly plunging fold belt. These folds are arranged in a set of value 1) to the lowest competence (with value 10). As the
alternating valley-forming synclines and ridge-forming anti- exposed Tertiary rock units in the core of the CTFB
clines [2, 88]. To the west, the surficial expressions of the anticlines are at a very low metamorphic level, therefore,
CTFB folds are limited chiefly along with the Chittagong Ramsay’s [93] classification can be applied. The folded rock
Coastal Fault (CCF). units in the CTFB area are mainly quartz sandstone (compe-
The bedding attitude analysis (Figure 3) reveals that the tence value 3), siltstone (competence value 7), and shale
axial planes of the eastern highly compressed zone anticlines (competence value 9). In the CTFB area, the presence of less
are dipping primarily to the west whereas the axial planes in competent rock units (e.g., shale, mud, and silty shale) in the
the western fold-thrust zone dipping both east and west. The form of separation surfaces and lateral changes in sedimen-
larger interlimb angles were observed in the western zone tary facies can cause the overall changes in fold wavelength,
(average 125°) compared to that of the eastern compressed amplitude, and style (Figure 3).
zone (average 97°) except for Utan Chatra and Sitakund The mechanical properties of rock units involved in fold-
anticlines (Table 3; Figure 3). A weak linear relationship ing have profound effect on the fold geometry. If similar
exists between fold interlimb angle and axial surface dip of sized-layers of incompetent and competent rock units are
the same folds (Figure 4(d)), except for the Barkal, Kasalong, forced to deform at a given rate, the differential stress will
and Gobamura, which are bent down below, possibly due to less in the incompetent unit compared to the competent
proximity to the Kaladan Thrust. Although the calculated unit. As the rock units in the study area are subjected to layer
shear strain (γ) of the CTFB area shows a small variation parallel compressive stress, buckling should be dominant
(Table 3), anticlines in the eastern zone have a relatively mechanism. In fold-thrust belts, the compressive stress that
higher shear strain than the western zone. Based on the drives thrusting causes the rock units to buckle and thereby
interlimb angle and shear strain of the studied anticlines, it shortening and thickening of the rock units. The buckling
is evident that the intensity of the folding in general can occur either through flexural shear/flow and/or flexural
decreases towards the west: the direction of the fold belt slip depending on the competency of the rock units. Accord-
propagation (Figures 4(e) and 4(f)). The exceptional values ing to Donath and Parker [94], the type of mechanism of
for the few anticlines in the western zone are possibly related folding is related to the competence contrast and mean com-
to the deformation caused by CCF. The CCF is a major petence of the rock units involved in folding. In the CTFB
boundary thrust known as “Thrust Front” that separates area, the average competency of the rock units (Table 2) is
the “Emerging Fold Belt” to the west and CTFB to the east medium to high, and the competence contrast is also
[17, 20]. The western flank and part of the axial zone of a medium to high as compared to Ramsay’s [93] classification.
few of the anticlines (e.g., Sitakund, Inani, and Dakhin Hence, by considering the Donath and Parker [94] kine-
Nhila) are highly deformed and under thrusted by the CCF matic model and Ramsay’s [93] competence classification,
and, therefore, show deviation from the general fold param- the dominant mechanism of folding in the CTFB area is pos-
eter results (Figure 8(a)). Overall, Hossain et al.’s [44] classi- sibly a flexural shear with some influence of a flexural slip.
fication of CTFB into two broad zones seems to be
supported by the current study with few exceptions. In the 5.3. Development of Brittle Structures and Paleostress
balance geological cross-section, it is also evident that there Conditions. In a fold-thrust belt, two major fracture systems
is a sharp variation in the spacing of anticlines. At the west- are generally developed, which are fold-related and fault-
ern fold-thrust zone, spacing between anticlines is higher related fractures [55, 56, 95]. Fractures with different orien-
than the eastern highly compressed zone, whereas the curva- tations can be attributed to changes in the stress field during
ture of the synclines is higher in the eastern highly com- folding [95]. Fold-related fractures may be axial, cross-axial,
pressed zone. and oblique fractures [55, 56]. In the CTFB area, a good
Multicurved geometry of some of the CTFB fold ridges relation prevails between fold-related fractures and the fold
(e.g., Sitapahar, Changotaung; Figure 2) are attributed to axes (Figure 6). The S1 fracture set (i.e., axial extensional

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 17

fracture) has the mean direction of N120°. Fold-related con- sional axis that creates the fold is equal to the maximum
jugate oblique fractures set S2 and S3 have the trend of principal stress axis (σ1). This compressive axis is aligned
~NNE and ~E, respectively. The bisector azimuth of these perpendicular to the axial plane of the folds. The average
two conjugate fracture sets (i.e., S2 and S3) is almost perpen- trend of the fold axial plane is ~155° (see Table S2;
dicular to the mean trend of the axial fracture set [95, 96]. Figure 10(d)), and therefore, the trend of the maximum
The geometric relation between these fracture sets and principal stress (σ1) is ~65°, which supports the result
axial plane suggests that the maximum shortening stress obtained through other methods.
(σ1) that formed the CTFB anticlines also formed the frac-
tures/lineaments simultaneously. Therefore, it can be argued 5.4. Deformation Kinematics vis-à-vis Regional Tectonics.
that the fracture sets S2 and S3 are the members of a conju- The mesoscale analysis of deformation structures performed
gate shear system, where the horizontal shortening is per- in this study is mainly based on the field kinematics of the
pendicular to the S1 fracture set. If so, then the σ1 (i.e., observed thrust faults (Figure 7 and Figure S5 of
maximum principal stress) bisects the acute angle between Supplement 5) and their relation with the map patterns of
S2 and S3, while the σ3 (i.e., minimum principal stress) the anticlines (Figure 2). The presence of approximately
bisects the obtuse angle between the fracture planes. How- east and west verging thrust parallel to subparallel of the
ever, the underlying assumption is that the fractures which NNW-SSE trending fold indicates the west-verging fault-
formed in homogenous materials under homogenous stress cored anticlines with ENE a subhorizontal shortening. In
field and in a single deformation event do not significantly general, bedding dip distributions, fault kinematics data,
perturb the stress field in their neighborhood and their ori- and overall map patterns indicate that the subhorizontal
entation has not rotated significantly since their commence- shortening axes are approximately perpendicular to the
ment [55, 57, 58, 74, 97]. In this study, the S2 and S3 axial trace of the CTFB (Figures 2 and 10). The faults
conjugate shear fractures in the CTFB area are probably observed locally in both western and eastern zones of the
formed simultaneously under the same stress conditions. study area (Figure 2) are most likely related to ENE
As these fracture sets were formed as a conjugate pair, the subhorizontal shortening due to westward propagation as
orientation of the bisector angle of the two sets (S2 and S3) the accretionary prism of the Indo-Burmese oblique
of the Sitakund (70°) and Sitapahar (63°) would show the subduction zone [2, 49]. In general, the deformation
maximum principal compressive stress (σ1). The mean pattern shows more complexity in the eastern zone and
shortening (σ1) direction measured from the fracture/linea- gradually becomes less intense towards the propagation
ments of the two anticlines is 67°, which is in good agree- front of the western IBR, i.e., CTFB, and the observation is
ment with the shortening direction (65°) determined from consistent with previous studies [17, 20]. The fault-slip
the axial plane orientation. analysis also helps analyze the tectonic stress regime and
The directions of maximum anisotropy intensity of the the detailed account of the stress tensor orientation. The
Sitakund and Sitapahar anticlines fracture systems are 82° tectonic stress regime of the four anticlines has been
and 79°, respectively, which are approximately parallel to determined (see Section 4.5), revealing the variation in
the mean shortening direction maximum compressive stress stress regime in the CTFB from the northern to the
(σ1). This AMOCADO-based results of maximum fractal southern part. The evaluation of the stress regime
anisotropy intensity direction (insets of Figures 5(b) and 5 (Supplement 4, Table S1) for the respective faults indicates
(e)) reasonably fit with the regional compressional direction that most faults lie within the pure compressional domain.
of the CTFB (Figure 10). The obtained stress regime also fits well with the regional
The fault-slip data collected from the four different anti- tectonics of the CTFB [1, 3, 99].
clines (i.e., Gobamura, Sitapahar, Bandarban, and Dakhin Three balanced geological cross-sections are drawn
Nhila) of the fold belt (Figure 2) are used to reveal the along the mean compressive stress direction in the northern,
paleostress direction which prevailed during deformation. middle, and southern part of the study area (Figure 2), and
It may be noted that Gobamura is situated in the eastern the estimated shortening percentage of the Tertiary sedi-
compressed zone, while Sitapahar, Bandarban, and Dakhin ments are ~16 km (15%), ~4.2 km (6%) and ~1.5 km
Nhila are from the western fold-thrust zone of the CTFB. (16.7%), respectively. This variation may be due to the fold
The results reported in Section 4.5 reveal an NW-SE to morphology of the different sedimentary layers. The north-
ENE-WSW directed maximum compression, which remains ern AA ′ section covers part of the eastern highly com-
consistent in all the four anticlines mentioned above. It is pressed zone where the folds are open to gentle and the
envisaged that all the thrust faults are related to CTFB defor- overall compression rate is high. The middle and the south-
mation on account of NW-SE to ENE-WSW directed maxi- ern part of the study area represent the western fold-thrust
mum compression, which is in good agreement with the zone where the folds are gentle and the percentage of short-
paleostress direction obtained from conjugate fracture set ening decreases. In the present study, the rate of shortening
as well as fractal anisotropy intensity direction. is not measured because data on thermochronology and geo-
Moreover, it is possible to determine the paleostress of a chronology is not available. Maurin and Rangin [18] esti-
folded area based on the axial plane orientation of the fold mated a crustal shortening of 11 km in the IBR with a
[98]. The axial plane orientation of the mapped (Figure 2) 5.5 mm/yr rate along a composite section. Betka et al. [17]
and analyzed anticlines (Figure 3) has been used to deter- reported a westward propagation rate of ≥15 km/Myr. The
mine the paleostress of the CTFB. In general, the compres- estimated shortening of the Tertiary sediments in the

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
18 Lithosphere

92°E 93°E
(a)

23°N
(b)
𝜎

22°N
(c)

21°N

(d)
Anticline
𝜎
Syncline
Thrust
Fault
Stress Filed (World Stress Map)
Paleostress (Fault)
Paleostress (Fracture/Axial-Plane)

Figure 10: Kinematic model of the CTFB area. (a) Simplified tectonic map of the CTFB (after [20]). Maximum principal stress orientations
(i.e., σ1) are taken from the World Stress Map database released in 2016 [99]. White arrow: GPS-derived velocity field (in mm yr-1) of the
present-day active convergence of the Indian Plate to the Shan Block (from [3]). Major earthquakes with magnitude ≥4.5 are marked with
filled black circles (after [2]). Earthquake focal mechanism is plotted as focal spheres/beach ball diagrams (magenta color: [104]; blue color:
Global Centroid-Moment-Tensor, CMT Catalog). Paleostress orientations (fault) are based on Figure 9. (b and c) Paleostress based on the
acute angle of the conjugate lineament/fracture sets the orientation of Sitapahar and Sitakund anticlines, respectively. The pink line marks
the S2 and S3 acute bisector orientation, equivalent to the maximum compressive stress (σ1) direction. Note: n: number of analyzed fracture/
lineament. (d) Great circle and pole density contour of the fold axial plane orientation (based on Figure 3). Arrow marked the orientation of
the maximum compressive stress (σ1). Note: n: number of the analyzed fold axial plane.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 19

present study is quite similar to the previous works. The bal- wedge fold belt are usually trench parallel [101, 102]. There-
anced cross-sections of the present study suggest that the fore, we can rationally conclude that the CTFB absorbs nor-
CTFB is a thin-skinned fold-thrust belt that reflects an active mal component (with respect to the fold belt orientation) of
accretion of deeper Foredeep basin sediments of the Bengal India-Burma oblique vector velocity motion and developed
Basin above a ~5–5.5 km deep décollement. We suppose that as a nearly fully partitioned accretionary wedge. We
the décollement, which is at ~7–7.5 km depth in the eastern assumed that the partitioning between the outer belt of
part of CTFB [42], is laterally continuous and rises to ∼2 km the IBR (i.e., CTFB) and the inner belt occurred along
shallower in the western margin of the CTFB. Betka et al. the Kaladan Fault (Figure 1).
[17] also suggest that the décollement laterally continuous
to the west and rises to a shallower level.
Structural data and kinematic analyses of this study sug- 6. Conclusions
gest that the CTFB was developed due to an approximately
east-trending subhorizontal shortening that is normal to Folds, fault-bounded anticlines, thrusts, and factures/linea-
the fold axial plane (Figures 2, 3, 9, and 10). Although ments are the major deformation structures of the outer belt
Maurin and Rangin [18] reported a dextral shear in the (i.e., CTFB) of the IBR. The development of the anticlines in
western IBR, we found no such field evidence in the IBR the western fold-thrust zone of the CTFB is more controlled
outer belt (i.e., CTFB). Our geologic mapping and paleos- by buckling than the anticlines in the eastern highly com-
tress analysis (see Supplement 4, Table S2) in the CTFB pressed fold-thrust zone, which are superposed in nature
area revealed no evidence of significant transpression or and developed firstly due to buckling and later through fault
strike-slip faulting in the study area. The measured ENE propagation mechanism.
shortening orientations based on the fold axial plane, The geometric relation between these fracture sets, axial
fractures/lineaments (Figure 10), fracture anisotropy plane, and axial trace suggests that the maximum shortening
(Figure 5), and paleostress analysis (Figure 9) is consistent stress (σ1) direction measured from the fracture/lineament
with the maximum horizontal stress orientations derived of the two anticlines is 67°, which is in good agreement with
from earthquake focal mechanism solutions and with the the shortening direction (65°) determined from the orienta-
absolute plate motion direction of the Indian Plate to tion of the fold axial plane. The stress tensors deduced from
Eurasian Plate (Figure 10) in the CTFB area. Though only fault-slip analysis of four anticlines show an NW-SE to ENE-
one earthquake focal mechanism solution in the middle of WSW directed maximum compression and remain consis-
the area shows evidence of dextral strike-slip (see in tent for the analyzed anticlines of the area. It is envisaged
Figure 10), all other evidences suggest compression rather that all the thrusts of the CTFB are related to the NW-SE
than transpression is dominant in the study area. The to ENE-WSW directed maximum compression, which is in
study conducted by Betka et al. [17] in the Tripura Fold good agreement with the other analyzed results in this study.
belt area and Wang et al. [19] in the Myanmar region also The geologic shortening percent of the Tertiary sedi-
found no evidence of dextral shear in the IBR outer belt. ments are ~13.8%, ~8.4%, and ~10.2% for the northern,
Moreover, the characteristics of synthetic Fault Plane middle, and southern parts, respectively, of the outer fold-
Solution (FPS) from a recent dynamic model involving belt component. Kinematics analysis also suggests that the
India-Burma oblique motion are consistent with those CTFB is a thin-skinned fold-thrust belt that reflects active
derived from our observations [6]. The synthetic source accretion of the Foredeep basin sediments of the Bengal
mechanisms from the dynamic model reveal reverse Basin above a deep regional décollement. Geologic mapping
faulting in the CCF and Kaladan faults. and kinematic analysis of the present study revealed no evi-
Our observations are also consistent with the shallow dence of significant transpression, which is consistent with
seismic events, maximum principal stress orientations (i.e., previously published studies and, therefore, suggest com-
σ1) extracted from the World Stress Map database [99], plete strain partitioning.
and GPS-derived present-day stress field (Figure 10). The The interpreted tectonic stress regime and resulted fold-
apparent inconsistency between the oblique GPS vector thrust system of the region will significantly improve the
velocity motion of the Indian Plate in the Bengal Basin understanding of deformation of any young and active oro-
(Figure 1) and the geologic shortening axes normal to the genic fold-thrust belt. Finally, it is the first attempt to pro-
structural trend of the CTFB (Figure 10) might result from vide a detailed geological mapping of the CTFB along with
strain partitioning of dominantly pure shear regimes per- a comprehensive analysis of the exposed deformation struc-
pendicular to the frontal/outer IBR margin structures. The tures and their stress regime. This study would help to fur-
structural analyses suggest a nearly complete strain parti- ther develop the kinematic and dynamic model of the
tioning in the frontal part of the IBR (i.e., the CTFB). The outer IBR in terms of present-day collision of the eastern
component of the plate motion which is normal to the fold and northeastern parts of the Indian Plate.
belt is absorbed by the CTFB, whereas the component paral-
lel to the fold belt is taken up along ~NS-striking major dex-
tral faults (i.e., Sagaing Fault, Churachandpur-Mao Fault, Data Availability
and Kabaw) in the inner part of the IBR [5, 17, 100]. A sim-
ilar strain partitioning has also been reported at other obli- All analyzed data are accessible at https://osf.io/x6ye9/?
que subduction zones where the axial traces of accretionary view_only=84091640f750442781e3f2ad8716727f.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
20 Lithosphere

Additional Points intermediate, and minimum principal stress axes; R = stress


ratio ðσ2 − σ3Þ/ðσ1 − σ3Þ; R ′ = stress regime index. Table
Key Points. (i) We present a complete structural map of the S2: axial plane trend of the anticline and paleostress orien-
Chittagong Tripura Fold Belt (CTFB) of the IBR outer belt tation (maximum principal stress, σ1). (Supplementary
for the first time. (ii) The maximum compressive stress Materials)
(σ1) shows eastward shortening normal to the fold axial
plane, and the estimated shortening is ~10.8%. (iii) Due to References
strain partitioning, the CTFB absorbs the component of
plate velocity motion that is normal to mean anticlinal trend. [1] V. K. Gahalaut, B. Kundu, S. S. Laishram et al., “Aseismic
plate boundary in the Indo-Burmese wedge, northwest Sunda
Conflicts of Interest Arc,” Geology, vol. 41, no. 2, pp. 235–238, 2013.
[2] M. S. Hossain, M. S. H. Khan, K. R. Chowdhury, and
The authors declare that they have no conflicts of interest. R. Abdullah, “Synthesis of the Tectonic and Structural Ele-
ments of the Bengal Basin and Its Surroundings,” in Tectonics
Acknowledgments and Structural Geology: Indian Context, S. Mukherjee, Ed.,
pp. 135–218, Springer International Publishing AG, 2019.
The study is financially supported by the National Natural [3] R. Mallick, E. O. Lindsey, L. Feng, J. Hubbard, P. Banerjee,
Science Foundation of China (41888101 and 41822204), and E. M. Hill, “Active convergence of the India-Burma-
the Science and Technology Major Project of Xinjiang sunda plates revealed by a new continuous GPS network,”
Uygur Autonomous Region, China (2021A03001) and the Journal of Geophysical Research: Solid Earth, vol. 124, no. 3,
Jahangirnagar University Research Grants through the Fac- pp. 3155–3171, 2019.
ulty of Mathematical and Physical Sciences. Md. Sakawat [4] M. S. Steckler, S. H. Akhter, and L. Seeber, “Collision of the
Hossain was supported by the CAS PIFI Visiting Scientist Ganges-Brahmaputra Delta with the Burma Arc,” Earth and
Fellowship (2019VCB0011) at the Institute of Geology and Planetary Science Letters, vol. 273, no. 3-4, pp. 367–378, 2008.
Geophysics, Chinese Academy of Sciences. A special thanks [5] M. S. Steckler, D. Mondal, S. H. Akhter et al., “Locked and
to the China-Pakistan Joint Research Center on Earth Sci- loading megathrust linked to active subduction beneath the
ences for supporting the implementation of this study Indo-Burman ranges,” Nature Geoscience, vol. 9, no. 8,
(131551KYSB20200021). pp. 615–618, 2016.
[6] I. Vorobieva, A. Gorshkov, and P. Mandal, “Modelling the
Supplementary Materials seismic potential of the Indo-Burman megathrust,” Scientific
Reports, vol. 11, no. 1, article 21200, 2021.
Figure S1: mean value and confidence interval of the shear [7] H. Fossen, B. Tikoff, and C. Teyssier, “Strain modeling of
strain (α) of the WFTZ and EHCFTZ of the CTFB. Error transpressional and transtensional deformation,” Norsk Geo-
bars represent 95% confidence interval. Figure S2: semigeo- logisk Tidsskrift, vol. 74, pp. 134–145, 1994.
logical profile along three sections (AA ′ , BB ′ , and CC ′ ) [8] B. D. Goscombe, D. Gray, R. Armstrong, D. A. Foster, and
based on fieldwork in the CTFB area (concept is based on J. Vogl, “Event geochronology of the Pan-African Kaoko Belt,
the [105]). Note: Black lines with teeth are the thrust faults. Namibia,” Precambrian Research, vol. 140, no. 3-4, pp. 103.
Figure S3: shortening % along section AA ′ . Here, the black e1–103.e41, 2005.
square represents cumulative shortening; the red square rep- [9] M. S. Hossain, M. S. H. Khan, R. Abdullah, and S. Mukherjee,
resents the shortening in kilometer, and the blue star repre- “Late Cenozoic transpression at the plate boundary: kinemat-
sents the shortening percentage for individual anticline. ics of the eastern segment of the Dauki fault zone (Bangladesh)
and tectonic evolution of the petroliferous NE Bengal basin,”
Figure S4: shortening % along section BB ′ . Here, the black
Marine and Petroleum Geology, vol. 131, p. 105133, 2021.
square represents cumulative shortening, the red square rep-
[10] S. Lin, D. Jiang, and P. F. Williams, “Discussion on transpres-
resents the shortening in kilometer, and the blue star repre-
sion and transtension zones,” Journal-Geological Society Lon-
sents the shortening percentage for individual anticline.
don, vol. 135, pp. 1–14, 1999.
Figure S5: (a) approximately vertical thrust plane between
[11] D. J. Sanderson and W. R. D. Marchini, “Transpression,”
sandstone and shale in the eastern flank of the Bandarban
Journal of Structural Geology, vol. 6, no. 5, pp. 449–458, 1984.
structure (Sangu River downstream). (b, c, and d) Evidence
[12] T. K. Mondal, “Evolution of fabric in Chitradurga granite
of fault kinematics in the upstream part of the Sangu River
(South India)-a study based on microstructure, anisotropy
section (East flank of the Bandarban structure). (e and f) of magnetic susceptibility (AMS) and vorticity analysis,” Tec-
Evidence of fault kinematics in the western flank of the Sita- tonophysics, vol. 723, pp. 149–161, 2018.
pahar structure. Note: The red broken line is the fault plane; [13] I. Davison, T. Faull, J. Greenhalgh, E. O. Beirne, and I. Steel,
the half-arrow indicates the hanging-wall movement direc- “Transpressional structures and hydrocarbon potential along
tion; the white dotted line is the striation on the fault plane; the Romanche fracture zone: a review,” in Transform Mar-
and CSB is the calcareous sandstone band. Table S1: stress gins: Development, Controls and Petroleum Systems, M. Nem-
regimes determined from fault-slip analysis (Right Dihe- čok, S. Rybár, S. T. Sinha, S. A. Hermeston, and L.
dron) of all faults. N: number of fault data accepted for stress Ledvényiová, Eds., vol. 431, pp. 235–248, Geological Society,
tensor determination at the lowest misfit angle (software Special Publications, London, UK, 2015.
generated); nt: total number of fault data measured; σ1, σ2, [14] M. Nemčok, S. Rybár, S. T. Sinha, S. A. Hermeston, and
and σ3: orientations (plunge/azimuth) of the maximum, L. Ledvényiová, “Transform margins: development, controls

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 21

and petroleum systems – an introduction,” London: Geologi- [30] B. Shen, T. Siren, and M. Rinne, “Modelling fracture propa-
cal Society, Special Publications, vol. 431, pp. 1–38, 2016. gation in anisotropic rock mass,” Rock Mechanics and Rock
[15] C. Benavente, S. Zerathe, L. Audin et al., “Active transpres- Engineering, vol. 48, no. 3, pp. 1067–1081, 2015.
sional tectonics in the Andean forearc of southern Peru quan- [31] L. Zeng, J. Zhao, S. Zhu, W. Xiong, Y. He, and J. Chen,
tified by10Be surface exposure dating of an active fault scarp,” “Impact of rock anisotropy on fracture development,” Prog-
Tectonics, vol. 36, no. 9, pp. 1662–1678, 2017. ress in Natural Science, vol. 18, no. 11, pp. 1403–1408,
[16] M. Meghraoui and S. Pondrelli, “Active faulting and trans- 2008.
pression tectonics along the plate boundary in North Africa,” [32] F. Zengchao, Z. Yangsheng, and Z. Dong, “Investigating the
Annals of Geophysics, vol. 55, no. 5, pp. 955–967, 2012. scale effects in strength of fractured rock mass,” Chaos, Soli-
[17] P. M. Betka, L. Seeber, S. N. Thomson, M. S. Steckler, tons & Fractals, vol. 41, no. 5, pp. 2377–2386, 2009.
R. Sincavage, and C. Zoramthara, “Slip-partitioning above a [33] D. C. P. Peacock and D. J. Sanderson, “Displacements, seg-
shallow, weak décollement beneath the Indo-Burman accre- ment linkage and relay ramps in normal fault zones,” Journal
tionary prism,” Earth and Planetary Science Letters, of Structural Geology, vol. 13, no. 6, pp. 721–733, 1991.
vol. 503, pp. 17–28, 2018. [34] R. Soliva and A. Benedicto, “A linkage criterion for seg-
[18] T. Maurin and C. Rangin, “Structure and kinematics of the mented normal faults,” Journal of Structural Geology,
Indo-Burmese Wedge: recent and fast growth of the outer vol. 26, no. 12, pp. 2251–2267, 2004.
wedge,” Tectonics, vol. 28, no. 2, article TC2010, 2009. [35] S. Bhowmick and T. Mondal, “Influence of fluid pressure
[19] Y. Wang, K. Sieh, S. T. Tun, K.-Y. Lai, and T. Myint, “Active changes on the reactivation potential of pre-existing frac-
tectonics and earthquake potential of the Myanmar region,” tures: a case study in the Archaean metavolcanics of the
Journal of Geophysical Research: Solid Earth, vol. 119, no. 4, Chitradurga region, India,” Geological Magazine, pp. 1–
pp. 3767–3822, 2014. 15, 2021.
[20] M. S. Hossain, M. M. Rahman, and R. A. Khan, “Active seis- [36] D. Delvaux, F. Kervyn, A. Macheyeki, and E. Temu, “Geo-
mic structures, energy infrastructures, and earthquake disas- dynamic significance of the TRM segment in the East African
ter response strategy - Bangladesh perspective,” International Rift (W-Tanzania): active tectonics and paleostress in the
Energy Journal, vol. 20, no. 3A, pp. 509–522, 2020. Ufipa plateau and Rukwa basin,” Journal of Structural Geol-
[21] A. Boruah, S. Verma, A. Rasheed, G. S. Gairola, and A. Gogoi, ogy, vol. 37, pp. 161–180, 2012.
“Macro-seepage based potential new hydrocarbon prospects [37] N. T. Dina, M. J. J. Rahman, M. S. Hossain, and A. S. M.
in Assam-Arakan Basin, India,” Scientific Report, vol. 12, arti- Sayem, “Provenance of the Neogene succession in the Ban-
cle 2273, 2022. darban structure, South-East Bengal Basin, Bangladesh:
[22] D. Das, T. K. Mondal, and M. S. Hossain, “Quantification of insights from petrography and petrofacies,” Himalayan Geol-
quartz reefs and mafic dykes of Bundelkhand, Craton, Central ogy, vol. 37, no. 2, pp. 141–152, 2016.
India: a study based on spatial and fractal analysis,” Journal of [38] Y. Najman, M. Bickle, M. BouDagher-Fadel et al., “The
the Geological Society of India, vol. 94, no. 3, pp. 227–237, 2019. Paleogene record of Himalayan erosion: Bengal Basin, Ban-
[23] M. S. Hossain and J. H. Kruhl, “Fractal geometry-based quan- gladesh,” Earth and Planetary Science Letters, vol. 273,
tification of shock-induced rock fragmentation in and around no. 1-2, pp. 1–14, 2008.
an impact crater,” Pure and Applied Geophysics, vol. 172, [39] Y. Najman, R. Allen, E. A. F. Willett et al., “The record of
no. 7, pp. 2009–2023, 2015. Himalayan erosion preserved in the sedimentary rocks of
[24] J. H. Kruhl, “Fractal-geometry techniques in the quantifica- the Hatia Trough of the Bengal Basin and the Chittagong Hill
tion of complex rock structures: a special view on scaling tracts, Bangladesh,” Basin Research, vol. 24, no. 5, pp. 499–
regimes, inhomogeneity and anisotropy,” Journal of Struc- 519, 2012.
tural Geology, vol. 46, pp. 2–21, 2013. [40] M. J. J. Rahman, W. Xiao, M. S. Hossain et al., “Geochemistry
[25] R. Pérez-López, C. Paredes, and A. Muñoz-Martín, “Rela- and detrital zircon U-Pb dating of Pliocene-Pleistocene sand-
tionship between the fractal dimension anisotropy of the spa- stones of the Chittagong Tripura Fold Belt (Bangladesh):
tial faults distribution and the paleostress fields on a Variscan implications for provenance,” Gondwana Research, vol. 78,
granitic massif (Central Spain): the F-parameter,” Journal of pp. 278–290, 2020.
Structural Geology, vol. 27, no. 4, pp. 663–677, 2005. [41] L. Yang, W. Xiao, M. J. J. Rahman et al., “Indo-Burma passive
[26] C. O. A. Filho and D. F. Rossetti, “Effectiveness of SRTM and amalgamation along the Kaladan fault: insights from zircon
ALOS-PALSAR data for identifying morphostructural linea- provenance in the Chittagong-Tripura Fold Belt (Bangla-
ments in northeastern Brazil,” International Journal of desh),” Geological Society of America Bulletin, vol. 132,
Remote Sensing, vol. 33, no. 4, pp. 1058–1077, 2012. no. 9-10, pp. 1953–1968, 2020.
[27] T. Ghosh, S. Hazra, and A. K. Das, “Potential of ALOS-2 [42] P. Bürgi, J. Hubbard, S. H. Akhter, and D. E. Peterson,
PALSAR-2 Strip Map data for lithofacies identification and “Geometry of the décollement below eastern Bangladesh
geological lineament mapping in vegetated fold-thrust belt and implications for seismic hazard,” Journal of Geophysical
of Nagaland, India,” Advances in Space Research, vol. 69, Research: Solid Earth, vol. 126, 2021.
no. 4, pp. 1840–1862, 2022. [43] P. R. Cummins, “The potential for giant tsunamigenic earth-
[28] F. A. Donath, “Experimental study of shear failure in aniso- quakes in the northern bay of Bengal,” Nature, vol. 449,
tropic rocks,” Geological Society of America Bulletin, vol. 72, no. 7158, pp. 75–78, 2007.
no. 6, pp. 985–990, 1961. [44] M. S. Hossain, W. Xiao, M. S. H. Khan, K. R. Chowdhury, and
[29] A. A. Griffith, “The phenomena of rupture and flow in S. Ao, “Geodynamic model and tectono-structural frame-
solids,” Philosophical Transactions of the Royal Society, Lon- work of the Bengal Basin and its surroundings,” Journal of
don, A., vol. 221, pp. 163–198, 1921. Maps, vol. 16, no. 2, pp. 445–458, 2020.

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
22 Lithosphere

[45] A. Das, S. Bose, S. Dasgupta, S. Roy, and B. Mukhopadhyay, [61] N. Woodward, S. E. Boyer, and J. Suppe, Balanced Geological
“Post-Oligocene evolution of Indo-Burma wedge: insights Cross-Sections: An Essential Technique in Geological Research
from deformation structures of Tripura Mizoram fold belt,” and Exploration, American Geophysical Union, Washington,
Journal of Structural Geology, vol. 154, article 104497, 2022. DC, USA, 1989.
[46] M. S. H. Khan, M. S. Hossain, and K. R. Chowdhury, “Geo- [62] J. Angelier, “Fault slip analysis and palaeostress construc-
morphic implications and active tectonics of the Sitapahar tion,” in Continental Deformation, P. L. Hancock, Ed.,
anticline–CTFB, Bangladesh,” Bangladesh Geoscience Jour- pp. 53–100, Pergamon Press, London, UK, 1994.
nal, vol. 23, pp. 1–24, 2017. [63] J. Angelier, “Inversion of field data in fault tectonics to obtain
[47] M. K. Alam, A. K. M. S. Hasan, M. R. Khan, and J. W. the regional stress – III. A new rapid direct inversion method
Whitney, “Geological map of Bangladesh,” in Scale by analytical means,” Geophysical Journal International,
1:1,000,000, Geological Survey of Bangladesh, 1990. vol. 103, pp. 363–376, 1990.
[48] M. Alam, M. M. Alam, J. R. Curray, M. L. R. Chowdhury, and [64] J. M. Dupin, W. Sassi, and J. Angelier, “Homogeneous stress
M. R. Gani, “An overview of the sedimentary geology of the hypothesis and actual fault slip: a distinct element analysis,”
Bengal Basin in relation to the regional tectonic framework Journal of Structural Geology, vol. 15, no. 8, pp. 1033–1043,
and basin-fill history,” Sedimentary Geology, vol. 155, no. 3- 1993.
4, pp. 179–208, 2003. [65] A. Etchecopar, G. Vasseus, and M. Daigniers, “An inverse
[49] Y. Najman, L. Bracciali, R. R. Parrish, E. Chisty, and problem in microtectonics for the determination of stress
A. Copley, “Evolving strain partitioning in the Eastern Hima- tensor from fault striation analysis,” Journal of Structural
laya: the growth of the Shillong plateau,” Earth and Planetary Geology, vol. 3, no. 1, pp. 51–65, 1981.
Science Letters, vol. 433, pp. 1–9, 2016. [66] N. Fry, “Striated faults: visual appreciation of their constraint
[50] N. Cardozo and R. W. Allmendinger, “Spherical projections on possible paleostress tensors,” Journal of Structural Geol-
with OSXStereonet,” Computers & Geosciences, vol. 51, ogy, vol. 21, no. 1, pp. 7–21, 1999.
pp. 193–205, 2013. [67] D. Gapais, P. R. Cobbold, O. Bourgeois, D. Rouby, and M. de
[51] A. Gerik and J. H. Kruhl, “Towards automated pattern quan- Urreiztieta, “Tectonic significance of fault–slip data,” Journal
tification: time efficient assessment of anisotropy of 2D pat- of Structural Geology, vol. 22, no. 7, pp. 881–888, 2000.
terns with AMOCADO,” Computers & Geosciences, vol. 35, [68] J.-A. Hansen, “Direct inversion of stress, strain or strain rate
no. 6, pp. 1087–1097, 2009. including vorticity: a linear method of homogenous fault-slip
[52] S. Volland and J. H. Kruhl, “Anisotropy quantification: the data inversion independent of adopted hypothesis,” Journal
application of fractal geometry methods on tectonic fracture of Structural Geology, vol. 51, pp. 3–13, 2013.
patterns of a Hercynian fault zone in NW-Sardinia,” Journal [69] R. J. Lisle, “Principal stress orientations from faults: an addi-
of Structural Geology, vol. 26, no. 8, pp. 1499–1510, 2004. tional constraint,” Annales Tectonicae-International Journal
[53] D. Healy, R. E. Rizzo, D. G. Cornwell et al., “FracPaQ: A of Structural Geology and Tectonics, vol. 1, pp. 155–158, 1987.
MATLAB™ toolbox for the quantification of fracture pat- [70] R. Marrett and R. W. Allmendinger, “Kinematic analysis of
terns,” Journal of Structural Geology, vol. 95, pp. 1–16, fault–slip data,” Journal of Structural Geology, vol. 12, no. 8,
2017. pp. 973–986, 1990.
[54] N. Kianizadeh, B. Rahimi, G. R. Lashkaripour, and V. Fakoor, [71] T. K. Mondal and S. S. Acharyya, “Fractured micro-granitoid
“Kinematic and statistical analysis of fold related fractures in enclaves: a stress marker,” Journal of Structural Geology,
the northeast of Kopet Dagh Folded Belt, Northeast Iran,” vol. 113, pp. 33–41, 2018.
Journal Geological Society of India, vol. 97, no. 4, pp. 428– [72] T. K. Mondal, S. Bhowmick, S. Das, and A. Patsa, “Paleostress
441, 2021. field reconstruction in the western Dharwar Craton, South
[55] K. Mobasher and H. A. Babaie, “Kinematic significance of India: evidences from brittle faults and associated structures
fold- and fault related fracture systems in the Zagros moun- of younger granites,” Journal of Structural Geology, vol. 135,
tains, southern Iran,” Tectonophysics, vol. 451, no. 1-4, article 104040, 2020.
pp. 156–169, 2008. [73] T. K. Mondal and M. A. Mamtani, “Palaeostress analysis of
[56] B. J. Stephenson, A. Koopman, H. Hillgartner et al., “Struc- normal faults in granite implications for interpreting Riedel
tural and stratigraphic controls on fold-related fracturing in shearing related to regional deformation,” Journal of the Geo-
the Zagros Mountains, Iran: implications for reservoir devel- logical Society, vol. 173, no. 1, pp. 216–227, 2016.
opment,” Geological Society London Special Publication, [74] J. G. Ramsay and R. J. Lisle, “The Techniques of Modern
vol. 270, no. 1, pp. 1.2–121, 2007. Structural Geology,” in Applications of Continuum Mechanics
[57] D. D. Pollard and A. Aydin, “Progress in understanding in Structural Geology, pp. 1–1061, Academic Press, London,
jointing over the past century,” Geological Society of America UK, 2000.
Bulletin, vol. 100, no. 8, pp. 1181–1204, 1988. [75] P. Thakur, D. C. Srivastava, and P. K. Gupta, “The genetic
[58] D. W. Stearns and M. Friedman, “Reservoirs in fractured algorithm: a robust method for stress inversion,” Journal of
rock: geological exploration methods,” AAP G Special Vol- Structural Geology, vol. 94, pp. 227–239, 2017.
ume, vol. 16, pp. 82–100, 1972. [76] M. D. Tranos, “TR method (TRM): a separation and stress
[59] T. Afrin, D. Hossain, and M. H. Imam, “Seismic characteriza- inversion method for heterogeneous fault-slip data driven
tion of the Semutang gas field, Bengal Basin, Bangladesh,” by Andersonian extensional and compressional stress
Journal Geological Society of India, vol. 86, no. 5, pp. 547– regimes,” Journal of Structural Geology, vol. 79, pp. 57–74,
552, 2015. 2015.
[60] C. D. A. Dahlstorm, “Balanced cross sections,” Canadian [77] R. J. Twiss and J. R. Unruh, “Analysis of fault slip inversions:
Journal of Earth Sciences, vol. 6, no. 4, pp. 743–757, 1969. do they constrain stress or strain rate?,” Journal of

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao
Lithosphere 23

Geophysical Research: Solid Earth, vol. 103, no. B6, of the Bengal Basin, Bangladesh,” Sedimentary Geology,
pp. 12205–12222, 1998. vol. 155, no. 3-4, pp. 209–226, 2003.
[78] A. Yamaji, “The multiple inverse methods: a new technique [93] J. G. Ramsay, “Rock ductility and its influence on the devel-
to separate stresses from heterogeneous fault-slip data,” Jour- opment of tectonic structures in mountain belts,” in Moun-
nal of Structural Geology, vol. 22, no. 4, pp. 441–452, 2000. tain Building Processes, K. J. Hsu, Ed., pp. 111–127,
[79] J. Žalohar and M. Vrabec, “Paleostress analysis of heteroge- Academic Press, London, UK, 1982.
neous fault–slip data: the Gauss method,” Journal of Struc- [94] F. A. Donath and R. B. Parker, “Folds and folding,” Geological
tural Geology, vol. 29, no. 11, pp. 1798–1810, 2007. Society of America Bulletin, vol. 75, no. 1, pp. 45–62, 1964.
[80] G. T. Blenkinsop, “Kinematic and dynamic fault slip analyses: [95] R. J. Twiss and E. M. Moores, “Structural Geology,” in Glaci-
implications from the surface rupture of the 1999 Chi-Chi, ology, p. 531, W.H. Freeman and Company, New York, NY,
Taiwan, earthquake,” Journal of Structural Geology, vol. 28, USA, 1992.
no. 6, pp. 1040–1050, 2006. [96] K. McClay, The mapping of geological structures, London
[81] J. W. Gephart and D. W. Forsyth, “An improved method for Handbook, Open University Press, Milton Keynes, 1987.
determining the regional stress tensor using earthquake focal [97] N. Price, Fault and joint development in brittle and semi-
mechanism data: application to the San Fernando earthquake brittle rocks, Pergamon Press, Oxford, UK, 1966.
sequence,” Journal Geophysical Research, vol. 89, no. B11,
pp. 9305–9320, 1984. [98] H. A. Ardakani, F. Ghaemi, and B. Rahimi, “Geometric and
kinematic analysis of Dorbadam anticline, north of Quchan,
[82] D. Delvaux, “Win-tensor, an interactive computer program
Iran,” Iranian Journal of Earth Sciences, vol. 10, pp. 121–
for fracture analysis and crustal stress reconstruction,” in
134, 2018.
EGU General Assembly, vol. 13, Geophysical Research
Abstract, Vienna, Austria, 2011. [99] O. Heidbach, M. Rajabi, K. Reiter, and M. Ziegler, World
Stress Map 2016, vol. 1.1, GFZ Data Services, 2016.
[83] D. Delvaux and B. Sperner, “Stress tensor inversion from
fault kinematic indicators and focal mechanism data: the [100] C. Nielsen, N. Chamot-Rooke, and C. Rangin, “From partial
TENSOR program,” in New Insights into Structural Interpre- to full strain partitioning along the Indo-Burmese hyper-
tation and Modelling, D. Nieuwland, Ed., vol. 212, pp. 75– oblique subduction,” Marine Geology, vol. 209, no. 1-4,
100, Geological Society, Special Publication, London, UK, pp. 303–327, 2004.
2003. [101] K. E. Bradley, L. Feng, E. M. Hill, D. H. Natawidjaja, and
[84] D. J. Sanderson, “The transition from upright to recumbent K. Sieh, “Implications of the diffuse deformation of the
folding in the Variscan fold belt of Southwest England: a Indian Ocean lithosphere for slip partitioning of oblique plate
model based on the kinematics of simple shear,” Journal of convergence in Sumatra,” Journal of Geophysical Research:
Structural Geology, vol. 1, no. 3, pp. 171–180, 1979. Solid Earth, vol. 122, no. 1, pp. 572–591, 2017.
[85] H.-T. Chu, J. -C. Lee, F. Bergerat et al., “Fracture patterns and [102] R. McCaffrey, P. C. Zwick, Y. Bock et al., “Strain partitioning
their relations to mountain building in a fold-thrust belt: a during oblique plate convergence in northern Sumatra: geo-
case study in NW Taiwan,” Bulletin de la Société Géologique detic and seismologic constraints and numerical modeling,”
de France, vol. 184, no. 4-5, pp. 485–500, 2013. Journal of Geophysical Research: Solid Earth, vol. 105,
[86] O. Lacombe, N. Bellahsen, and F. Mouthereau, “Fracture pat- no. B12, pp. 28363–28376, 2000.
terns in the Zagros simply Folded Belt (Fars, Iran): con- [103] Z. Huang, L. Wang, M. Xu, D. Zhao, N. Mi, and D. Yu, “P and
straints on early collisional tectonic history and role of S wave tomography beneath the SE Tibetan Plateau: Evidence
basement faults,” Geological Magazine, vol. 148, no. 5-6, for lithospheric delamination,” Journal of Geophysical
pp. 940–963, 2011. Research: Solid Earth, vol. 124, no. 10, pp. 292–308, 2019.
[87] D. Delvaux, R. Moeys, G. Stapel et al., “Paleostress recon- [104] A. Kumar, S. Mitra, and G. Suresh, “Seismotectonics of the
structions and geodynamics of the Baikal region, Central Eastern Himalayan and indo-burman plate boundary sys-
Asia, part 2. Cenozoic rifting,” Cenozoic rifting. Tectonophy- tems,” Tectonics, vol. 34, no. 11, pp. 2279–2295, 2015.
sics, vol. 282, no. 1-4, pp. 1–38, 1997. [105] B. Wang, L. Wu, W. Li et al., “A semi-automatic approach for
[88] J. D. Das, A. K. Saraf, and Y. Shujat, “A remote sensing tech- generating geological profiles by integrating multi-source
nique for identifying geometry and geomorphological fea- data,” Ore Geology Reviews, vol. 134, article 104190, 2021.
tures of the Indo-Burman frontal fold belt,” International
Journal of Remote Sensing, vol. 31, no. 16, pp. 4481–4503,
2010.
[89] D. Davis, J. Suppe, and F. A. Dahlen, “Mechanics of fold-and
thrust belts and accretionary wedges,” Journal of Geophysical
Research: Solid Earth, vol. 88, no. B2, pp. 1153–1172, 1983.
[90] W. J. Jamison, “Geometric analysis of fold development in
overthrust terranes,” Journal of Structural Geology, vol. 9,
no. 2, pp. 207–219, 1987.
[91] S. Mitra, “Effects of deformation mechanism on reservoir
potential in the central Appalachian,” American Associa-
tion of Petroleum Geologists Bulletin, vol. 72, pp. 536–
554, 1988.
[92] A. Sikder and M. M. Alam, “2-D modelling of the anticlinal
structures and structural development of the eastern fold belt

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/doi/10.2113/2022/6058346/5587560/6058346.pdf


by The Institute of Geology and Geophysics, Chinese Academy of Sciences, Prof. Songjian Ao

You might also like