Tectonics and Structural Geology Indian Context 2019
Tectonics and Structural Geology Indian Context 2019
Tectonics and Structural Geology Indian Context 2019
Tectonics and
Structural
Geology: Indian
Context
Springer Geology
The book series Springer Geology comprises a broad portfolio of scientific books,
aiming at researchers, students, and everyone interested in geology. The series
includes peer-reviewed monographs, edited volumes, textbooks, and conference
proceedings. It covers the entire research area of geology including, but not limited
to, economic geology, mineral resources, historical geology, quantitative geology,
structural geology, geomorphology, paleontology, and sedimentology.
123
Editor
Soumyajit Mukherjee
Department of Earth Sciences
Indian Institute of Technology Bombay
Powai, Mumbai, Maharashtra, India
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
...geology contents in geological text books
for compulsory education is not regularly
updated, so new paradigms are included
belatedly (as happened, e.g., with plate
tectonics), and this is one of the reasons why
younger students lag behind in Geosciences.
—Brusi et al. (2016)
Annett Buettener and Helen Ranchner (Springer) are thanked for handling this
book proposal positively. The Springer proofreading team is acknowledged for
support. I thank the authors and the reviewers for their participation. Thesis students,
interns and visitors in the Geodynamics lab during 2017–2018: Narayan Bose,
Dripta Dutta and Tarunkanti Das (IIT Bombay), Prof. Seema Singh and Ajay
Kumar (Panjab University), Swagato Dasgupta (Haliburton), Troyee Dasgupta
(Reliance Industries Limited), Chandan Majumdar (Schlumberger), Tuhin Biswas
(ONGC), Rajkumar Ghosh (Geological Survey of India), Chanel Vidal (Iowa State
University), Saber Idriss (University of SFax), Puja Banerjee (Institut De Physique
Du Globe De Paris), Ishiqua Agarwal (IIT Kharagpur), Naimisha Vanik
and Haroon Saikh (MS University Baroda), Shiba Nikalje (St. Xavier’s College,
Mumbai), Amey Dashputre and Renuka Kale (Fergusson College), Rucha
Kanchan and Samidha Shinde (Pune University), Lokesh Tayade (IISER Pune),
Rohit Shaw, Madhurima Bose, Anuva Chowdhury and Jayesh Mukherjee
(Presidency University, Kolkata) helped in various ways. A research sabbatical
provided by IIT Bombay to me for the year 2017 helped much to edit this book.
vii
Contents
ix
x Contents
Soumyajit Mukherjee
S. Mukherjee (&)
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai
400076, Maharashtra, India
e-mail: smukherjee@iitb.ac.in; soumyajitm@gmail.com
In their very detailed review on the Bengal basin, Hossain et al. (2019) in
Chap. 6 present the basic division of this basin, fault distribution, and how these
divisions evolved temporally with or without volcanism.
Goswami and Upadhyay (2019) in Chap. 7 study the structural geology and
geochemistry of the Kadiri schist belt (Cuddapah) and decipher an ocean-continent
subduction tectonics and a volcanic arc setting of the terrain.
Detailed field investigation of the structural geology of the Nallamalai Fold Belt
(Cuddapah) by Tripathy et al. (2019) in Chap. 8 reveals a Pan-African thin-skinned
tectonics, which link with the tectonics of the East Gondwana fragments.
Multi disciplinary geoscientific studies by Mazumder et al. (2019) in Chap. 9
reveal that a number of E trending steeply dipping shear zones pass through the
northern part of the Cauvery Basin that was later reactivated.
Dasgupta (2019) in Chap. 10 reviews the Cauvery basin’s tectonics. Half gra-
bens in its all the three sub basins signify a rift origin of the basin. This article
analyzes the transfer zone geometries from the Cauvery basin that are crucial in
developing hydrocarbon trap conditions.
Misra et al. (2019) in Chap. 11 study the field structural geology of the Ramgarh
impact structure (SE Rajasthan), and especially its fracture patterns. They conclude
that impacting happened at the palaeo-channel of the river Parvati.
Dinkar et al. (2019) in Chap. 12 describe in detail field structural geology from
the Lalitpur district (Uttar Pradesh). The notable information are E/ENE trending
axial traces and Proterozoic to Neoproterozoic reactivation plausible in the southern
part of the study area.
Singh and Awasthi (2019) in Chap. 13 discuss the tectonics of the Kangra region
(Himachal Pradesh), which is presumably devoid of any weak layer below itself.
Overpressure condition at depth possibly due to fluid activity had helped to
propagate this crustal wedge towards the foreland side.
Kumar et al. (2019a) in Chap. 14 describe from the field along with attractive
photographs the damage zone associated with the Munsiari Thrust, a strand of the
Main Central Thrust, from the Mandakini river section, Higher Himalaya. The
authors document more landslides from the damage zone and perform engineering
geological studies from such zones.
Mahato et al. (2019) in Chap. 15 perform detailed field studies from the
Mussoorie syncline and the nearby regions from the Uttarakhand Lesser Himalaya.
Top-to-N/NE back shear and Himalayan arc-parallel shears (such as top-to-NW) are
the new meso scale findings in this work.
Banerjee et al. (2019) in Chap. 16 too document orogen-parallel shear from the
Darjeeling Group of rocks from the Sikkim Lesser Himalaya. A more detail work
from the same research group has been submitted in a journal where such defor-
mation is reported from the Siwalik Himalaya (Dutta et al. submitted).
Kumar et al. (2019b) in Chap. 17 discuss the database of lead (Pb) content in the
Indian Gondwana coal (207Pb/206 Pb = 0.7150–0.8845; 208Pb/206 Pb = 1.9484–
2.2231; Pb concentration = 3.2–566 mg kg−1). This study will have a far-reaching
implication in India-Antarctica plate reconstruction.
Introduction 3
Readers without any instructors, especially students (in some unfortunate cases),
are requested to go through few recent books on structural geological and tectonic
principles and Indian case studies (e.g., Sharma 2010; Mukherjee 2013a, b, 2014,
2015a, b; Mukherjee et al. 2017; Mukherjee and Mulchrone 2015; Mukherjee et al.
2015, 2017; Valdiya 2016; Bose and Mukherjee 2017; Dasgupta and Mukherjee
2017; Chetty 2018; Misra and Mukherjee 2018; Roy and Purohit 2018; Acharyya,
in press) before going through this book.
Refer this book as follows:
• Mukherjee S (2019) Tectonics and Structural Geology: Indian Context.
Springer International Publishing AG, Cham. ISBN 978-3-319-99340-9.
pp. 1–455.
Refer individual chapters of this book as follows:
• Banerjee S, Bose N, Mukherjee S (2019) Field structural geological studies
around Kurseong, Darjeeling-Sikkim Himalaya, India. In: Mukherjee S
(ed) Tectonics and Structural Geology: Indian context. Springer International
Publishing AG, Cham. ISBN 978-3-319-99340-9. pp. 425–440.
References
Acharyya SK (in press) Tectonic setting and Gondwana Basin architecture in the Indian Shield. In:
Mukherjee S (eds) Developments in structural geology and tectonics. Elsevier.
ISBN: 978-0-12-815218-8
Babar MD, Kaplay RD, Mukherjee S, Mahato S, Gurav C (2019) NE-SW strike-slip fault in the
granitoid from the margin of the South East Dharwar Craton, Degloor, Nanded district,
Maharashtra, India. In: Mukherjee S (ed) Tectonics and structural geology: Indian context.
Springer International Publishing AG, Cham, pp 115–134. ISBN 978-3-319-99340-9
Banerjee S, Bose N, Mukherjee S (2019) Field structural geological studies around Kurseong,
Darjeeling-Sikkim Himalaya, India. In: Mukherjee S (ed) Tectonics and structural geology:
Indian context. Springer International Publishing AG, Cham, pp 425–440. ISBN
978-3-319-99340-9
Bose N, Mukherjee S (2017) Map interpretation for structural geologists. In: Mukherjee S
(ed) Developments in structural geology and tectonics. Elsevier, Amsterdam. ISBN:
978-0-12-809681-9. ISSN: 2542-9000
Chetty TRK (2018) Proterozic orogens of India: a critical Window to Gondwana. Elsevier,
Amsterdam. ISBN: 9780128044414
Dasgupta S (2019) Implication of transfer zones in rift fault propagation: example from Cauvery
basin, Indian east coast. In: Mukherjee S (ed) Tectonics and structural geology: Indian
contexts. Springer International Publishing AG, Cham, pp 313–326. ISBN 978-3-319-99340-9
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology 125, 561–591
Dinkar GK, Bhattacharya AR, Verma AK, Sharma P (2019) Geology, structural architecture and
tectonic framework of the rocks of Southern Lalitpur District, Uttar Pradesh, India: an epitome
of the Indian Peninsular Shield. In: Mukherjee S (ed) Tectonics and structural geology: Indian
context. Springer International Publishing AG, Cham, pp 353–379. ISBN 978-3-319-99340-9
4 S. Mukherjee
1 Introduction
We presently believe that the continental crust grew episodically and that smaller
continents were united and disintegrated several times in the past *4000 Ma
(Hoffmann 1989; Rogers 1996). This process of plate jostling and its subsequent
destruction, commonly termed as ‘supercontinental cycle’, eventually controls the
harmonic interactions among lithosphere, hydrosphere and atmosphere over geo-
logical time (Worsley et al. 1985, 1986; Piper 2013). The antiquity and the
petrological diversity of the rocks in the Indian shield offer unique opportunity to
configure the supercontinents that existed in the geological past (Acharyya 2003;
Meert et al. 2010). One of the outstanding and hotly debated problems is the
position of the Indian shield in the proposed Precambrian supercontinents in general
and timing of suturing of India and east Antarctica in particular (Torsvik et al. 2001;
Dasgupta and Sengupta 2003; Pisarevsky et al. 2003; Bhowmik et al. 2012). With
regard to the latter, two competing views exist. One view is India and Antarctica
were united at least from *1000 Ma (Hoffmann 1989; Dalziel 1991; Li et al.
2008). The disclaimers of this view propose that the two continents got juxtaposed
not earlier than *900 Ma (Bhowmik et al. 2012) if not after *750 Ma (Merdith
et al. 2017; Torsvik et al. 2001). In the reconstructed Rodinia, a Proterozoic
supercontinent, the CGGC and the Eastern Ghats Mobile Belt juxtaposed against
the east Antarctic Precambrian basement (Dasgupta and Sengupta 2003; Chatterjee
et al. 2010; Mukherjee et al. 2017a). Therefore, rocks of these two areas of the
Indian shield are likely to provide the clinching evidence about the timing of
Indo-Antarctic suturing.
Till the end of twentieth century, limited petrological information and rudi-
mentary geochronological data were available from the CGGC that is positioned
The CGGC is an east-west trending mobile belt that belongs to the east Indian
Shield and is exposed across the states of Jharkhand, Bihar, West Bengal and
Chhattisgarh covering an area of over 100,000 km2 (reviewed in Mahadevan 2002;
Acharyya 2003). The northern margin of the CGGC is covered by quaternary
sediments of Gangetic alluvium (Fig. 2a). Sediments of the Bengal Basin mark the
eastern boundary of the terrain and Mesozoic volcanics of Rajmahal Trap covers
the northeastern fringe of the terrain. The western margin of CGGC is dominantly
covered by Gondwana deposits of Permian to mid-Cretaceous age (Mahadevan
2002). However, in the northwestern part Vindhyan sediments and the Mahakoshal
group of rocks are in contact with the Proterozoic rocks of the CGGC. Towards the
south the contact between CGGC and Proterozoic rocks of the North Singhbhum
Fold Belt (NSFB) is marked by the east-west trending crustal scale shear zone
called the South Purulia Shear Zone (SPSZ) or the Tamar-Porapahar-Khatra Shear
zone.
Fig. 1 Proposed configuration of India and Antarctica during the assembly of Rodinia; modified
after Dasgupta and Sengupta (2003). The schematic map of India and it does not represent the
erstwhile configuration as northeastern and northwestern margin (Tethyan sequence) formed later
10 S. Mukherjee et al.
Proterozoic Crustal Evolution of the Chotanagpur Granite … 11
JFig. 2 a Geological map of the Chotanagpur Granite Gneiss Complex (CGGC) and the showing
major domains (modified after Acharyya 2003). EITZ: Eastern Indian Tectonic Zone, NPSZ:
North Purulia Shear Zone, SPSZ: South Purulia Shear Zone, RT: Rajmahal Trap. b Geological
map of the Chotanagpur Granite Gneiss Complex (CGGC) showing five subdivisions proposed by
Mahadevan (2002) and Sanyal and Sengupta (2012), modified after Sanyal and Sengupta (2012)
3.1 Domain IA
This domain covers rocks exposed in the southernmost part of the CGGC and is bounded
by GBF and the SPSZ (Fig. 2a). The geological information albeit sparse, clusters
around the Bankura-Saltora-Bero area in the east, the Raghunathpur-Adra-Ranchi areas
12 S. Mukherjee et al.
in the central part and Raikera-Kunkuri region in the western part of the domain (Fig. 2a).
Petrology and geochronology of the rocks exposed around Bankura-Saltora-Bero—have
been extensively studied by several workers (Manna and Sen 1974; Roy 1977;
Bhattacharyya and Mukherjee 1987; Sen and Bhattacharya 1993; Mukherjee et al. 2005;
Chatterjee et al. 2008; Maji et al. 2008), which reveals that variably deformed migmatitic
felsic orthogneisses, holding dismembered rafts of mafic granulite and calc-silicate
gneisses is the dominant rocktype intruded by massif type anorthositic rocks (Bengal
anorthosite). Metamorphic grade varies from amphibolite to granulite grade conditions.
Documentation of granulite and amphibolite facies rocks mostly coming from the eastern
and the western part of the domain respectively (Karmaker et al. 2011; Goswami and
Bhattacharya 2010, 2013; Maji et al. 2008; Chatterjee et al. 2008; Sanyal and Sengupta
2012). However the published information does not support any systemic variation in the
estimated metamorphic conditions along any geographic direction. The rocks in this
region are folded to east-west closing folds and an E-W trending axial planar fabric
dipping steeply towards north (Maji et al. 2008).
In the easternmost parts, near Bero-Saltora-Santuri, high grade metapelitic rocks
and migmatitic quartzofeldspathic rocks are exposed. Three deformation (D1-3)
phases accompanied by four metamorphic events (M1-4) have been inferred from
the area by Maji et al. (2008). Earliest tectonothermal event (M1), considered to be
of granulite grade (minimum P–T estimates of 5–6 kbar and 750–850 °C), pro-
duced the migmatitic banding of the orthogneisses (S1) (Sen and Bhattacharya
1993; Maji et al. 2008) and occurred at ca. 1.70 Ga, inferred from chemical dating
of monazite (Chatterjee et al. 2010). Both the subsequent deformations (D2-3)
occurred under amphibolite facies (M2-3) between 1.3 and 1.1 Ga and replaces the
older granulite mineralogy to variable extant (Maji et al. 2008). Towards the eastern
fringe of the domain, near Saltora (West Bengal) (Fig. 2a) intrusion of massif
anorthosite, called the Bengal anorthosite, occurred at ca. 1.55 Ga between the D1
and D2 (Bhattacharyya and Mukherjee 1987; Chatterjee et al. 2008; Maji et al.
2008). The most pervasive metamorphic event is inferred to have occurred between
1.0 and 0.95 Ga, that has been designated as M4 by Maji et al. (2008). However the
P–T conditions of the event is debated. Maji et al.(2008) inferred that the meta-
moprhism culminated at 650 ± 50 °C at 4–5 kbar whereas Chatterjee et al. (2008)
recovered a high grade conditions (850–900 °C and 8.5–11 kbar) from gabbro
anorthositic rocks. Near Kankarkiari (West Bengal), migmatitic felsic orthogneisses
has been intruded by nepheline-bearing syenite and subsequently got deformed and
metamorphosed (Das et al. 2016; Goswami and Bhattacharyya 2010). No
geochronological data is available to constrain this tectonothermal event experi-
enced by the syenitic rocks. However, field observations suggest that they have
intruded after the Grenvillian metamorphic event and subsequently got deformed
and metamorphosed under amphibolite grade yielding a P–T condition of
700–750 °C and *10 kbar, associated with development of foliation and folding
(Das et al. 2016). Younger Neoproterozoic ages (ca. 900–820 Ma) recovered from
the overgrowths on older monazite grains have been documented by several
Proterozoic Crustal Evolution of the Chotanagpur Granite … 13
workers (Maji et al. 2008; Chatterjee et al. 2010), which further attest for the Late
Tonian–Early Cryogenian tectonothermal event. One of the source of heating could
be deformation itself (Mukherjee and Mulchrone 2013; Mulchrone and Mukherjee
2015, 2016; Mukherjee 2017).
The area around Raghunathpur (West Bengal), west of Bero, reveals an
ensemble of different generations of felsic orthogneisses (porphyritic and deformed
granitoids), metapelitic and calcareous enclaves (Dunn 1929; Baidya et al. 1987,
1989; Ray Barman and Bishui 1994; Goswami and Bhattacharyya 2010, 2013;
Karmakar et al. 2011). Three major deformational events (D1-D3) have been
identified from the area associated with two major metamorphic phases (M1-M2).
Non-porphyritic granite intrusion occurred at 1178 ± 61 Ma (Rb–Sr whole rock
isochron, Ray Barman and Bishui 1994) and is synchronous with development of
S1 during D1 (Goswami and Bhattacharyya 2010). A porphyritic charnockite that
was emplaced prior to D2 shows a Rb–Sr whole rock age of 1071 ± 64 Ma (Ray
Barman and Bishui 1994; Goswami and Bhattacharya 2010). These porphyritic
granites have been geochemically classified as Shoshonitic to high K-calc alkalic
intrusives that were formed via mixing of mantle-derived mafic magma and crustal
melts followed by fractional crystallization in continent-continent collisional set-
tings (Goswami and Bhattacharyya 2013). Towards the southeast of Raghunathpur,
near Adra (West Bengal), migmatitic quartzofeldspathic gneiss contains enclave of
Mg–Al granulite and mafic granulites. Chemical ages of monazite from these Mg–
Al pelite and migmatitic gneiss indicate that the most pervasive and prominent
tectonothermal event of the area occurred between ca. 990–940 Ma that culminated
at *870 °C and 11 kbar pressure followed by a steeply decompressive path
(Karmakar et al. 2011). Goswami and Bhattacharyya (2010) determined a similar
temperature (*800 °C) but lower pressure (6.5–7.5 kbar) and argued that M1
spanned over during both D1 and D2. Youngest monazite age population of ca.
850–775 Ma recovered by Karmakar et al. (2011) are consistent with the
Neoproterozoic dates (870 ± 40 Ma: K–Ar biotite of porphyritic granite and
810 ± 40 Ma: K–Ar muscovite of leucogranite) reported by Baidya et al. (1987)
from the western part of the area near Jaipur, West Bengal. Although petrological
manifestations are lacking, the youngest age clusters presumably constrains the
third tectonothermal event (D3).
Litho-package exposed in the south-central of the terrain, south of Ranchi,
resembles rocktypes of the eastern part, containing garnetiferous migmatitic felsic
gneiss, pelitic schist and minor calc-silicate bodies that have been intruded by
porphyritic granite (Sarkar and Jha 1985; Rekha et al. 2011). Zircon and monazite
geochronological studies of both the older migmatitic gneiss and metapelite reveals
Neoproterozoic ages (944 ± 9 and 921 ± 18 Ma), inferred to be metamorphic,
whereas younger granites yield an emplacement age of 928 ± 23 Ma with older
inherited components of 1072 ± 17 and 1239 ± 66 Ma (Rekha et al. 2011).
At the south-western part of the terrain, near Raikera–Kunkuri region
(Chhattisgarh), different generation of granite bodies are associated with pelitic
schist (with chlorite, biotite and hornblende schist), quartzite and dolerite
14 S. Mukherjee et al.
dykes/sills (Singh and Krishna 2009). Two-mica bearing grey granites, derived
from juvenile crustal sources (SrI 0.7047 and low high field strength elements),
intruded the crust at 1005 ± 51 Ma (Singh and Krishna 2009). On the other hand,
younger Rb–Sr isochron age of 815 ± 47 Ma and high SrI (0.7539) determined
from the pink granite are inferred to be the late metasomatic event associated with
the Y-mineralization in the area (Singh and Krishna 2009).
3.2 Domain IB
during D1. The foliation in the host gneiss represents the dominant foliation of the
area and strikes E-W. They refolded during subsequent D2 and D3 deformation
(Sanyal and Sengupta 2012). Protolith of porphyritic charnockite emplaced in
between D1 and D2 at 1515 ± 5 Ma (Acharyya 2003). A swarm of mafic dykes
(plagioclase + amphibole + clinopyroxene + chlorite + epidote + calcite + quartz
+ ilmenite) cut across the foliation of host gneiss but is folded by open D3 folds. The
Mg–Al granulitic enclave rocks develop a mineral assemblage of aluminous (7 wt%
Al2O3) orthopyroxene + magnetite-hercynite + sillimanite +quartz + garnet + melt
which is a good indicator of ultrahigh-temperature metamorphism (>900 °C) at a
pressure in excess of *8 kbar (Sanyal and Sengupta 2012). From conventional
geothermobarometry of the mafic enclaves, Sanyal and Sengupta (2012) has obtained
distinctly lower temperatures 825–850 °C at 8–9 kbar indicating subsequent cooling
followed by the UHT metamorphism. Geothermobarometry of the porphyritic
charnockite constrain the conditions of M3 metamorphism synchronous to D3, at
700 ± 50 °C and 6.5 ± 1 kbar (Sanyal and Sengupta 2012).
Maximum petrological and geochronological information is available from
further north-west, in between Dumka and Deoghar town (Jharkhand); especially
from northern part of Dumka. The country rock of felsic orthogneiss hosts km to cm
scale rafts of pelitic rocks (garnet-sillimanite-biotite-K-feldspar-plagioclase-
quartz ± spinel), mafic rocks (plagioclase + clinopyroxene + orthopyrox-
ene + garnet + hornblende + ilmenite + rutile), calc silicates (clinopyroxene +
plagioclase + titanite ± garnet ± amphibole ± scapolite ± calcite), granulites
and augen gneisses (K-feldspar + plagioclase + quartz + biotite + hornblende
+ apatite). Mineralogically the host gneiss varies from charnockitic
(orthopyroxene + clinopyroxene + garnet + plagioclase + K-feldspar + quartz +
hornblende + biotite + ilmenite) to biotite-hornblende gneiss to hornblende-biotite
gneiss (garnet + plagioclase + K-feldspar + quartz +hornblende + biotite + il-
menite) (Mukherjee et al. 2017a). Geochemical and isotopic studies of the host
felsic gneiss confirm that they have a ferroan (A type) character (Mukherjee et al.
2017a, 2018). The ortho-gneisses show a prominent N-S trending migmatitic
banding and this regional fabric locally swerves around the enclaves. The pelitic
enclaves contain voluminous (>30%) leucosomal segregations (S1; Fig. 3a). The S1
which are discordant to and are dragged to parallelism with the pervasive foliation
(S2) of the host felsic gneiss (Fig. 3a). Numerical modelling with appropriate bulk
for the metapelites constrain an early high grade event M1 occurred at 7 ± 1 kbar
and 1000 ± 50 °C i.e. at MP (medium pressure)-UHT condition which generated
voluminous S1 leucosomal foliation (Dey et al. under review, a). This event was
followed by another metamorphism whereby the felsic orthogneisses and the
meta-sedimentary gneisses developed a prominent migmatitic foliation that are
currently *N-S trending. Numerical modelling combined with conventional
geo-thermobarometry of the host gneiss as well as the pelitic enclaves constrain the
peak of this metamorphism at 770 ± 50 °C and 9 ± 1 kbar (Dey et al. under
review, a; Chatterjee et al. 2008; Mukherjee et al. 2017a). However, petrological
study of a suite of mafic enclave reveals much higher pressure (12 ± 1 kbar and
800 ± 50 °C) for the same (Dey et al. under review, b). In all these studies
16 S. Mukherjee et al.
Fig. 3 a Internal foliation (S1) of metapelitic enclaves and external foliation (S2) within host felsic
gneisses; b folding of S2, designated by coarse leucosomes, and development of axial planar S3
within felsic orthogneiss; c folding of the mafic dyke during D2 and d development of S3 along the
axial planes of folded mafic dyke
P–T of 7.3 ± 0.1 kbar, 615 ± 15 °C and 4.3 ± 0.7 kbar, 600 ± 60 °C, respec-
tively for this metamorphism (Dey et al. under review, a). Bhattacharjee et al.
(2012) reported a gabbro-anorthosite body (Hizla anorthosite) near Dumka area,
Jharkhand. This anorthosite body intruded the charnockitic country rock and is
deformed and metamorphosed along with the latter rock. These rocks that are likely
to be the product of D2-M2 tectonothermal event recorded an unusually wide range
of pressure and temperature of (511–915 °C and 5.0–7.5 kbar) for the metamor-
phism of the anorthosite (Bhattacharjee et al. 2012). The published petrological
information is not robust enough for draw any conclusion about this wide range of
metamorphic P–T values. Multiple generations of pegmatitic veins criss-cross all
the lithounits. No structural data on the orientation of these veins are available.
Eastern margin of the Domain I is marked by a highly tectonized N- to NNE-zone
shear deformation. Shear driven non-cylindrical folds (D2-D3 fold interference) has
been described by Chatterjee et al. (2010). This tectonic zone is termed as the
Eastern Indian Tectonic Zone (EITZ; Chatterjee et al. 2010). A linear gravity high
follows the trend of EITZ over a length of approximately 400 km (Singh et al.
2004). From a quartzo-feldspathic gneiss from northern part of Dumka, Chatterjee
et al. (2010) has quantified the peak conditions of the shearing and anatexis along
EITZ at *11 kbar and <800 °C followed by isothermal decompression.
U–Pb zircon and monazite age of the metapelitic enclaves attests the age of M1
metamorphism at 1630 ± 50 Ma (Dey et al. under review, a). Similar
Paleoproterozoic metamorphic ages has also been obtained by Rekha et al. (2011)
and by Chatterjee et al. (2010) from metapelites and anatectic quartzofeldspathic
gneiss near Dumka. From detrital zircon study of the metapelitic enclaves, Dey
et al. (2017) has constrained the age of deposition of the sedimentary protoliths of
these enclaves within 1700–1680 Ma and has ascribed that the sediments were
derived from Archean and Paleoproterozoic sources (Rekha et al. 2011; Dey et al.
2017). The protolith of the felsic orthogneiss was derived via high temperature
partial melting of Paleoproterozoic crustal source (*1800–1600 Ma) with limited
mantle input (Lu–Hf model age: Mukherjee et al. 2018) and was presumably
emplaced in a continental rift setting at *1450 Ma (U–Pb zircon dating;
Mukherjee et al. 2017a, 2018). Mesoproterozoic dates have also been reported from
U–Pb dating of zircon from metapelites and anatectic quartzofeldspathic gneiss in
this domain (Rekha et al. 2011; Dey et al. under review, a). Rb–Sr whole-rock date
the emplacement of a syenitic rock (1457 ± 63 Ma) and a charnockite
(1331 ± 125 Ma) in Dumka-Jamua, Jharkhand area of Jharkhand (Ray Barman
et al. 1994). U–Pb zircon and Th–U–Pb monazite age constrained the age of the
granulite grade metamorphism at *1000–930 Ma (Dey et al. under review, a;
Chatterjee et al. 2008, 2010; Rekha et al. 2011; Mukherjee et al. 2017a). Chatterjee
et al. (2010) has retrieved a younger age population of *870–650 Ma from the
monazite dates, which they linked with the tectonothermal activity and sinistral
shearing along the EITZ. Zircon and monazite from metapelite and felsic gneiss
yield ages for D3 and M3, ranging within 900–800 Ma (Chatterjee et al. 2010;
Dey et al. under review, a; Mukherjee et al. 2017a) with very few ages as low
18 S. Mukherjee et al.
as *650 Ma (Chatterjee et al. 2010). These age of the D3 deformation of the area
have been correlated with the age duration of shearing along the EITZ.
Not much geological and geochronological information are available from fur-
ther west of Jasidih, in and around Dhanbadand and Hazaribag (Jharkhand). It is
noteworthy that the strike of the schists and gneisses of these areas varies between
E-W to NW-SE. Parts of the Dhanbad district (Jharkhand) exposes banded gneisses
of amphibolite facies (Roy Chowdhury 1979) with enclaves of hornblende schists
and gneisses, olivine-orthopyroxene bearing metanorites and quartzites. The rocks
folded thrice with development of E-W to NW-SE trending foliation axial planar to
the second generation folding (Sarangi and Mohanty 1998).
Further west, near Hazaribagh (Jharkhand), amphibolites, pelitic and calcareous
rocks dominate with a persistent E-W trending schistosity. The rocks have also
been folded thrice with first two folds being tight and coaxial, and the last one being
open warp (Roy Chowdhury 1979; Mahadevan 2002). The whole rock Rb–Sr
isochron date of biotite-K-feldspar from a migmatite record one of the youngest
ages (481 ± 18 Ma; Pandey et al. 1986b).
The area around Daltonganj (Jharkhand) exposes granite gneisses and migma-
tites interbanded with Graphite-bearing pelitic schists, limestones, quartzites and
intruded by mafic-ultramafic, anorthositic-komatiitic rocks (Bhattacharya et al.
2010). These ultramafic rocks host magnetite (Ghose 1983; Sinha and Bhattacharya
1995), fluorite (Soni et al. 1991) and base metal mineralization (Ghose 1992;
Bhattacharya et al. 2010). Thermometry of granite gneiss confirms that the gneissic
foliation formed under high grade condition (850–800 °C; Chatterjee and Ghosh
2011). The granite gneisses are intruded by non-foliated granitic to granodioritic
rocks (Rode, 1948; Ghose 1983; Mazumdar 1988). Anatexis of granulite to form
leucogranite, and partial melting of metapelite forming these non-foliated granitoids
probably occurred at 650–800 °C and 2–6 kbar (Srivastava and Ghose 1992).
Whole-rock Rb–Sr dating of the granite gneiss suggests an event of granitic
magmatism at 1741 ± 65 Ma (Ray Barman and Bishui 1994) followed by high
grade metamorphism and anatexis. Chatterjee and Ghosh (2011) has dated this high
grade metamorphism from xenotime at 975 ± 67 Ma. The granite gneiss and
metapelitic sequence deformed thrice (D1-D3) (Lahiri and Das 1984; Patel 2007).
Interference of D1 and D2 produced NW-SE trending domal structure with devel-
opment of prominent axial planar schistosity. Post-D3 pegmatite veins and mafic
dykes are common in the area. In the Tatapani area, the calc-silicate rocks develop
the rare assemblage vesuvianite + grossular garnet + diopside + wollastonite +
quartz. Patel (2007) estimated the conditions of vesuvianite formation at <4 kbar
and 590–650 °C in presence of a highly aqueous fluid.
3.3 Domain II
Sandwiched between Domain III (in the north) and GBF (in the south), this domain
is the ENE-WSW trending the Bihar Mica belt (BMB) exposes a distinct
Proterozoic Crustal Evolution of the Chotanagpur Granite … 19
lithological ensemble in the CGGC (Fig. 2a). The geological formation in this belt
include meta-sedimentary sequence dominantly constituting muscovite-biotite
schist interbedded with micaceous quartzites, conglomerates, calc-silicate rocks
and hornblende schists. The carbonate metasediments are occasionally found
associated with economic grade base-metal mineralization (Ghose 1992). These
metasedimentary rocks of BMB rest on the high grade rocks of CGGC with an
erosional conglomerate base (Ghose and Mukherjee 2000). The Pb/Pb age of galena
from base-metal deposits brackets the sedimentation age in the BMB within the
time span of 1700–1650 Ma (Singh et al. 2001). Rb–Sr whole rock dating of a
migmatitic granite gneissic basement from south of BMB yields an age of
1717 ± 102 Ma (Mallik et al. 1991). The meta-sedimentary package recorded three
phases of folding (D1, D2 and D3), where first two folds are most dominant and
produce E-W trending axial planar foliation while the third fold is weak and less
pervasive. The whole litho-unit is extensively intruded by large bodies of granitoid
rocks, dolerite dykes, gabbro anorthosite and mica pegmatites (sometimes REE and
rare metal rich). Granitoids were emplaced syn- to post-D2 folding with ages of
most of the granite intrusives clustering within 1300–1100 Ma and age of oldest
granite intrusion around 1600 Ma (Whole-rock and Rb–Sr mineral isochron ages;
Pandey et al. 1986a; Misra and Dey 2002). K–Ar dating of granitoids yields lower
ages (1080–850 Ma) (Sarkar 1980). The BMB plutons characteristically show very
high SrI ratio which suggests their derivation from sialic crustal sources (Misra and
Dey, 2002). Different generations of mica-bearing granite–pegmatites emplaced
pre-D2 to post-D3 folding (Mahadevan 2002). The conditions of emplacement as
well as formation of the BMB granite plutons is estimated *5 kbar and slightly
>1000 °C in a relatively anhydrous condition within a possible post-orogenic set-
ting (Misra and Dey 2002). The age of emplacement of the oldest REE-rare
metal-bearing mica pegmatites of the BMB has been restricted within
960 ± 50 Ma (Pb/Pb age; Vinogradov et al. 1964) and 910 ± 20 Ma (U–Pb and
Pb/Pb ages of the Columbite–Tantalite minerals; Krishna et al. 2003). Fission-track
ages of garnet (830 Ma), muscovite (760 Ma), biotite (590 Ma) and apatite
(590 Ma) date the final exhumation of the mica pegmatites of this region (Lal et al.
1976). It suggests that the rocks of this area had not received any significant thermal
pulse, which by implication means cratonization, since 590 Ma (Lal et al. 1976). In
summary, this domain bears history of Paleoproterozoic and Mesoproterozoic
granitic activities. Evidence of Paleoproterozoic sedimentation is noteworthy. The
abundant intrusion of pegmatites took place in Neoproterozoic which cooled
gradually.
The northern fringe of CGGC, north of the Bihar mica Belt, exposes an ensemble of
migmatitic quartzofeldspathic gneisses and supracrustals. Sedimentary rocks of the
Munger Group and the Rajgir Group showing basement–cover relationship with the
20 S. Mukherjee et al.
4 Discussions
This section collates the geological and geochronological information (Table 1) and
reconstructs a plausible event stratigraphy that shaped the CGGC.
Table 1 Available geochronological data from different domains of CGGC; ages are in Ma
Age Method Type of age References
Domain IA
595 U–Pb monazite Alteration Chatterjee et al. (2010)
784 U–Pb monazite Metamorphism Chatterjee et al. (2010)
810 ± 40 K–Ar muscovite Magmatism Baidya et al. (1987)
815 ± 47 Rb–Sr whole rock Metasomatism Singh and Krishna (2009)
825–818 U–Pb monazite Metamorphism Maji et al. (2008)
850–775 U–Pb monazite Deformation event Karmakar et al. (2011)
859 U–Pb monazite Metamorphism Chatterjee et al. (2010)
860–830 U–Pb monazite Metamorphism Chatterjee et al. (2010)
870 ± 40 K–Ar biotite Magmatism Baidya et al. (1987)
921 U–Pb monazite Metamorphism Rekha et al. (2011)
921 U–Pb monazite Metamorphism Rekha et al. (2011)
923 U–Pb zircon Metamorphism Rekha et al. (2011)
928 U–Pb zircon Metamorphism Rekha et al. (2011)
937 U–Pb monazite Metamorphism Chatterjee et al. (2010)
944 U–Pb monazite Metamorphism Rekha et al. (2011)
946 U–Pb monazite Metamorphism Chatterjee et al. (2010)
947 ± 27 U–Pb zircon Metamorphism Chatterjee et al. (2008)
965 U–Pb monazite Metamorphism Chatterjee et al. (2010)
990–940 U–Pb monazite Metamorphism Karmakar et al. (2011)
1005 ± 51 Rb–Sr whole rock Magmatism Singh and Krishna (2009)
1021–967 U–Pb monazite Metamorphic/magmatic Maji et al. (2008)
1025 U–Pb zircon Metamorphism Rekha et al. (2011)
1059 ± 104 Rb–Sr whole rock Magmatism Krishna et al. (1996)
1065 ± 74 Rb–Sr whole rock Magmatism Krishna et al. (1996)
1071 ± 64 Rb–Sr whole rock Retrogression Ray Barman et al. (1994)
1072 U–Pb zircon Metamorphism Rekha et al. (2011)
1138 ± 193 Rb–Sr whole rock Magmatism Krishna et al. (1996)
1176 U–Pb monazite Metamorphism Maji et al. (2008)
1178 ± 61 Rb–Sr whole rock Retrogression Ray Barman et al. (1994)
1200–1100 U–Pb monazite Older thermal events Karmakar et al. (2011)
1239 U–Pb zircon Metamorphism Rekha et al. (2011)
1331 ± 42 Rb–Sr whole rock Krishna et al. (1996)
1422–1305 U–Pb monazite Older thermal events Chatterjee et al. (2010)
1550 ± 12 U–Pb zircon Magmatism Chatterjee et al. (2008)
1717–1446 U–Pb monazite Older thermal event?? Chatterjee et al. (2010)
1800 U–Pb monazite Inherited Karmakar et al. (2011)
1870–1691 U–Pb monazite Older thermal event?? Chatterjee et al. (2010)
(continued)
22 S. Mukherjee et al.
Table 1 (continued)
Age Method Type of age References
Domain IB
481 ± 18 Rb–Sr Bt, Kfs ? Pandey et al. (1986a)
649 U–Pb monazite Metamorphism Chatterjee et al. (2010)
850–750 U–Pb monazite Metamorphism Sanyal et al. (2007)
872–838 U–Pb monazite Metamorphism Chatterjee et al. (2010)
876 U–Pb monazite Metamorphism Chatterjee et al. (2010)
902 U–Pb zircon Metamorphism Mukherjee et al. (2017a)
907 ± 49 U–Pb zircon Metamorphism Mukherjee et al. (2018)
910 U–Pb monazite Metamorphism Rekha et al. (2011)
937 U–Pb monazite Metamorphism Chatterjee et al. (2010)
943 U–Pb zircon Metamorphism Mukherjee et al. (2017a)
945 U–Pb monazite Metamorphism Rekha et al. (2011)
948 ± 22 U–Pb zircon Metamorphism Mukherjee et al. (2018)
950 ± 20 U–Pb monazite Retrogression Chatterjee et al. (2008)
954 U–Pb monazite Metamorphism Chatterjee et al. (2010)
965–930 U–Pb monazite Metamorphism Chatterjee et al. (2010)
975 ± 67 U–Pb monazite Metamorphism Chatterjee and Ghose (2011)
995 ± 24 U–Pb monazite Metamorphism Chatterjee et al. (2008)
1009 U–Pb zircon Metamorphism Rekha et al. (2011)
1100–930 U–Pb monazite Metamorphism Sanyal et al. (2007)
1119 Rb–Sr whole rock ? Sarkar et al. (1986)
1183 U–Pb monazite Older thermal events Chatterjee et al. (2010)
1190 ± 26 U–Pb monazite Metamorphism Chatterjee et al. (2008)
1270 U–Pb zircon ? Rekha et al. (2011)
1272 U–Pb monazite Older thermal events Chatterjee et al. (2010)
1278 U–Pb monazite Older thermal events Chatterjee et al. (2010)
1331 ± 125 Rb–Sr whole rock Cooling age Ray Barman et al. (1994)
1333 U–Pb zircon ? Rekha et al. (2011)
1377 U–Pb zircon ? Rekha et al. (2011)
1435 U–Pb zircon ? Rekha et al. (2011)
1446 ± 7 U–Pb zircon Magmatism Mukherjee et al. (2018)
1447 U–Pb zircon Magmatism Mukherjee et al. (2017a)
1457 ± 63 Rb–Sr whole rock Older thermal events Ray Barman et al. (1994)
1462 U–Pb zircon ? Rekha et al. (2011)
1470 ± 2 U–Pb zircon Magmatism Mukherjee et al. (2018)
1480 U–Pb monazite ? Rekha et al. (2011)
1515 ± 5 U–Pb zircon Magmatism Acharyya (2003)
(continued)
Proterozoic Crustal Evolution of the Chotanagpur Granite … 23
Table 1 (continued)
Age Method Type of age References
1522 ± 71 Rb–Sr whole rock Magmatism Mallik et al. (1991)
1580 ± 33 Rb–Sr whole rock Magmatism Mallik et al. (1991)
1599 ± 33 Rb–Sr whole rock Magmatism Mallik et al. (1991)
1624 ± 5 U–Pb zircon Magmatism Acharyya (2003)
1649 U–Pb zircon Metamorphism Rekha et al. (2011)
1660–1270 U–Pb monazite Metamorphism Sanyal et al. (2007)
1720 U–Pb monazite Older thermal event Chatterjee et al. (2010)
1741 ± 65 Rb–Sr whole rock ? Ray Barman and Bishui (1994)
1824–1659 U–Pb monazite Older thermal event Chatterjee et al. (2010)
1870–1720 U–Pb monazite UHT metamorpism Sanyal et al. (2007)
2600–1900 U–Pb zircon Detrital grains Rekha et al. (2011)
Domain II
590 Apt Fission Track Cooling age Lal et al. (1976)
595 Bt Fission Track Cooling age Lal et al. (1976)
760 mus Fission Track Cooling age Lal et al. (1976)
830 grt Fission Track Cooling age Lal et al. (1976)
855 ± 25 Mica Rb–Sr Magmatism Pandey et al. (1986b)
910 ± 19 U–Pb, Pb–Pb Magmatism Krishna et al. (2003)
960 ± 50 Pb–Pb Magmatism Vinogradov et al. (1964)
1086–850 K–Ar ? Sarkar (1980)
1020 ± 46 Rb–Sr whole rock Magmatism Mallik (1993)
1100–700 Fission track mica Cooling age Lal et al. (1976)
1242 ± 34 Rb–Sr whole rock ? Pandey et al. (1986a)
1238 ± 33 Rb–Sr whole rock Magmatism Mallik (1993)
1285 ± 108 Rb–Sr whole rock Magmatism Mallik (1993)
1300–1100 Rb–Sr whole rock ? Pandey et al. (1986a, b)
1590 ± 30 Rb–Sr whole rock Magmatism Pandey et al. (1986b)
Domain III
557 ± 99 U–Pb xenotime ? Chatterjee and Ghose (2011)
768 ± 11 U–Pb uraninite Metamorphism Chatterjee and Ghose (2011)
929 U–Pb xenotime Metamorphism Chatterjee and Ghose (2011)
1044 ± 35 U–Pb zircon Magmatism Wanjari et al. (2012)
1337 ± 26 Rb–Sr whole rock Magmatism Wanjari et al. (2012)
1583 ± 50 U–Pb xenotime Cooling age Chatterjee and Ghose (2011)
1697 ± 17 U–Pb Monazite Magmatism Chatterjee and Ghose (2011)
1737–1664 U–Pb zircon Magmatism Saikia et al. (2017)
24 S. Mukherjee et al.
authigenic datable minerals, few studies from northern part of CGGC (domain IB &
II) have attempted to constrain the timing of sedimentation of the protolith of
meta-sedimentary rocks. From the Pb–Pb dating of galena Singh et al. (2001) has
inferred a sedimentation age of 1700–1650 Ma in parts of BMB. Dey et al. (2017)
has analysed and studied 207Pb/206Pb apparent dates of the detrital zircons from
UHT pelitic granulite from Dumka and Deoghar (Jharkhand) (Fig. 2a). The ages of
the youngest analysed detrital zircon cores and the oldest metamorphic overgrowth
constrain the age of sedimentation of the precursor of metapelites within a narrow
age bracket of *1700–1680 Ma. From the detrital zircon dates, the authors sug-
gested that protoliths of these sediments were sourced from the *2700–1700 Ma
old domains in the adjoining cratonic areas and also from the rocks of similar age
now occurring in the Lesser Himalayan region. Both the studies are consistent with
the interpretation that during *1700–1650 Ma the (unknown) basement of CGGC
experienced a phase of basin formation and sedimentation (Fig. 4a). A part of these
sediments got subsequently metamorphosed by later tectonothermal events.
Proterozoic Crustal Evolution of the Chotanagpur Granite … 25
The oldest tectonothermal event reported so far from northern and central CGGC
occurred in the late Paleoproterozoic (ca. 1750–1640 Ma) (Mallik et al. 1991; Ray
Barman and Bishui 1994; Chatterjee and Ghose 2011; Saikia et al. 2017). U–Pb
zircon ages distinctly indicates the intrusion of arc-related bimodal volcanics in the
northern part (Saikia et al. 2017). However, contemporaneous (1697 Ma; Chatterjee
and Ghose 2011) porphyritic granites from the northern domain are inferred to be of
anorogenic affinity (Yadav et al. 2014). In the eastern fringe of the terrain, intrusion
of massif anorthosite body, occurred at 1550 Ma and inferred to be the manifes-
tation of post-orogenic magmatism (Fig. 4b) (Chatterjee et al. 2008). Subsequently
in the eastern part of the domain Mesoproterozoic (ca. 1450 Ma) ferroan granites
have intruded the Paleoproterozoic crust in an extensional setting (Fig. 4c)
(Mukherjee et al. 2017a, 2018). A body of syenite near Dumka (Jharkhand) yields a
Rb–Sr whole rock age of 1457 ± 63 Ma (Ray Barman and Bishui 1994). Early
Neoproterozoic felsic magmatism so far has been reported only from southern part
of the CGGC (Singh and Krishna 2009). However, post-Grenvillian (<950 Ma)
magmatic pulses, including granite, syenite, mafic dykes and pegmatite intrusions
have been documented from different parts the terrain (Fig. 4e). Biotite bearing
granites have considered to intrude the early crust *870–810 Ma (Baidya et al.
1987). Although no geochronological dates are available, a body of syenite intruded
the granulitised crust in the southwestern part of the terrain, associated with NPSZ
(Das et al. 2016, 2017c). A set of mafic dykes also intruded the granulitised crust in
the northeastern part of the CGGC (Sanyal and Sengupta 2012; Mukherjee et al.
2017a). Numerous mica-bearing pegmatites with intrusion age of 960–855 Ma
have been reported from the northern part of the terrain (Pandey et al. 1986b;
Krishna et al. 2003; Vinogradov et al. 1964).
1000–950 Ma (M2). From the northeastern margin of the terrain, several workers
(Mukherjee et al. 2017a, Dey et al. under review, a, Dey et al. under review, b) have
inferred that the metamorphism culminated at 770–800 °C and 9–12 kbar pressure,
followed by a steep decompressive path. Similar P–T path has been recorded from the
southern part of the terrain where the peak culminated at 870 °C and 11 kbar followed
by decompression (Karmakar et al. 2011). Such high pressure event associated with
clockwise path is inferred to be the manifestation of continent-continent collision
(Fig. 4d). Post-Grenvillian metamorphic event (M3) has been reported from north-
eastern margin of the terrain where U–Th–Pb monazite constrains the age of the event
between 900 and 780 Ma (Chatterjee et al. 2010; Mukherjee et al. 2017a, Dey et al.
under review, a). Chatterjee et al. (2010) estimated a high-pressure peak for M3
culminating at 820 °C and 11 kbar which correlates with the development of Eastern
Indian tectonic Zone (Fig. 4f).
Structural analyses from different parts of the terrane reveal that the CGGC has
experienced multiple deformational events (Mahadevan 2002; Maji et al. 2008;
Goswami and Bhattacharyya 2010; Karmakar et al. 2011; Sanyal and Sengupta
2012; Mukherjee et al. 2017a). However, structural data from scattered places and
scarcity of detailed geochronological information are some of the major hindrance
for correlating the different deformational and metamorphic events reported by
different workers over the entire CGGC. Unfortunately, sizable amount of structural
and geochronological information are available only from the rocks of Domain I. In
the following section an attempt has been made to develop a structural history of the
CGGC using the structural, metamorphic and geochronological information from
Domain I. Three major deformational phases seem to have affected the rocks of
Domain I. The earliest deformational event (D1) is preserved in the metasedi-
mentary enclaves within the felsic orthogneiss. This event is marked by a mig-
matitic banding (S1) which is inferred to be associated with M1 metamorphic event
that is dated to be *1650 Ma (Fig. 4a) (Sanyal and Sengupta 2012; Mukherjee
et al. 2017a; Dey et al. under review a; Dey et al. under review, b). The next major
deformational event (D2) is manifested by the development of S2 foliations within
the felsic orthogneisses and transposition of S1 within the metasedimentary
enclaves. The D2 is associated with *1000–950 Ma old granulite facies meta-
morphism (M2) (Fig. 4c), especially from the eastern margin of the terrane (Maji
et al. 2008; Karmakar et al. 2011; Mukherjee et al. 2017a; Dey et al. under review a;
Dey et al. under review, b). Folding of the late intrusives (mafic dykes in the
northeastern part and nepheline syenite in the southern part), along with the S2 of
the host felsic orthogneisses, depicts the third major deformational event (D3).
There is no detail information on the age of emplacement of the mafic dykes and
subsequent D3 deformation. The EITZ and with its characteristic asymmetrical
folds, N-NNE trending planar fabric and high pressure metamorphism appear to be
coeval with the D3-M3 tectonothermal event (Fig. 4f). Similar shear-related folds
also reported by Mukherjee et al. (2015b) and Mukherjee (submitted manuscript).
Proterozoic Crustal Evolution of the Chotanagpur Granite … 27
Several prominent crustal-scale shear zones cross through the CGGC. The
‘Tamar-Porapahar-Khatra’ shear zone is the southernmost lineament affecting the
Proterozoic gneissic rocks of CGGC and separating it from the low grade meta-
morphites of North Singhbhum Fold Belt (NSFB). This ENE-WSW trending
150 km long shear zone passes through the south of the Purulia town (West Bengal)
and also known as ‘South Purulia Shear Zone’ (SPSZ). *150 km long, the ‘North
Purulia Shear Zone (NPSZ) is another prominent E-W to NE-SW trending linea-
ment that affects the CGGC rocks occurring north of Purulia.
Approximately 150 km long arcuate shear zone which is referred to as the South
Purulia Shear Zone (SPSZ: Mazumdar 1988), or Northern Shear Zone (NSZ: Kumar
et al. 1978), or Tamar-Porapahar-Khatra Fault Zone (TPKF: Mahadevan 2002), is a
crustal-scale ENE-WSW trending brittle-ductile shear zone that marks the southern
margin of the CGGC and separates the terrain from the low to medium grade
meta-sedimentary and meta-igneous rocks of the North Singhbhum Fold Belt
(Fig. 2a; Mahadevan 1992; Acharyya et al. 2006; Mahato et al. 2008; Sanyal and
Sengupta 2012). This shear zone has been extensively deformed during several
tectonothermal events (Bhattacharyya et al. 1992; Sengupta et al. 2005) and represents
an ensemble of a nepheline-bearing syenite, tuffaceous rocks, tourmalinite, carbon-
atite, apatite bearing rocks and granitoids altered hydrothermally (Banerji 1985; Basu
1993; Acharyya et al. 2006; Chattopadhyay et al. 2015a). Abundance of carbonatite
and alkaline connotes crustal rifting (Acharyya et al. 2006; Chakrabarty and Sen
2010). Intrusion of the alkaline rocks at ca. 922 Ma, evident from the U–Pb zircon
geochronology (Reddy et al. 2009), constrains the timing of the extensional event,
which was followed by another tectonothermal event that deformed the
carbonatite-alkaline rocks forming DARCs (Deformed Alkaline Carbonatite Rocks)
(Chattopadhyay et al. 2015a). Economically viable apatite deposit, tuffaceous rocks
and Fe-ore are reported from the area (Banerji 1985).
The NPSZ is an E-W to ENE-WSW trending crustal-scale shear zone with a
prominent steep northerly dip traversing the metamorphites of CGGC. This lineament
passes through Jhalda-Jaypur in the west through Raghunathpur in the central part and
can be traced up to north of Murlu near Saltora (West Bengal) in the east (Baidya et al.
1989; Mahadevan 1992; Dasgupta et al. 2000; Som et al. 2007; Maji et al. 2008;
Goswami and Bhattacharyya 2010). The shear zone is best developed in the granitoids
where an early mylonitic fabric has developed in deep structural level in amphibolite
facies condition. The crystal plastic deformation of quartz and feldspar indicates the
temperature *650–700 °C during this early ductile deformation (Vernon 2004).
The NPSZ has the signatures of protracted period of ductile deformation where the
early shear foliation folded. Later stage of brittle deformation has affected the early
mylonitic fabric of the NPSZ. Discrete bodies of nepheline syenite is found to be
emplaced along this crustal lineament near Kankarkiari–Kusumda area near Saltora
(West Bengal) cross-cutting the gneissic fabric of felsic gneiss and khondalite and get
metamorphosed and deformed when NPSZ formed (Goswami and Bhattacharyya
28 S. Mukherjee et al.
2010; Das et al. 2017b). Das et al. (2017a) has also reported carbonatite in the central
part of NPSZ where apatite-Fe-oxide/hydroxide-silica bearing veins coexist with
alkali metasomatised granitoids. In the western part of NPSZ late stage pegmatite
veins within granite and calc-silicate rocks have been explored for Cs, Li and other
rare elements (Som et al. 2007).
The Eastern Indian Tectonic Zone (EITZ), roughly N- to NNW trending
crustal-scale shear zone, is observed towards the eastern margin of the CGGC
(Fig. 2a) (Chatterjee et al. 2010). The shear fabric, manifested by asymmetric folds,
developed over the older lithounits and is traceable along CGGC, North Singhbhum
Fold Belt and Singhbhum Craton. Detailed petrological studies by Chatterjee et al.
(2010) have shown that the metamorphism associated with the development of the
shear zone culminated at *11–12 kbar and 800 °C. Monazite geochronology
constrains the timing of the of the tectonothermal event at ca. 870–780 Ma, which
agrees with the field observations that EITZ developed after the major tectonic
event at CGGC i.e. granulite grade metamorphism dated *950 Ma.
Quaternary sediments of Gangetic alluvium cover the northern margin of the CGGC.
Further north, the Late Paleoproterozoic rocks of the Lower Lesser Himalaya occurs
interleaved with the Phanerozoic rocks (Mukherjee et al. 2013, 2015a; Mukherjee
2013, 2015). These units expose metamorphosed and unmetamorphosed sedimen-
tary sequence and granitic plutons of age *1900–1700 Ma (Fig. 5). Geochemical
signatures of the granitoids point towards their emplacement in a continental arc
setting (Kohn et al. 2010) existing along the northern boundary of India at that time.
Detrital zircon populations of the sedimentary units are characterized by
Paleoproterozoic–Late Archean ages (2.6–1.8 Ga), with no younger population
(Richards et al. 2005). According to Richards et al.(2005) “Detrital zircon ages are
younger than their respective Hf-isotope derived crustal formation ages by 0.7–
2.1 Ga, indicating that the source regions of the detrital zircons consisted of older
terranes with considerable amounts of reworking and renewed magmatism”. Many
workers have suggested that these Proterozoic rocks of the Lesser Himalaya con-
stitute the remnants of the Greater Indian Landmass, subducted below the
Himalayan orogenic belt (DeCelles 2000). Petrological and limited geochronolog-
ical studies suggest that the Paleoproterozoic felsic orthogneisses exposed within the
quaternary alluvium towards the north of the CGGC resembles the basement
orthogneisses of the CGGC. Felsic magmatism in the northern part of the CGGC,
near Bathani, shows enrichment of Th, U, and Pb over Nb, Ta and Ti which cor-
roborates with the geochemical characters of the Paleoproterozoic granitoids of the
lower Lesser Himalaya. Elevated magmatic temperature, derived from the zircon
saturation thermometry and predominance of crustal source also poses the possibility
Proterozoic Crustal Evolution of the Chotanagpur Granite … 29
The Mohakoshal Belt (MB) and the Sausar Mobile Belt (SMB) occur to the NW
and SW of CGGC respectively (Fig. 2a) and constitute the northern and the
southern structural blocks of Central Indian Tectonic Zone (CITZ) that sutures two
Achaean blocks: Bundelkhand at the north and Bastar at the south (Acharyya 2003;
Jain et al. 1991; Mishra et al. 2000; Radhakrishna and Naqvi 1986; Yedekar et al.
1990).
The Mahakoshal belt is composed of intercalated riftogenic volcano-sedimentary
successions (Roy and Prasad 2003 and the references therein), metamorphosed and
intruded with post-orogenic granitoids aging 1800–1700 Ma (Rb/Sr whole rock age
and U/Pb zircon age; Bora et al. 2013; Pandey et al. 1998; Sarkar 1998; Srivastava
et al. 2000). The sedimentary sequence was metamorphosed in low pressure–
medium temperature conditions in a possible back arc setting (Roy and Prasad
2003; Wani and Mondal 2016). On contrary, a recent study proposed that the
sediments were metamorphosed in moderate pressure–low temperature conditions
in a collisional setting followed by rapid exhumation (Deshmukh et al. 2017).
The Sausar Mobile Belt occurs at the southern margin of CITZ and is a collage
of different litho-tectonic components. The southern Bhandara-Balaghat granulite
(BBG) domain exposes enclaves of supracrustal and meta-igneous granulites
occurring within felsic gneisses. They have been metamorphosed under lower
crustal, ultra-high temperature (UHT) granulite facies conditions along a
counter-clockwise P–T path (Bhowmik 2006; Bhowmik et al. 2005) in a possible
back arc setting associated with emplacement of arc magma (Bhowmik et al. 2011).
Although early workers predicted an Archean–Early Paleoproterozoic age
(Bhowmik 2006; Bhowmik et al. 2005; Roy et al. 2006), precise U–Pb zircon and
texturally controlled monazite dating constrains the timing of this metamorphism
and magmatism at *1600 Ma (Bhandari et al. 2011; Bhowmik et al. 2011, 2014).
This event was followed by several events of granulite-amphibolite grade over-
printing (BasuSarbadhikari and Bhowmik 2008; Bhowmik et al. 2014) with a
terminal phase of metamorphism at *1400 Ma (Bhandari et al. 2011; Bhowmik
et al. 2011). The central Sausar Group of rocks deformed multiply and metamor-
phosed in greenschist to amphibolite facies during *1060–950 Ma in a collisional
orogeny (Bhowmik et al. 1999, 2011, 2012; Bhowmik and Roy 2003;
Chattopadhyay et al. 2015b; Pal and Bhowmik 1998) accompanied by felsic
magmatism (Chattopadhyay et al. 2015b; Pal and Bhowmik 1998). The northern
Ramakona-Katangi granulite (RKG) domain contains mafic, felsic and pelitic rocks
as rafts, which are metamorphosed in high pressure upper amphibolite to granulite
facies along a clockwise P–T path indicating a continent-continent collision
(Bhowmik and Roy 2003; Bhowmik and Spiering 2004) at 1040–950 Ma
(Bhowmik et al. 2012). This was followed by emplacement of post-orogenic
30 S. Mukherjee et al.
Fig. 5 Timing of magmatic and metamorphic events in different Proterozoic mobile belts of India, c
Lower Lesser Himalaya and east Antarctica. Age scale changes at 1000 Ma. References: CITZ
(Central Indian Tectonic Zone): Bhandari et al. (2011), Bora et al. (2013), Bhowmik et al.
(2005), Roy et al. (2006), Panigrahi et al. (2004), Deshmukh et al. (2017), Pandey et al. (1986b),
Bhowmik et al. (2011); Aravalli-Delhi Fold Belt: Kaur et al. (2009), (2011), (2017a, b); Buick
et al. (2006), Chatterjee et al. (2017), Deb et al. (2002), Biju-Sekhar et al. (2003), Pandit et al.
(2003), Bhowmik et al. (2010), Wiedenbeck et al. (1996), Dharma Rao et al. (2011, 2013),
Mukhopadhyay et al. (2000), Gupta et al. (1998), Sivaraman and Raval (1995); North
Singhbhum Fold Belt: Bhattacharya et al. (2015), Mahato et al. (2008), Rekha et al. (2011),
Ramakrishnan and Vaidyanadhan (2008), Roy et al. (2002); CGGC: Chatterjee et al. (2008),
Chatterjee and Ghose (2011), Karmakar et al. (2011), Rekha et al. (2011), Chatterjee et al. (2010),
Mallik et al. (1991), Ray Barman and Bishui (1994), Dey et al. (2017), Mukherjee et al. (2017a,
2018), Saikia et al. (2017), Sanyal et al. (2007); Lower Lesser Himalaya: Kohn et al. (2010),
Sakai et al. (2013), Larson et al. (2016), Chambers et al. (2008), Treloar and Rex (1990), Richards
et al. (2005), Miller et al. (2000), Célérier et al. (2009), DiPietro and Isachsen (2001), Yin et al.
(2010), Liao et al. (2008), Upreti et al. (2003), Decelles et al. (2000), Long et al. (2008), Daniel
et al. (2003); EGMB (Eastern Ghats Mobile Belt): Dobmeier and Simmat (2002), Simmat and
Raith (2008), Bose et al. (2011), Das et al. (2011), Upadhyay et al. (2006), Upadhyay and Raith
(2006), Sarkar and Schenk (2016), Das et al. (2017a, b, c), Rickers et al. (2001), Korhonen et al.
(2013), Krause et al. (2001), Dharma Rao et al. (2012), Mezger and Cosca (1999), Dobmeier and
Raith (2003), Henderson et al. (2014); SMGC (Shillong-Meghalaya Gneissic Complex): Kumar
et al. (2017), Chatterjee et al. (2007); East Antarctica: Kelly et al. (2002), Corvino et al. (2008),
Liu et al. (2013, 2017, 2009), Morrissey et al. (2015), Tsunogae et al. (2014, 2016), Owada et al.
(2003), Zhang et al. (2012b), Asami et al. (2002), Elburg et al. (2015), Goodge et al. (2008)
In the eastern margin, the CGGC is bordered by the Cenozoic sediments of Bengal basin
that ranges more than 200,000 km2 comprising the largest alluvium delta in the world
covering parts of West Bengal and Tripura and Bangladesh (Hossain et al. 2019).
Sedimentation occurred in the continental passive margin (pre-Oligocene) to a remnant
ocean basin (beginning of Miocene) (Alam et al. 2003). Basement rocks below the
Phanerozoic sedimentary cover compositionally ranges from tonalite, diorite to granite
(Ameen et al. 1998, 2007; Kabir et al. 2001; Hossain et al. 2007). U–Pb zircon
geochronology of basement rocks recovered from subcrop investigations near
Maddhapara, in the northeastern Bangladesh, reveals an intrusion ages of 1722 ± 6 Ma
from a tonalite suite (Ameen et al. 2007) and 1730 ± 11 Ma from a diorite suite
(Hossain et al. 2007). Limited geochemical data suggest ‘pre-plate collision’ affinity for
the tonalite and ‘syn-collisional’ affinity for granitic basements (Ameen et al. 2007)
whereas contemporaneous diorite are inferred to have a calc-alkaline origin associated
with a subduction zone settings (Hossain and Tsunogae 2014).
Towards the northeastern part of the Bengal basin, the Shillong-Meghalaya
Gneissic Complex (SMGC) comprises an ensemble of basement gneisses, meta-
morphosed to amphibolite-granulite grade, overlain by the Shillong Group of
sediments (Ahmed 1983; Nandy 2001). Different generations metamorphosed
granite and garnodiorites and porphyritic granitoids (Fig. 4), having a wide range of
isotopic age (ca. 1700–550 Ma), are reported from the terrane (Mazumdar 1976;
Ghosh et al. 1994; Bhattacharya and Ray Barman 2000; Ghosh et al. 2005; Kumar
et al. 2017). Lal et al. (1978) calculated P–T conditions of 750 °C and 5 kbar from
the metapelitic rocks near Sonapahar, eastern SMGC, whereas Chatterjee et al.
(2007) determined a higher P–T of 850 °C and 7.5 kbar associated with a
counter-clockwise P–T–t path. Chemical monazite dating constrains the age of
the metamorphism at *1596 ± 15 Ma whereas in the western part near
Garo-Goalpara, the dominant metamorphism is dated to be 500 ± 14 Ma
(Chatterjee et al. 2007). Mitra (1998) determined Pb–Pb ages of 1550–1530 Ma
from the detrital zircons from the supracrustals. In a recent publication Kumar et al.
(2017) documented that granitoids from Rongjeng and Guwahati (Assam) intruded
the Neo-Archean crust at 1778 ± 37 and 1630 ± 16 Ma respectively. Detailed
geochemical studies reveal dominant crustal source and intrusion of the granitoids
during syn-collisional settings.
Lithological and geochronological similarities indicate that the SMGC has been
a part of the Indian Shield (Evans 1964; Crawford 1974) or extension of CGGC
(Desikachar 1974; Chatterjee et al. 2007). Considering the distinct geochronolog-
ical similarity, it is also suggested that basement of the Bengal basin is the
extension of the Paleoproterozoic crust of CGGC and CITZ (Hossain et al. 2007;
Hossain and Tsunogae 2014). All these inferences further confirms the idea of
Paleoproterozoic crust o CGGC being extended up to SMGC as the Garo-Rajmahal
gap represents a shallow basement ridge (Desikachar 1974).
Proterozoic Crustal Evolution of the Chotanagpur Granite … 33
The contact between the Paleoproterozoic CGGC and Archean Singhbhum craton is
demarcated by the North Singhbhum Fold belt (NSFB), a 200 km long and 50 km
wide curvilinear orogenic belt. Extant geological information suggest that the NSFB
represents Proterozoic rift-related sedimentation along major suture zones (Sarkar
1982; Sarkar and Saha 1983; Gupta and Basu 2000; Sarkar 2000; Bhattacharya and
Mahapatra 2008; Bhattacharya et al. 2015). The Dalma Ophiolite Belt (DOB), an
ensemble of dominantly mafic- ultramafic rocks with minor felsic components
representing bimodal mafic-felsic volcanics, forms the most conspicuous part of
the NSFB. Geochemical evidences suggest that the magmatism occurred at
1619 ± 38 Ma (Rb–Sr whole rocks isochron age: Roy et al. 2002) under an exten-
sional regime, presumably in a back-arc like settings (Chakrabarti 1985; Bose et al.
1989). Metasedimentary rocks of the NSFB have been deformed-metamorphosed in
at least three tectonothermal events (Mahato et al. 2008) representing a greenschist
facies that grades into amphibolite facies at the contact of the CGGC (Bose 1954; Ray
and Gangopadhyay 1971; Mahato et al. 2008). Available geochronological study
from the southern part of the NSFB (S-NSFB) reveals that the metamorphism
occurred between 1.72 and 1.55 Ga in the southern part of the Dalma Volcanics
(Mahato et al. 2008; Chatterjee et al. 2010; Rekha et al. 2011). In the northern part
(N-NSFB) geochronological data are scarce and reveals an Early Neoproterozoic
tectonothermal event (*0.95 Ga; Rekha et al. 2011).
Discordant structural analyses, mismatching metamorphic grade and distinctly
different geochronological records suggest that Singhbhum Craton and CGGC
evolved separately in distant geological time and got amalgamated afterwards. The
contact between the Singhbhum Craton and the CGGC has been severely debated
and scarcity of geochronological and petrological data further blurs the nature and
the timing of the amalgamation and in this perspective the NSFB plays an important
role in the formation of the East Indian Shield (EIS). Several tectonic models have
been proposed to construct the evolutionary history of the EIS (Mahadevan 2002),
which mostly revolves around two major ideas: (1) subduction of the Singhbhum
Craton beneath the CGGC followed by collision between two continents (Sarkar
and Saha 1962, 1977; Sarkar 1982; Mahato et al. 2008; Rekha et al. 2011). and
(2) plume-driven basin formation accompanied by mafic magmatism and crustal
shortening (Mukhopadhyay 1990; Gupta and Basu 2000). Former model (Sarkar
and Saha 1962, 1977; Sarkar 1982; Mahato et al. 2008; Rekha et al. 2011) con-
siders the Singhbhum Craton and the CGGC as separate blocks that collided
together and the NSFB represented a milieu of arc magmatism, sedimentation and
obducted oceanic crust. Latter model indicates the presence of a unified crustal
section suggesting a prolonged connection between the Singhbhum Craton and the
CGGC where NSFB represented basin formed via incipient rifting. In a recent
publication Bhattacharya et al. (2015) argued that different sedimentary basins and
magmatic bodies in the NSFB did not evolve synchronously suggesting that the
34 S. Mukherjee et al.
juxtaposition of the CGGC and the Singhbhum Craton is not straight forwards and
warrants further geochronological and petrological investigations.
Late Paleoproterozoic history of the Earth has been largely influenced by the
presence of Columbia Supercontinent that was inferred to be amalgamated between
2.1 and 1.8 Ga (Rogers and Santosh 2009). However, several recent publications
argued for more prolonged amalgamation and placed the timing of final packing at
ca. 1450 Ma (Sarkar and Schenk 2016; Meert and Santosh 2017). The configuration
of the supercontinent is also debated owing to the scarcity of adequate petrological,
geochronological and paleomagnetic data. Abundance of Late Paleoproterozoic
magmatism and metamorphism (Figs. 5 and 6) and available paleomagnetic data
indicate that Indian plate played an important role during the formation of
Columbia supercontinent (Rogers and Santosh 2002; Zhao et al. 2002; Zhang et al.
2012a). Apart from its position in the supercontinent assembly, the configuration of
Greater Indian Landmass raised considerable debate as the timing of the CITZ,
through which different Archean-Early Paleoproterozoic cratons of India (Yedekar
et al. 1990; Acharyya 2003; Roy and Prasad 2003) amalgamated remained unclear.
Initial notion of a Late Paleoproterozoic-Mesoproterozoic (*1.8–1.5 Ga) suturing
(Acharyya 2003; Roy and Prasad 2003; Bhandari et al. 2011) has been challenged
by the proposition of Neoproterozoic (1.0–0.95 Ga) assembly of the CITZ
(Chattopadhyay and Khasdeo 2011; Bhowmik et al. 2012). However, high-grade
metamorphic events at ca. 1800–1600 Ma reported from the northern and southern
part of CITZ (Bhandari et al. 2011; Deshmukh et al. 2017) and arc magmatism at
ca. 1750 Ma (Bora et al. 2013) points towards an accretion orogen.
In the Ongle domain (Andhra Pradesh) of Eastern Ghats Mobile Belt (EGMB)
arc magmatism occurred at 1750–1710 Ma, followed by an UHT metamorphism at
*1610 Ma (Sarkar and Schenk 2014; Sarkar et al. 2015). Calc-alkaline magmatism
of arc-affinity with intrusion age of 1791–1771 Ma is also reported from Vinjamuru
domain of the EGMB (Vadlamani et al. 2013). Paleoproterozoic (*1860–
1810 Ma) arc magmatism is also reported from Aravalli-Delhi Mobile Belt (Kaur
et al. 2009, 2017a).
Although geochronological data are scarce from the vast terrain of the CGGC,
Paleoproterozoic dates, corresponding to tectonothermal and magmatic events, are
reported from different parts of the terrain (Fig. 6). Dey et al. (2017) and Dey et al.
(under review, a) reported Late Paleoproterozoic (*1650 Ma) UHT metamorphism
from the northeastern part of the terrane and the inferred P–T path corroborate
with convergent margin settings. Saikia et al. (2017) reported Paleoproterozoic
volcano-sedimentary sequence from the northern part of the terrain and the asso-
ciated felsic magmatism (ca. 1750–1660 Ma) resembles an arc settings. Evidences
of such tectonothermal events from parts of CGGC indicate its involvement in the
formation of the Greater Indian landmass during the assembly of Columbia
supercontinent.
The Mesoproterozoic extensional events throughout the world, manifested by
rift-related magmatism, deemed to be linked with the breakdown of the super-
continent (Rogers and Santosh 2002). Contemporaneous rift-related magmatism,
manifested by several ferroan granite and alkaline batholiths (Vinukonda granite:
1589 Ma: Dobmeier et al. 2006; Elchuru alkaline rocks: 1442 ± 30 Ma; Upadhyay
et al. 2006; Errakonda and Uppalapadu alkaline rocks: 1352 ± 2 Ma; Vijaya
Kumar et al. 2007) has been widely reported from western margin of the Eastern
Ghats belt which reflect the extensional events. Despite their restricted occurrence,
extension-related Mesoproterozoic (ca. 1450 Ma) anorogenic granites have intru-
ded the Paleoproterozoic crust in the northeastern part of CGGC (Mukherjee et al.
2017a, 2018). The multiple tectonothermal events (*1.75–1.3 Ga) that are recor-
ded in the EGMB and CGGC are also recorded from the fragments of the erstwhile
Columbia Supercontinents (Australia: Hand et al. 2007; Cutts et al. 2013; East
Antarctica: Kelly et al. 2002; Halpin et al. 2007b; Namibia: Kroner et al. 2010;
Scotland : Friend and Kinny 2001). This observation points to the fact that the
Indian shield was a part of the Columbia supercontinent.
Configuration and position of the Greater Indian landmass during the formation of
Rodinia supercontinent remained highly debatable (e.g. Bhowmik et al. 2012).
Widely inferred notion places the Greater Indian landmass adjacent to Australia and
East Antarctica along the marginal part of the Rodinia supercontinent (Dalziel
1991; Hoffman 1991; Torsvik et al. 1996; Weil et al. 1998; Dasgupta and Sengupta
2003; Li et al. 2008). However, the view has been opposed by several workers
(Fitzsimons 2000; Powell et al. 2001; Torsvik et al. 2001; Powell and Pisarevsky
2002; Merdith et al. 2017) based on available Paleomagnetic data. Numerous
studies have shown that Early Neoproterozoic (ca. 1000–900 Ma) tectonothermal
Proterozoic Crustal Evolution of the Chotanagpur Granite … 37
events experienced by the EGMB and the Rayner Complex of East Antarctica are
very much similar, which lends the idea of the Greater Indian landmass being
connected to Antarctica (Dasgupta and Sengupta 2003). Neoproterozoic
tectonothermal events presented in this review is consistent with evolution of the
CGGC in a continent-continent collision zone (Chatterjee and Ghose 2011;
Karmakar et al. 2011; Sanyal and Sengupta 2012). The NSFB that is located to the
south of the CGGC does not record strong impress of *1000–950 Ma
tectonothermal event (reviewed in Chatterjee et al. 2010). The pervasive planar
fabric in most part of the CGGC is E-W (Mahadevan 2002). This feature and the
structural attributes supporting that the NSFB is overridden by the CGGC (re-
viewed in Mahadevan 2002). This deduction by implication points to the fact that
the overthrusted unit possibly came from the north. The CGGC is surrounded by
Phanerozoic sedimentary cover on its three sides (Fig. 2a). For this reason the
nature of northern continent could not be studied.
Grenvillian tectonothermal events similar to CGGC are also recorded from the
EGMB and the entire Rayner Complex of east Antarctica (reviewed in Mukherjee
et al. 2017a; Li et al. 2008). Neoproterozoic clockwise P–T paths recovered from
different parts of the CITZ further corroborates the reworking of the
Paleoproterozoic mobile belts of India and amalgamation of the Greater Indian
landmass (Fig. 7). However, geochronological studies have shown that although
most of the cratons were amalgamated by *1000 Ma, India was not attached to the
supercontinent assembly until * 950 to 900 Ma (Li et al. 2008).
Timing of the break-up of Rodinia supercontinent has been controversial and
mostly placed at ca. 825 Ma, as manifested by widespread plume-related mag-
matism throughout the world (reviewed in Li et al. 2008). However, late
Tonian-Cryogenian (ca. 900–800 Ma) subduction-related magmatic activity has
been reported from different peripheral parts of Rodinia (reviewed in Cawood et al.
2016). Post Grenvillian (<950 Ma) tectonothermal activity, reported from different
parts of the CGGC, EGMB and East Antarctica suggest that the final assembly and
breakup of the Rodinia was prolonged in these parts. Contemporaneous (ca. 870–
780 Ma) high grade metamorphism event and shearing associated with the devel-
opment of EITZ further poses the possibility that fragmentation of amalgamated
East Antarctica and Greater Indian landmass did not initiate until ca. 780 Ma
(Chatterjee et al. 2010).
(A) Intra- and inter domain correlation of the magmatic, metamorphic and tectonic
history is of paramount importance to understand the evolution of the CGGC.
As mentioned above, the present data base are meagre (and imprecise too) to do
this correlation. Strain patterns can change with space and metamorphic
intensity may vary from place to place. Consequently, geometry of structure
and estimated P–T values may vary from place to place. This may lead to a
wrong interpretation about the tectonic evolution of the CGGC. It is not clear if
the contrasting metamorphic and structural information that are reported from
different parts of the CGGC reflects imprecision of the data base or they are
indicative of special/temporal variation of strain pattern and metamorphic
intensity. This problem can be approached by the following ways:
Proterozoic Crustal Evolution of the Chotanagpur Granite … 39
of basaltic compositions, yet some contact effects are expected in view of the
large volume of the felsic magma in the BMB.
(G) Felsic orthogneisses constitute the major component of the Domain I. Existing
petrological and geochronological information is available from a few scattered
areas. It is not clear if the magmatic protoliths of the entire felsic orthogneisses
were emplaced within a short time during crustal extension (Mukherjee et al.
2018). Or else, different pulses of felsic magmatism were emplaced in diverse
tectonic setting over a protracted period of time.
(H) The relation between the CGGC (Domain I) and the NSFB require to be studied
in great detail. In view of the last high grade event at 1000–950 Ma in Domain I
and its absence in NSFB, suturing of the two unit during Paleo-
Mesoproterozoic time seems implausible. Timing and mechanism of juxtapo-
sition between CGGC and NSFB are to be studied.
Acknowledgements S.M. and A.D. acknowledge the financial support in the form of research
fellowships from the University Grant Commission, New Delhi and Council of Scientific and
Industrial Research, New Delhi respectively. P.S. and S.S. acknowledge the grants received from
the programs awarded to the Department of Geological Sciences, Jadavpur University: University
Potential for Excellence (UPE-Phase II), Promotion of University Research and Scientific
Excellence and Fund for Improvement of Science and Technology (FIST-Phase II) from
Department of Science and Technology and Center of Advance Studies (CAS-phase VI).We
express our sincere thanks to Soumyajit Mukherjee for providing very constructive detailed
review. Mukherjee (2019) summarizes this work.
References
Acharyya A, Roy S, Chaudhuri BK et al (2006) Proterozoic rock suites along South Purulia Shear
Zone, Eastern India: evidence for rift related setting. Journal of Geological Society of India 68,
1069–1086
Acharyya SK (2003) The nature of mesoproterozoic Central indian tectonic zone with exhumed
and reworked older granulites. Gondwana Research 6, 197–214
Ahmed M (1983) Depositional environment of the basal conglomerate of the Barapani Formation,
Shillong Group, Khasi Hills, Meghalaya. Quarternary Journal Geological, Mining, and
Metallurgical Society of India 55, 62–68
Alam M, Alam MM, Curray JR et al (2003) An overview of the sedimentary geology of the
Bengal Basin in relation to the regional tectonic framework and basin-fill history. Sedimentary
Geology 155, 179–208
Ameen SMM, Khan MSH, Akon E, Kazi AI (1998) Petrography and major oxide chemistry of
some Precambrian crystalline rocks from Maddhapara, Dinajpur. Bangladesh Geoscience
Journal 4, 1–19
Ameen SMM, Wilde SA, Kabir Z et al (2007) Paleoproterozoic granitoids in the basement of
Bangladesh : a piece of the Indian shield or an exotic fragment of the Gondwana jigsaw ?
Gondwana Research 12, 380–387
Asami M, Suzuki K, Grew ES (2002) Chemical Th-U-total Pb dating by electron microprobe
analysis of monazite, xenotime and zircon from the Archean Napier complex, East Antarctica:
evidence for ultra-high-temperature metamorphism at 2400 Ma. Precambrian Research 114,
249–275
Proterozoic Crustal Evolution of the Chotanagpur Granite … 41
Baidya TK, Chakravorty PS, Drubetskoy E, Khiltova VJ (1987) New geochronologic data on
some granitic phases of the Chotanagpur granite gneiss complex in the northwestern Purulia
dist., West Bengal. Indian Journal of Earth Science 14, 136–141
Baidya TK, Maity N, Biswas P (1989) Tectonic phases and crustal evolution in a part of the
Eastern Chotanagpur Gneissic Complex. Journal of Geological Society of India 34, 318–324
Banerji AK (1985) On the nature of a part of boundary between Chotanagpur plateau and
Singhbhum orogenic belt and its role in mineralization. Bulletin of the Geology, Mining and
Metallurgical Society of India 53, 171–180
Basu SK (1993) Alkaline-carbonatite complex in Precambrian of South Purulia Shear Zone,
Eastern India: its characteristics and mineral potentialities. Indian Minerals 47, 179–194
Basu A, Patranabis-Deb S, Schieber J, Dhang PC (2008) Stratigraphic position of the *1000 Ma
Sukhda Tuff (Chhattisgarh Supergroup, India) and the 500 Ma question. Precambrian Research
167, 383–388
Bhandari A, Pant NC, Bhowmik SK, Goswami-Banerjee S (2011) 1.6 Ga ultrahigh‐temperature
granulite metamorphism in the Central Indian Tectonic Zone_insights from metamorphic
reaction history, geothermobarometry and monazite chemical ages. Geological Journal 46,
198–216
Bhattacharjee N, Ray J, Ganguly S, Saha A (2012) Mineralogical study of gabbro-anorthosite from
Dumka, Chhotanagpur Gneissic complex, Eastern Indian Shield. Journal of Geological Society
of India 80, 481–492
Bhattacharya BP, Ray Barman T (2000) Precambrians of Meghalaya: a concept. Proc Dr MS
Krishnan Birth Centen Sem Geol Surv, Ind Spec, Publ 55, 95–100
Bhattacharya DK, Mukherjee D, Barla VC (2010) Komatiite within Chhotanagpur Gneissic
Complex at Semra, Palamau district, Jharkhand: petrological and geochemical fingerprints.
Journal of Geological Society of India 76, 589–606
Bhattacharya HN, Mahapatra S (2008) Evolution of the Proterozoic rift margin sediments–North
Singhbhum Mobile Belt, Jharkhand-Orissa, India. Precambrian Research 162, 302–316
Bhattacharya HN, Nelson DR, Thern ER, Altermann W (2015) Petrogenesis and geochronology of
the Arkasani Granophyre and felsic Dalma volcanic rocks: implications for the evolution of the
Proterozoic North Singhbhum Mobile Belt, east India. Geological Magazine 152, 492–503
Bhattacharyya HN, Chatterjee A, Chowdhury S (1992) Tourmalinite from Cu–U belt of
Singhbhum, Bihar, India. Journal of Geological Society of India 39, 191–195
Bhattacharyya PK, Mukherjee S (1987) Granulites in and around the Bengal anorthosite, eastern
India: genesis of coronal garnet and evolution of the granulite–anorthosite complex. Geological
Magazine 124, 21–32
Bhowmik SK (2006) Ultra high temperature-metamorphism and its significance in the Central
Indian Tectonic Zone. Lihos 92, 484–505
Bhowmik SK, Bernhardt HJ, Dasgupta S (2010) Grenvillian age high-pressure upper
amphibolite-granulite metamorphism in the Aravalli-Delhi Mobile Belt, Northwestern India:
new evidence from monazite chemical age and its implication. Precambrian Research 178,
168–184
Bhowmik SK, Pal T, Roy A, Pant NC (1999) Evidence for pre-Grenvillian high-pressure granulite
metamorphism from the northern margin of the Sausar mobile belt in Central India. Journal of
Geological Society of India 53, 385–399
Bhowmik SK, Roy A (2003) Garnetiferous metabasites from the Sausar Mobile Belt: petrology,
P-T path and implications for the tectonothermal evolution of the Central Indian Tectonic
Zone. Journal of Petrology 44, 387–420
Bhowmik SK, Sarbadhikari AB, Spiering B, Raith MM (2005) Mesoproterozoic reworking of
Palaeoproterozoic ultrahigh-temperature granulites in the Central Indian Tectonic Zone and its
implications. Journal of Petrology 46, 1085–1119
Bhowmik SK, Spiering B (2004) Constraining the prograde and retrograde P-T paths of granulites
using decomposition of initially zoned garnets : an example from the Central Indian Tectonic
Zone. Contributions to Mineralogy and Petrology 147, 581–603
42 S. Mukherjee et al.
Humboldt Kolleg “Earth and material sciences for sustainable societal developments” (13–15
Jan), Raichak, India. pp 50–51
Das S, Sanyal S, Karmakar S, Sengupta P (2016) Petrology of a suite of deformed alkaline rocks in
Kankarkiari area of Purulia district, West Bengal: evidence of intra-continental suturing in parts
of Eastern Indian Shield during Neoproterozoic time. In: International conference on electron
microscopy & XXXVII annual meeting of EMSI (2–4 June), Varanasi, India, Abstract
Volume, pp 85–86
Dasgupta S, Narula PL, Acharyya SK, Banerjee J (2000) Seismotectonic atlas of India and its
environs. In: Narula PL et al. (ed), Geological survey of India. Geological Survey of India
Dasgupta S, Sengupta P (2003) Indo-Antarctic correlation: a perspective from the Eastern Ghats
Granulite Belt, India. Geological Society London Special Publications 206, 131–143
Deb M, Thorpe R, Krstic D (2002) Hindoli group of rocks in the Eastern Fringe of the
Aravalli-Delhi Orogenic Belt-Archean secondary Greenstone Belt or Proterozoic Supracrus
tals? Gondwana Research 5, 879–883
Decelles PG, Gehrels GE, Quade J et al (2000) Tectonic implications of of U-Pb U-Pb Zircon ages
of of the Himalayan orogenic belt in Nepal. Science 288, 497–499
Deshmukh T, Prabhakar N, Bhattacharya A, Madhavan K (2017) Late Paleoproterozoic clockwise
P–T history in the Mahakoshal Belt, Central Indian Tectonic zone: implications for Columbia
supercontinent assembly. Precambrian Research 298, 56–78
Desikachar SV (1974) A review of the tectonic and geological history of eastern India in terms of
‘plate tectonics’ theory. Journal of Geological Society of India 15, 137–149
Dey A, Karmakar S, Mauricio Ibanez-Mejia SM et al (under review, a) Petrology and
geochronology of metapelitic enclaves within felsic orthogneiss from the NE Chotanagpur
Granite Gneissic Complex, eastern India: evidence for Stenian-Tonian reworking of a late
Paleoproterozoic crust. Precambrian Research
Dey A, Karmakar S, Mukherjee S et al (under review, b) High Pressure metamorphism of mafic
granulites from the Chotanagpur Granite Gneissic Complex, India: evidence for collisional
tectonics during assembly of Rodinia. Journal of Geodynnamics
Dey A, Mukherjee S, Sanyal S et al (2017) Deciphering sedimentary provenance and timing of
sedimentation from a suite of metapelites from the Chotanagpur Granite Gneissic complex,
India: implications for Proterozoic tectonics in the East-Central part of the Indian shield. In:
Mazumder R (ed) Sediment Provenance. Influences on compositional change from source to
sink, Elsevier Ltd, pp 453–486
Dharma Rao CV, Santosh M, Dong Y (2012) U-Pb zircon chronology of the Pangidi-Kondapalle
layered intrusion, Eastern Ghats belt, India: constraints on Mesoproterozoic arc magmatism in
a convergent margin setting. Journal of Asian Earth Sciences 49, 362–375
Dharma Rao CV, Santosh M, Kim SW, Li S (2013) Arc magmatism in the Delhi Fold Belt:
SHRIMP U-Pb zircon ages of granitoids and implications for Neoproterozoic convergent
margin tectonics in NW India. Journal of Asian Earth Sciences 78, 83–99
Dharma Rao CV, Santosh M, Purohit R et al (2011) LA-ICP-MS U-Pb zircon age constraints on
the Paleoproterozoic and Neoarchean history of the Sandmata Complex in Rajasthan within the
NW Indian Plate. Journal of Asian Earth Sciences 42, 286–305
DiPietro JA, Isachsen CE (2001) U-Pb zircon ages from the Indian plate in northwest Pakistan and
their significance to Himalayan and pre-Himalayan geologic history. Tectonics 20, 510
Dobmeier C, Lütke S, Hammerschmidt K, Mezger K (2006) Emplacement and deformation of the
Vinukonda meta-granite (Eastern Ghats, India)—Implications for the geological evolution of
peninsular India and for Rodinia reconstructions. Precambrian Research 146, 165–178
Dobmeier C, Simmat R (2002) Post-Grenvillean transpression in the Chilka Lake area, Eastern
Ghats Belt—implications for the geological evolution of peninsular India. Precambrian
Research 113, 243–268
Dobmeier CJ, Raith MM (2003) Crustal architecture and evolution of the Eastern Ghats Belt and
adjacent regions of India. Geological Society London Special Publications 206, 145–168
Dunn JA (1929) The geology of north Singhbhum including parts of Ranchi and Singhbhum
districts. Memoirs of the Geological Survey of India 54, 1–280
Proterozoic Crustal Evolution of the Chotanagpur Granite … 45
Hoffman PF (1991) Did the breakout of Laurentia turn Gondwanaland inside-out ? Hoffman PF,
Published by : American Association for the Advancement of Science
Hoffmann P (1989) Speculations on Laurentia’s first gigayear (2.0–1.0 Ga). Geology 17, 135–138
Hossain I, Tsunogae T (2014) Crystallization conditions and petrogenesis of the paleoproterozoic
basement rocks in Bangladesh : an evaluation of biotite and coexisting amphibole mineral
chemistry. Journal of Earth Sciemce 25, 87–97
Hossain I, Tsunogae T, Rajesh HM (2007) Palaeoproterozoic U – Pb SHRIMP zircon age from
basement rocks in Bangladesh : a possible remnant of the Columbia supercontinent. Competes
Rendus Geoscience 339, 979–986
Hossain Md S, Khan SH, Chowdhury KR, Abdullah R (2019) Synthesis of the tectonic and structural
elements of the Bengal Basin (and its surroundings). In: Mukherjee S. (Ed) Tectonics and
structural geology: Indian context. Springer International Publishing AG, Cham, pp 135–218.
ISBN 978-3-319-99340-9
Jain SC, Yedekar DB, Nair KKK (1991) Central Indian shear zone: a major Pre-Cambrian crustal
boundary. Journal of Geological Society of India 37, 521–531
Kabir MZ, Khalil RC, Akon E et al (2001) Petrogenetic study of Precambrian basement rocks from
Maddhapara, Dinajpur, Bangladesh. Bangladesh Geoscience Journal 7, 1–18
Karmakar S, Bose S, Sarbadhikari AB, Das K (2011) Evolution of granulite enclaves and
associated gneisses from Purulia, Chhotanagpur Granite Gneiss Complex, India: evidence for
990–940 Ma tectonothermal event(s) at the eastern India cratonic fringe zone. Journal of Asian
Earth Sciences 41, 69–88
Kaur P, Chaudhri N, Raczek I et al (2009) Record of 1.82 Ga Andean-type continental arc
magmatism in NE Rajasthan, India: insights from zircon and Sm-Nd ages, combined with
Nd-Sr isotope geochemistry. Gondwana Research 16, 56–71
Kaur P, Chaudhri N, Raczek I, et al (2011) Zircon ages of late Palaeoproterozoic (ca.
1.72-1.70 Ga) extension-related granitoids in NE Rajasthan, India: regional and tectonic
significance. Gondwana Research 19, 1040–1053
Kaur P, Zeh A, Chaudhri N (2017a) Palaeoproterozoic continental arc magmatism, and
Neoproterozoic metamorphism in the Aravalli-Delhi orogenic belt, NW India: new constraints
from in situ zircon U-Pb-Hf isotope systematics, monazite dating and whole-rock geochem-
istry. Journal of Asian Earth Sciences 136, 68–88
Kaur P, Zeh A, Chaudhri N, Eliyas N (2017b) Two distinct sources of 1.73–1.70 Ga A-type
granites from the northern Aravalli orogen, NW India: constraints from in situ zircon U-Pb
ages and Lu-Hf isotopes. Gondwana Research 49, 164–181
Kelly NM, Clarke GL, Fanning CM (2002) A two-stage evolution of the Neoproterozoic Rayner
structural episode: new U-Pb sensitive high resolution ion microprobe constraints from the
Oygarden group, Kemp Land, East Antarctica. Precambrian Res 116, 307–330
Kelly NM, Harley SL (2004) Orthopyroxene–corundum in Mg–Al-rich granulites from the
Oygarden Islands, east Antarctica. Journal of Petrolology 45, 1481–1512
Kohn MJ, Paul SK, Corrie SL (2010) The lower lesser Himalayan sequence : a Paleoproterozoic
arc on the northern margin of the Indian plate. GSA Bulletin 122, 323–335
Korhonen FJ, Clark C, Brown M et al (2013) How long-livedis ultrahigh temperature
(UHT) metamorphism? Constraints from zircon andmonazite geochronology in the Eastern
Ghats orogenic belt, India. Precambrian Research 234, 322–350
Kovach VP, Berezhnayal NG, Salnikoval EB, Kotovl AB (2001) The Western Charnockite zone
of the Eastern Ghats Belt, India—an Laurentia-Siberia connection revisited again : an overview
of U-Pb Zircon. 6–7
Krause O, Dobmeier C, Raith MM, Mezger K (2001) Age of emplacement of massif-type
anorthosites in the Eastern Ghats Belt, India: constraints from U-Pb zircon dating and structural
studies. Precambrian Research 109, 25–38
Krishna V, Prasad RN, Pandey UK, et al (1996) Rb-Sr geochronology of Chotanagpur
gneiss-granulite complex around Kailashnathgufa area, Raigarh district, MP. In: Proceedings
of seventh national symposium on mass spectrometry
Proterozoic Crustal Evolution of the Chotanagpur Granite … 47
Krishna V, Sastry D, Pandey BK, Sinha RP (2003) U-Pb and Pb-Pb ages on columbite-tantalite
minerals from pegmatites of Bihar Mica Belt, Jharkhand, India. In: ISMAS silver jubilee
symposium on mass spectrometry. V. 2: contributed papers
Kroner A, Rojas-Agramonte Y, Hegner E, et al (2010) SHRIMP zircon dating and Nd isotopic
systematics of Palaeoproterozoic migmatitic orthogneisses in the Epupa metamorphic complex
of northwestern Namibia. Precambrian Research 183, 50–69
Kumar MN, Das N, Dasgupta S (1978) Geology and mineralization along the Northern Shear
Zone, Purulia District, West Bengal—an up-to-date appraisal. Records of the Geological
Survey of India 133, 25–31
Kumar S, Rino V, Hayasaka Y et al (2017) Contribution of Columbia and Gondwana
Supercontinent assembly- and growth-related magmatism in the evolution of the Meghalaya
Plateau and the Mikir Hills, Northeast India: constraints from U-Pb SHRIMP zircon
geochronology and geochemistry. Lithos 277, 356–375
Lahiri G, Das S (1984) Petrology of the area east of Daltonganj, Palamau district, Bihar. Journal of
Geological Society of India 25, 490–504
Lal N, Saini HS, Nagpaul KK, K. Sharma K (1976) Tectonic and cooling history of the Bihar Mica
Belt, India, as revealed by fission-track analysis. Tectonophysics 34, 163–180
Lal RK, Ackermand D, Seifert F, Haldar SK (1978) Chemographic relationships in
sapphirine-bearing rocks from Sonapahar, Assam, India. Contributions to Mineral Petrol 67,
169–187
Larson KP, Kellett DA, Cottle JM et al (2016) Anatexis, cooling, and kinematics during
orogenesis: miocene development of the himalayan metamorphic core, east-central nepal.
Geosphere 12, 1575–1593
Li ZX, Bogdanova SV, Collins AS et al (2008) Assembly, configuration, and break-up history of
Rodinia: a synthesis. Precambrian Research 160, 179–210
Liao QA, Li DW, Lu L et al (2008) Paleoproterozoic granitic gneisses of the Dinggye and
LhagoiKangri areas from the higher and northern Himalaya, Tibet: geochronology and
implications. Science China, Section D Earth Science 51, 240–248
Liu X, Zhao Y, Chen H, Song B (2017) New zircon U – Pb and Hf – Nd isotopic constraints on the
timing of magmatism, sedimentation and metamorphism in the northern Prince Charles
Mountains, East Antarctica. Precambrian Research 299, 15–33
Liu X, Zhao Y, Hu J (2013) The c. 1000–900 Ma and c. 550 – 500 Ma tectonothermal events in
the Prince Charles Mountains—Prydz Bay region, East Antarctica, and their relations to
supercontinent evolution. In: Harley SL, Fitzsimons ICW, Zhao Y (eds) Antarctica and
supercontinent evolution. Geological Society, London, Special Publications, pp 95–112
Liu X, Zhao Y, Song B et al (2009) SHRIMP U-Pb zircon geochronology of high-grade rocks and
charnockites from the eastern Amery Ice Shelf and southwestern Prydz Bay, East Antarctica:
constraints on late Mesoproterozoic to Cambrian tectonothermal events related to supercon-
tinent assembly. Gondwana Research 16, 342–361
Long SP, McQuarrie N, Tobgay T et al (2008) Tectonostratigraphy of the lesser Himalaya of
Bhutan: deducing the paleostratigraphy of the northern Indian margin. In: AGU fall meeting
abstracts
Mahadevan TM (1992) Geological evolution of the Chhotanagpur gneissic complex in parts of
Purulia district. Indian Journal of Geology 64, 1–22
Mahadevan TM (2002) Geology of Bihar & Jharkhand. Geological Society of India, Bangalore
Mahato S, Goon S, Bhattacharya A et al (2008) Thermo-tectonic evolution of the North
Singhbhum Mobile Belt (eastern India): a view from the western part of the belt. Precambrian
Research 162, 102–127
Maji AK, Goon S, Bhattacharya A et al (2008) Proterozoic polyphase metamorphism in the
Chhotanagpur Gneissic Complex (India), and implication for trans-continental Gondwanaland
correlation. Precambrian Research 162, 385–402
Mallik AK (1993) Dating of the granite plutons in Bihar mica belt. Bihar Records of the
Geological Survey of India 126:27–29
48 S. Mukherjee et al.
Mallik AK, Gupta SN, Barman Ray T (1991) Dating of early Precambrian granite-greenstone
complex of the eastern Indian Precambrian shield with special reference to the Chotanagpur
granite gneiss complex. Records of Geological Survey of India 125, 20–21
Malone SJ et al (2008). Paleomagnetism and Detrital Zircon geochronology of the upper Vindhyan
sequence, Son Valley and Rajasthan, India: a ca. 1000 Ma Closure age for the Purana basins?
Precambrian Research 164, 137–159
Mandal S, Robinson DM, Kohn MJ et al (2016) Zircon U-Pb ages and Hf isotopes of the Askot
klippe, Kumaun, northwest India: implications for Paleoproterozoic tectonics, basin evolution
and associated metallogeny of the northern Indian cratonic margin. Tectonics 35, 965–982
Mandal P (2016) Shear-wave splitting in Eastern Indian Shield: detection of a Pan-African suture
separating Archean and Meso-Proterozoic terrains. Precambrian Research. Elsevier B.V. 275,
278–285
Manna SS, Sen SK (1974) Origin of garnet in basic granulites around Saltora, West Bengal, India.
Contributions to Mineralogy and Petrology 44, 195–218
Mazumdar SK (1976) A summary of the Precambrian Geology of Khasi Hills, Meghalaya.
Geologoical Survey of India Miscillineous Publications 23, 311–334
Mazumdar SK (1988) Crustal evolution of the Chotanagpur gneissic complex and the Mica Belt of
Bihar. In: Mukhopadhyay D (ed) Precambrian of the Eastern Indian shield. Geological Society
of India Memoir 8, 49–84
Meert JG, Pandit MK, Pradhan VR et al (2010) Precambrian crustal evolution of Peninsular India:
a 3.0 billion year odyssey. Journal of Asian Earth Sciences 39, 483–515
Meert JG, Santosh M (2017) The Columbia supercontinent revisited. Gondwana Research 50, 67–
83
Merdith AS, Collins AS, Williams SE et al (2017) A full-plate global reconstruction of the
Neoproterozoic. Gondwana Research 50, 84–134
Mezger K, Cosca MA (1999) The thermal history of the Eastern Ghats Belt (India) as revealed by
U-Pb and 40Ar/39Ar dating of metamorphic and magmatic minerals: implications for the
SWEAT correlation. Precambrian Research 94, 251–271
Miller C, Klotzli U, Frank W et al (2000) Proterozoic crustal evolution in the NW Himalaya
(India) as recorded by circa 1.80 Ga mafic and 1.84 Ga granitic magmatism. Precambrian
Research 103, 191–206
Misra S, Dey S (2002) Bihar Mica Belt plutons-an example of post-orogenic granite from eastern
Indian shield. Journal of Geological Society of India 59, 363–37
Mishra DC, Singh B, Tiwari VM, et al (2000) Two cases of continental collisions and related
tectonics during the Proterozoic period in India - insights from gravity modelling constrained
by seismic and magnetotelluric studies. Precambrian Research 99:149–169. https://doi.org/10.
1016/S0301-9268(99)00037-6
Mitra SK (1998) Structure, sulphide mineralization and age of the Shillong group of rocks,
Meghalaya. In: SK Krishnan centenary commemorative national seminar, p 118–119
Morrissey LJ, Hand M, Kelsey DE (2015) Multi-stage metamorphism in the Rayner—Eastern
Ghats Terrane : P–T–t constraints from the northern Prince Charles Mountains, east Antarctica.
Precambrian Research 267, 137–163
Mukherjee D, Ghose NC, Chatterjee N (2005) Crystallization history of a massif anorthosite in the
eastern Indian shield margin based on borehole lithology. Journal of Asian Earth Sciences 25,
77–94
Mukherjee S (In press) Kinematics of pure shear ductile deformation within rigid walls: New
analyses. In: Billi A, Fagereng A. (eds) Problems and solutions in structural geology and
tectonics. Series editor: Mukherjee S, Developments in structural geology and tectonics book
series. Elsevier. ISBN: 9780128140482. ISSN: 2542–9000
Mukherjee S (2013) Channel flow extrusion model to constrain dynamic viscosity and Prandtl
number of the Higher Himalayan Shear Zone. International Journal of Earth Sciences 102,
1811–1835
Proterozoic Crustal Evolution of the Chotanagpur Granite … 49
Pandit MK, Carter LM, Ashwal LD et al (2003) Age, petrogenesis and significance of 1 Ga
granitoids and related rocks from the Sendra area, Aravalli Craton, NW India. Journal of Asian
Earth Sciences 22, 363–381
Panigrahi MK, Bream BR, Misra KC, Naik RK (2004) Age of granitic activity associated with
copper-molybdenum mineralization at Malanjkhand, Central India. Mineralium Deposita 39,
670–677
Patel SC (2007) Vesuvianite-wollastonite-grossular-bearing calc-silicate rock near Tatapani,
Surguja district, Chhattisgarh. Journal of Earth System Science 116, 143–147
Piper JDA (2013) Continental velocity through Precambrian times: the link to magmatism, crustal
accretion and episodes of global cooling. Geoscience Frontiers 4, 7–36
Pisarevsky SA, Wingate MTD, Powell CM et al (2003) Models of Rodinia assembly and
fragmentation. Geological Society, London, Special Publications 206, 35–55
Powell CM, Jones DL, Pisarevsky S, Wingate MTD (2001) Palaeomagnetic constraints on the
position of the Kalahari craton in Rodinia. Precambrian Research 110, 33–46
Powell CM, Pisarevsky SA (2002) Late neoproterozoic assembly of East Gondwana. Geology 30,
3–6
Radhakrishna BP, Naqvi SM (1986) Precambrian continental crust of India and its evolution.
Journal of Geology 94, 145–166
Ramakrishnan M, Vaidyanadhan R (2008) Geology of India. Geological Society of India
Bangalore
Ray Barman T, Bishui PK (1994) Dating of Chotanagpur gneissic complex of eastern Indian
Precambrian shield. Records of Geological Survey of India 127, 25–27
Ray Barman T, Bishui PK, Mukhopadhyay K, Ray JN (1994) Rb-Sr geochronology of the
high-grade rocks from Purulia, West Bengal and Jamua-Dumka sector, Bihar. Indian Minerals
48, 45–60
Ray J, Saha A, Ganguly S et al (2011a) Geochemistry and petrogenesis of Neoproterozoic
Mylliem granitoids, Meghalaya Plateau, northeastern India. Journal of Earth System Science
120, 459–473
Ray S, Gangopadhyay PK (1971) Metamorphic belt of Singhbhum, Manbhum and Chottanagpur.
Journal of Geological Society of India 12, 286–294
Ray S, Sanyal S, Sengupta P (2011b) Mineralogical control on rheological inversion of a suite of
deformed Mafic Dykes from parts of the Chottanagpur Granite Gneiss complex of Eastern
India, 263–276
Reddy S, Clarke C, Mazumder R (2009) Temporal constraints on the evolution of the Singhbhum
Crustal Province from U-Pb SHRIMP data. In: Mazumder R, Saha D (eds) Paleoproterozoic
Supercontinents and global evolution international association for Gondwana research
conference series, 9 abstract volume. pp 17–18
Rekha S, Upadhyay D, Bhattacharya A et al (2011) Lithostructural and chronological constraints
for tectonic restoration of Proterozoic accretion in the Eastern Indian Precambrian shield.
Precambrian Research 187, 313–333
Richards A, Argles T, Harris N et al (2005) Himalayan architecture constrained by isotopic tracers
from clastic sediments. Earth and Planetary Science Letters 236, 773–796
Rickers K, Mezger K, Raith MM (2001) Evolution of the continental crust in the Proterozoic
Eastern Ghats Belt, India and new constraints for Rodinia reconstruction: implications from
Sm-Nd, Rb-Sr and Pb-Pb isotopes. Precambrian Research 112, 183–210
Rode KP (1948) On charnockitic rocks of Palamau, Bihar, India. Schweiz, Min Petr Mitt 28,
288–307
Rogers JJW (1996) A history of continents in the past 3 billion years. Journal of Geology 104,
91–107
Rogers JJW, Santosh M (2002) Configuration of Columbia, a Mesoproterozoic supercontinent.
Gondwana Research 5, 5–22
Rogers JJW, Santosh M (2009) Tectonics and surface effects of the supercontinent Columbia.
Gondwana Research 15, 373–380
Proterozoic Crustal Evolution of the Chotanagpur Granite … 51
Sarkar SN, Saha AK (1962) A revision of the Precambrian stratigraphy and tectonics of
Singhbhum and adjacent regions. Quarternary Journal Geological Mining and Metallurgical
Society of India 34, 97–136
Sarkar SN, Saha AK (1977) The present status of the Precambrian stratigraphy, tectonics and
geochronology of Singhbhum-Keonjhar-Mayurbhanj region, eastern India. Indian Journal of
Earth Sciences, S. R. Volume 37–65
Sarkar SN, Saha AK (1983) Structure and tectonics of the Singhbhum-Orissa iron ore craton,
eastern India. In: Roy S (ed) Structure and tectonics of Precambrian rocks in India. Hindustan
Publishing Corporation, New Delhi, Recent Researches in Geology 10, 1–25
Sarkar T, Schenk V (2014) Two-stage granulite formation in a Proterozoic magmatic arc (Ongole
domain of the Eastern Ghats Belt, India): part 1. Petrology and pressure–temperature evolution.
Precambrian Research 255, 485–509
Sarkar T, Schenk V (2016) Early mesoproterozoic (1.6–1.5 Ga) granulite facies events in the
Ongole domain: geodynamic significance and global correlation. Journal of Metamorphic
Geology 34, 765–784
Sarkar SN, Ghosh D, St Lambert RJ (1986) Rubidium-strontium and lead isotopic studies on the
soda-granites from Mosaboni, Singhbhum Copper Belt, E. India. Indian Society of Earth
Sciences 13:101–116
Sarkar T, Schenk V, Berndt J (2015) Formation and evolution of a Proterozoic magmatic arc:
geochemical and geochronological constraints from meta-igneous rocks of the Ongole domain,
Eastern Ghats Belt, India. Contributions to Mineralogy and Petrology 169, 1–27
Sarkar, SN(1980) Precambrian stratigraphy and geochronology of Peninsular India: a review.
Indian Journal of Earth Sciences 7, 12–26
Sen SK, Bhattacharya A (1993) Post-peakpressure–temperature–fluid history of the granulites
around Saltora, West Bengal. In: Proceedings of the National academy of sciences of India,
63(A), 282–306
Sengupta N, Mukhopadhyay D, Sengupta P, Radegund H (2005) Tourmaline-bearing rocks in the
Singhbhum shear zone, eastern India: Evidence of boron infiltration during regional
metamorphism. American Mineralogist 90, 1241
Sengupta P, Sen J, Dasgupta S et al (1999) Ultra-high temperature metamorphism of metapelitic
granulites from Kondapalle, Eastern Ghats Belt: implications for the Indo-Antarctic correlation.
Journal of Petrology 40, 1065–1087
Sengupta P, Raith MM, Kooijman E, Talukdar M, Chowdhury P, Sanyal S, Mezger K,
Mukhopadhyay D (2015) Provenance, timing of sedimentation and metamorphism of
metasedimentary rock suites from the Southern granulite terrane, India. In: Mazumder R,
Eriksson PG (eds) Precambrian basins of India: stratigraphic and tectonic context. Geological
Society, London, Memoirs 43, 297–308
Simmat R, Raith MM (2008) U-Th-Pb monazite geochronometry of the Eastern Ghats Belt, India:
timing and spatial disposition of poly-metamorphism. Precambrian Research 162, 16–39
Singh AP, Kumar N, Singh B (2004) Magmatic underplating beneath the Rajmahal Traps: gravity
signature and derived 3-D configuration. Journal of Earth System Science 113, 759–769
Singh RN, Thorpe R, Kristic D (2001) Galena Pb-isotope data of base metal occurrences in the
Hesatu-Belbathan belt, eastern Precambrian shield. Journal of Geological Society of India 57,
535–538
Singh S, Claesson S, Jain AK, et al (1994) Geochemistry of the Proterozoic peraluminous
granitoids from the Higher Himalayan Crystalline (HHC) and Jutogh Nappe, NW-Himalaya,
Himachal Pradesh, India. Journal of Nepal Geological Society 10, 125
Singh Y, Krishna V (2009) Rb–Sr geochronology and petrogenesis of granitoids from the
Chotanagpur granite gneiss complex of Raikera–Kunkuri region, Central India. Journal of
Geological Society of India 74, 200–208
Sinha AK, Bhattacharya DK (1995) Geochemistry of the magnetite deposits around Sua, Palamau
district, Bihar. Journal of Geological Society of India 46, 313–316
Sivaraman T V, Raval U (1995) U-Pb isotopic study of zircons from a few granitoids of
Delhi-Aravalli belt. Journal of Geological Society of India 46, 461–475
Proterozoic Crustal Evolution of the Chotanagpur Granite … 53
Som SK, Bandyopadhyay KC, Prasad RK (2007) Cesium enrichment and resource evaluation in
aplite and pegmatite at the Southern slope of cesium enrichment and resource evaluation in
aplite and pegmatite. Journal of Geological Society of India 70, 273–281
Soni S, Mukherjee AB, Sengupta DK (1991) A new fluorite deposit in the Palamau District, Bihar
and the associated iron-fluorine-tungsten skarns and hornfelses. Journal of Geological Society
of India 38, 504–510
Srivastava AK, Pandey KH, Kumar G (2000) Geochemical characteristics of the Jhirgadandi
granitoid, Sonbhadra district, Uttar Pradesh. In: Proceedings of national seminar on
tectonomagamatism, geochemistry and metamorphism of Precambrian terrains, pp 189–199
Srivastava SC, Ghose NC (1992) Petrology of the highgrade gneisses and granites around Chianki,
south of daltonganj, district Palamau, Bihar. Indian Journal of Geology 64, 122–142
Torsvik TH, Carter LM, Ashwal LD, Bhushan SK (2001) Rodinia refined or obscured :
palaeomagnetism of the Malani igneous suite (NW India). Precambrian Research 108, 319–333
Torsvik TH, Smethurst MA, Meert JG et al (1996) Continental break-up and collision in the
Neoproterozoic and Palaeozoic—a tale of Baltica and Laurentia. Earth-Science Reviews 40,
229–258
Treloar PJ, Rex DC (1990) Cooling and uplift histories of the crystalline thrust stack of the Indian
Plate internal zones west of Nanga Parbat, Pakistan Himalaya. Tectonophysics 180, 323–349
Tsunogae T, Dunkley DJ, Horie K et al (2014) Petrology and SHRIMP zircon geochronology of
granulites from Vesleknausen, Lützow-Holm complex, East Antarctica: Neoarchean magma-
tism and Neoproterozoic high-grade metamorphism. Geoscience Frontiers 5, 167–182
Tsunogae T, Yang QY, Santosh M (2016) Neoarchean–Early Paleoproterozoic and early
Neoproterozoic arc magmatism in the Lützow–Holm complex, East Antarctica: insights
from petrology, geochemistry, zircon U–Pb geochronology and Lu–Hf isotopes. Lithos 263,
239–256
Upadhyay D, Raith MM (2006) Intrusion age, geochemistry and metamorphic conditions of a
quartz-monzosyenite intrusion at the craton-Eastern Ghats Belt contact near Jojuru, India.
Gondwana Research 10, 267–276
Upadhyay D, Raith MM, Mezger K, Hammerschmidt K (2006) Mesoproterozoic rift-related
alkaline magmatism at Elchuru, Prakasam Alkaline Province, SE India. Lithos 89, 447–477
Upreti BN, Rai SM, Sakai H et al. (2003) Early Proterozoic granite of the Taplejung Window, far
eastern lesser Nepal Himalaya. Journal of Nepal Geological Society 28, 9–18
Vadlamani R, Kröner A, Vasudevan D et al (2013) Zircon evaporation ages and geochemistry of
metamorphosed volcanic rocks from the Vinjamuru domain, Krishna Province: evidence for
1.78 Ga convergent tectonics along the southeastern margin of the Eastern Dharwar Craton.
Geological Journal 48, 293–309
Vernon RH (2004) A practical guide to rock microstructure. 594 p. Cambridge University Press.
ISBN 0 521 89133 7
Vijaya Kumar K, Frost CD, Frost BR, Chamberlain KR (2007) The Chimakurti, Errakonda, and
Uppalapadu plutons, Eastern Ghats Belt, India: an unusual association of tholeiitic and alkaline
magmatism. Lithos 97, 30–57
Vinogradov A, Tugarinov AL, Zhykov C et al (1964) Geochronology of Indian Precambrian. In:
Report of the 22nd international congress, New Delhi. pp 553–567
Wani H, Mondal MEA (2016) Geochemical evidence for the Paleoproterozoic arc–Back arc basin
association and its importance in understanding the evolution of the Central Indian Tectonic
Zone. Tectonophysics 690, 318–335
Wanjari NR, Chaturvedi R, Mahanta DN (2012) Specialised thematic mapping in Munger–Rajgir
Group of rocks to examine structural and stratigraphic set up in and around Gaya–Rajgir areas
in parts of Gaya, Nawada and Jahanabad districts of Bihar
Weil AB, Van der Voo R, Mac Niocaill C, Meert JG (1998) The Proterozoic supercontinent
Rodinia: paleomagnetically derived reconstructions for 1100 to 800 Ma. Earth and Planetary
Science Letters 154, 13–24
Wiedenbeck M, Goswami JN, Roy AB (1996) An ion microprobe study of single zircons from the
Amet granite, Rajasthan. Journal of Geological Society of India 48, 127–137
54 S. Mukherjee et al.
Worsley TR, Moody JB, Nance RD (1985) Proterozoic to recent tectonic tuning of biogeochemical
cycles. Carbon cycle atmospheric CO natural variations archean to present, 561–572
Worsley TR, Nance RD, Moody JB (1986) Tectonic cycles and the history of the Earth’s
biogeochemical and paleoceanographic record. Paleoceanography 1, 233–263
Yadav BS, Wanjari N, Ahmad T, Chaturvedi R (2014) Geochemistry and petrogenesis of
proterozoic granitic rocks from northern margin of the chotanagpur gneissic complex (CGC).
Journal of Earth System Science 125, 1041–1060
Yedekar DB, Jain SC, Nair KKK, Dutta KK (1990) The Central Indian collision suture.
Precambrian of Central India. Geological Survey of India Special Publications 28, 1–37
Yin A, Dubey CS, Kelty TK et al (2010) Geologic correlation of the Himalayan orogen and Indian
craton: part 2. Structural geology, geochronology, and tectonic evolution of the eastern
Himalaya. Bulletin of Geological Society of America 122, 360–395
Zeitler PK, Sutter JF, Williams IS et al (1989) Geochronology and temperature history of the
Nanga Parbat–Haramosh massif, Pakistan. Geological Society of America Special Paper 232,
1–22
Zhang S, Li ZX, Evans DAD et al (2012a) Pre-Rodinia supercontinent Nuna shaping up: a global
synthesis with new paleomagnetic results from North China. Earth and Planetary Sciences
Letters 353–354, 145–155
Zhang SH, Zhao Y, Liu XC et al (2012b) U-Pb geochronology and geochemistry of the bedrocks
and moraine sediments from the Windmill islands: implications for Proterozoic evolution of
East Antarctica. Precambrian Research 206–207, 52–71
Zhao G, Cawood PA, Wilde AS, Sun M (2002) Review of global 2.1-1.8 Ga orogens: implications
for a pre-Rhodinia supercontinent. Earth-Science Reviews 59, 125–162
Geomorphic Characteristics
and Morphologic Dating of the Allah
Bund Fault Scarp, Great Rann
of Kachchh, Western India
1 Introduction
The 16 June 1819 Allah Bund earthquake is the largest known earthquake in
western India that produced spectacular geomorphological changes in its epicentral
area in the Great Rann of Kachchh (Lyell 1855). The earthquake though not
instrumentally recorded, is described in several historical accounts (MacMurdo
1824; Baker 1846; Wynne 1872; Oldham 1926). The earthquake produced
*90 km long E-W trending and south facing low fault scarp to its north, blocking
the distributary of the Indus river that flowed through the Kori creek into the
Arabian sea and caused subsidence of the area around Sindri, which was submerged
by a local tsunami. The scarp is located in the logistically challenging and largely
inaccessible saline terrain of Great Rann, which shows peculiar geomorphic char-
acters (Roy and Merh 1982; Merh 2005). This is the most likely reason for the fact
that very few field-based studies have been carried out on the Allah Bund Fault
scarp (Rajendran and Rajendran 2001; Thakkar et al. 2012). The Great Rann is a
tectonically formed basin bounded by the Kachchh Mainland Fault (KMF) in the
south and the Nagar Parkar Fault (NPF) in the north (Biswas 1987). Towards the
west the basin opens up to the Arabian Sea and the Indus delta region. The basin
formed a major sink for sediments during Pleistocene-Holocene that are dominantly
fine-grained (clayey silt to silty clay) and presumably sourced from the Himalaya
The original version of this chapter was revised: Missed out author corrections in figure captions
have been incorporated. The correction to this chapter is available at
https://doi.org/10.1007/978-3-319-99341-6_18
through the now extinct Vedic Saraswati river and the Indus delta and deposited in
shallow marine marginal gulf environment (Valdiya 2002; Maurya et al. 2013;
Khonde et al. 2017a, b). In this article, we describe the geomorphological charac-
teristics along the Allah Bund Fault scarp in detail and demonstrate the application of
morphological dating technique to determine the age of the scarp. We begin with a
detailed description of the enigmatic landscape characteristics of the Great Rann to
understand the geomorphic setting and evolution of the Allah Bund Fault scarp.
The major geomorphic components of the Great Rann are the almost flat and
hyper-saline surface, hereafter described as the Rann surface, several islands
(locally called ‘bet’) of different shapes and sizes and the * E-W Allah Bund scarp
(Fig. 2). Part of the extensive Rann surface is salt-encrusted while the remainder is
free of salt crust though the sediments are inherently saline. The most overbearing
part of the Rann is the flatness of the Rann surface that lies 2–6 m amsl. The Rann
surface consists of several large to small islands that remain above the submergence
level. There are several smaller islands rising up to 1–10 m above the Rann surface,
especially in the northern part of the Great Rann and consist of sediments resem-
bling the Rann surface raised to a higher level (Khonde et al. 2013). The top cover
of these islands is usually made up of 1–2 m thick aeolian sediment blown from the
wind-swept Rann surface.
The major factor responsible for the unique present day environmental condi-
tions of the Rann surface is its periodic submergence by sea water and annual
monsoon precipitation. The flat Rann surface and the negligible westward gradient
allows extensive inundation by sea water from the Arabian sea in the west and by
river waters from the northeast and south during the monsoon season (Glennie and
Evans 1976; Roy and Merh 1982). The submergence pattern of the Rann surface is
however not uniform as it shows very small variations in elevation which has
resulted in variable geomorphic characteristics. The Great Rann is therefore geo-
morphologically divisible into four geomorphic units (Fig. 2 and 3) the Banni plain,
Supra tidal salt flat, Inland saline flat and the Bet zone. Owing to imperceptible
gradient, the boundaries between the various geomorphic units are gradational.
The Banni plain is a vast flat terrain, highly vegetated with thorny shrubs and
grasses, and extending from the mainland Kachchh in the south and the Pachham
island in the north (Fig. 1a). It is a distinct geomorphic surface of the Great Rann
that occurs at the highest elevation and is consequently completely free of present
day marine influence (Figs. 2, 3 and 4a). The entire terrain of the Banni plain is
regarded as a vast raised mudflat (Kar 1995) that coincides with the subsurface
Geomorphic Characteristics and Morphologic Dating … 57
Fig. 1 a Digital elevation model (DEM) of Kachchh region prepared from SRTM data (http://
srtm.csi.cgiar.org). DEM prepared by Patidar (2010). Inset-Location map. b Digital Elevation
Model (DEM) of Allah Bund Fault Scarp prepared from SRTM data (http://srtm.csi.cgiar.org)
showing the various geomorphological characteristics. Profile location indicated is for profiles
given in Fig. 8
Median High. Based on variations in elevation, Kar (1995) has divided the Banni
plain into three sub-units-high level mudflat, an undifferentiated and sloping low
level mudflat and a residual depression.
This is a vast but linear and narrow E-W trending low lying zone with several
centimeters thick salt crust between the Banni plain in the south and the Bet zone to
the north (Fig. 2 and 4b–d). This zone occurs at the lowest elevation (*2 m) in the
west and gradually rises to *4 m towards east. Being at the lowest elevation, this
zone forms the main pathway through which the saline waters of the Arabian sea in
the west enter and spread out (Roy and Merh 1982) submerging about two-thirds of
the Rann surface to varying degrees depending upon the volume/magnitude of the
58 A. Padmalal et al.
Fig. 2 Satellite image of Great Rann of Kachchh basin showing the variations in surface
morphology. The image shows the completely dry Rann surface as it appeared in the extreme arid
season in May, 2003. Source www.earthobservatory.nasa.gov. The geomorphological divisions of
the Great Rann (1–4) are also indicated. 1-Banni plain, 2-Supra tidal salt flat, 3-Inland saline flats
and 4-Bet zone. (Pa-Pachham island, Kh-Khadir island, Be-Bela island, Ch-Chorar island)
Fig. 3 N-S topographic profile across the Great Rann basin showing the geomorphic divisions.
Vertical scale is highly exaggerated. The elevation data is based on the SOI topographical maps
(survey years-1960–66)
Fig. 4 a Photograph showing the typical nature of the surface of the Banni plain. Sand storms as c
seen in the picture are a common sight during the peak summers. b Photograph showing the
typical extensive flat surface of the supra tidal salt flat. The salt crust is the result of regular marine
inundation of the surface. c Close view of the large polygonal cracks in the salt crust. d Photograph
showing the thickness (*10 cm) of the salt crust. e View of the Inland saline flat to the north of
Bela island. The scarp in the background marks the geomorphic expression of the Island Belt Fault
(IBF). f View of the typical salt crust free surface of the Bet zone. g Northward view of the
developing gullies in the northern most part of the Bet zone. The northward upslope nature of the
surface is also clearly visible. h Surface of the Bet zone covered with numerous bivalve shells
which thrive during periods of submergence
Geomorphic Characteristics and Morphologic Dating … 59
60 A. Padmalal et al.
ingression resulting in a thick salt crust (Fig. 4d). Evaporation to dryness results in
several centimeteres thick, residual salt crusts, which characterize the surface of the
supra tidal salt flats (Fig. 4b–d).
This zone comprises the easternmost part of the Great Rann (Fig. 2 and 4e) that is
not influenced by marine submergence, but is inundated by monsoon precipitation
and by the rivers from the east and north. The elevation rises towards the margins of
the zone, giving it a shallow bowl like morphology and comprise inherently saline
Rann sediments (Roy and Merh 1982).
It comprises the flat Rann surface in the northwestern part of the Great Rann shows
several bets occurring few metres above the Rann surface (Fig. 4f–h). The Bet zone
is delimited by the Kuar and Bedia bets in the east and the supra tidal salt flat in the
south (Fig. 2). In addition to the bets with well defined margins, several elevated
parts within the Rann surface also exist which gradually rise above the Rann
surface. Many of these are comparable to the bets in terms of their size. Their
surfaces comprise aeolian sediments and generally support small vegetation like
scrubs and grasses of various size. Morphologically, the shapes of all bets and
almost all elevated tracts described above are elongated in N-S direction, which, on
a map gives the misleading impression of wide N-S trending channels separated by
bets. However, these are not ‘channels’ in any sense as they are basically flat Rann
surfaces several tens of kilometers wide comprising the inter-bet regions (Khonde
et al. 2013). The surface of the Bet zone shows a very gradual northward slope
away from the Allah Bund scarp which testifies to the uplift along the ABF (Fig. 3).
The western part of the Bet zone i.e. the area around Vigukot and further west
shows submergence characteristics which is slightly different from the rest of the
Bet zone and the Great Rann (Khonde et al. 2013). The 1819 Allah Bund earth-
quake caused drastic geomorphological changes along with subsequent flooding
events of the now defunct distributary of Indus River (Burnes 1835; Oldham 1926;
Bilham 1998). The now disconnected distributary of Indus—the Nara river/Pooran
flowed into this region and joined Arabian sea through the Kori creek in the south
before the earthquake. This part therefore formed the eastern margin of the Indus
Delta till the 1819 earthquake changed the morphology of the area. Presently also,
the region to the west and SW of Vigukot is prone to flooding frequently by river
floods from the north and relatively less frequently by sea water influx from the
south. In addition numerous small shallow (*0.5 m deep) channels of uncertain
affinity also exist. Overall, the role of rivers is evidently more pronounced in the
Geomorphic Characteristics and Morphologic Dating … 61
western most part than the rest of the Bet zone. Contrary to the rest of the Great
Rann, the Bet zone shows wide variation in elevation due to the presence of bets
and is characterized by several seasonal short distance channels, pools i.e. local
depressions and elevated surfaces (Khonde et al. 2013).
The E-W trending intrabasinal structural features within the Great Rann like the
Island Belt Fault (IBF), the Allah Bund Fault (ABF) and the subsurface Banni faults
correlate remarkably with the geomorphic divisions described above. A critical
evaluation of the elevation differences of the various geomorphic divisions of the
Great Rann shows a strong correlation with the above mentioned regional structural
elements worked out by Biswas (1987, 1993). All geomorphological units of the
Great Rann show prominent structural control (Maurya et al. 2013). The Banni
plain astrides the subsurface Median High, while the supra tidal salt flat and the
inner saline flats occupy a structural depression to the north of the subsurface
Bhirandiyala high and the Island Belt Fault (IBF). The supra tidal flats extends into
a depression to the south of the island belt between the Pachchh and Khadir islands,
the Wagad highland to the east and the mainland in the south. The Bet zone occurs
on the upthrown northern block of the roughly E-W trending Allah Bund Fault.
Based on the distinct geomorphic divisions attributed to variable submergence
pattern and their correlation with structural elements, it is inferred that the emer-
gence of the Rann surface may have occurred gradually due to differential tectonic
activity along the various subsurface intrabsinal faults in the recent past, which
formed the morphologic units viz. the Banni plain, the Bet zone and the supra tidal
salt flat (Maurya et al. 2013, 2016). The Banni plain astrides the subsurface Median
high and is separated from the remainder of the Rann basin by the Banni Fault to
the north (Biswas 1974). Similarly, the Bet zone is delimited by the Allah Bund
Fault to its south (Roy and Merh 1982). The close association of these units with
faults suggest differential tectonic activity along subsurface faults within the Great
Rann basin may have played a major role in the emergence of various morphologic
units at different times (Maurya et al. 2013). Based on elevation and present day
submergence characteristics, the Banni plain appears to be the first to emerge
followed by the Bet zone, the inner saline flat and the supra tidal salt flat, which still
gets submerged by marine waters regularly.
intensity report, Gutenberg and Richer (1954) assigned a magnitude of 8.4 to the
event. Later on Chandra (1977) and Johnston and Kanter (1990) estimated the
magnitude to be 7.8 and epicenter on 23.6° N and 69.6° E, while Quittmeyer and
Jacob (1979) estimated the epicenter at 24° N and 69° E. According to Bilham
(1998), the 1819 event was a near the surface (0–10 km depth) reverse-slip rupture
on a 90-km-long, 50° to 70° N-dipping fault plane which matches the measured
elevation changes from the event. To et al. (2004) estimated a 50-km-long rupture
dipping 45 to the north with 3–8 m slip.
Prior to the 1819 earthquake, the region was a flourishing trade route extending
northward into Sind through Kori creek and the distributary of the Indus (Grindlay
1808). The Sindri fort was a major halting point and a revenue collection centre
along this route. Extensive surface effects resulting due to the 1819 earthquake have
been documented in several historical records (MacMurdo 1824; Baker 1846;
Wynne 1872; Oldham 1926; Frere 1870. A detailed review of historical accounts
for this earthquake is available in Bilham (1998). The earthquake produced an
astonishing variety of surface effects that includes vertical movements of the
ground, flooding of regions near sea level, widespread liquefaction, and a local
tsunami and the complete damming of a distributary of the Indus river. Two sig-
nificant changes occurred in the region around Sindri fort at the western part of
Great Rann. The foundation of fort and surrounding Rann surface subsided by more
than 1 m and the region 7 km north of fort got elevated to 3–6 m, shutting out
northward navigation into the Sind province due to blocking of the distributary of
Indus river. The northern region (the present Bet zone) was uplifted forming a
10–15 ft high south facing scarp which caused the complete damming of the river.
The natural dam so formed was named by local people as the Allah Bund (dam of
God). The scarp presently marks the sharp boundary between the Bet zone in the
north and the active tidal flats of Kori creek and the Bet zone Notably all the above
surface changes documented in the available historical records are from the western
part of the Great Rann of Kachchh. However, it is also mentioned that several
temporarily active rivers with subsequent sand venting caused transient flooding of
major part of Rann surface (Baker 1846; Oldham 1926).
The Allah Bund Fault scarp is an elongated steeply dipping south facing scarp
bordered on the south by the salt encrusted surface of Rann. The Bet zone com-
prises the flat Rann surface in the northwestern part of the Great Rann shows
several bets occurring few metres above the Rann surface. The Bet zone is
delimited by the Kuar and Bedia bets in the east and the supra tidal salt flat in the
south. To the north lie the sand ridges of Sind (Pakistan). Towards the west, the Bet
zone imperceptibly merges with the Indus delta. Historical documents suggest that
the south facing fault scarp was produced during the 1819 earthquake, causing
uplift of the northern part of the Great Rann (Burnes 1835; Oldham 1926),
Geomorphic Characteristics and Morphologic Dating … 63
Fig. 5 Topographic profile drawn over the crest of the Allah Bund scarp. The supra tidal salt flat
surface is also shown to indicate the height of the scarp. Vertical scale is highly exaggerated. The
elevation data is based on the Survey of India topographical maps (survey years-1960–66)
which corresponds with the present day Bet zone. The scarp trends E-W and
laterally extends for *90 kms (Oldham 1926). However, presently the scarp is
visibly identifiable for about *60 kms length in the western and central part of the
extension as per Oldham (1926). In the westernmost part, the scarp rises gently
northward rise of the ground above the intertidal flats of the Kori creek (Figs. 5 and
6). Further east, the scarp gains elevation and therefore is more distinctly identi-
fiable (Fig. 5). The scarp height continues to increase eastward with the highest
elevation recorded *4–5 m from the supra tidal salt flat in the central part (Figs. 5
and 6). Beyond this the scarp again gradually reduces in height and finally disap-
pears in the flatness of the Rann surface to the southeast of Shakti bet. As per
Oldham (1926), the scarp continued eastward where it is presumably represented by
the southern cliffy margins of the Gainda bet, Mori bet and the Kuar bet. Sediments
of the bund consist of thinly laminated alternate layers of dark brown silt and clays
encrusted with salt and ferruginous tubules; these differ from the land-derived
fluvial bet sediments that impinge against the bund in the north (Rajendran and
Rajendran 2001).
At places, the height of the scarp is accentuated by the deposition of 1–2 m thick
wind-blown saline silty sediments. This especially observed in the central part
where the scarp attains highest elevation. A major significant characteristic that
defines the scarp as an erosional scarp is its deeply gullied nature (Thakkar et al.
2012). The gullies are 1–3 m deep and usually form a dendritic pattern over the
crest of the scarp. Another significant characteristic feature of the scarp is the
subvertical northward face of the scarp surface, which finally merges with the flat
Rann surface. The northward slope developed over the scarp is attributed to the
back tilting of the Rann surface due to upliftment along the scarp. The backtilting
being subtle is not recognisable in the field at many places as the elevation drop is
marginal (1–2 m) that occurs over the distance of few kilometers (Fig. 6). The back
tilted surface over the scarp suggests that the scarp is tectonically formed and is in
sharp contrast to the flat surface of the Great Rann.
64 A. Padmalal et al.
Fig. 6 Topographic cross sections drawn across the E-W trending Allah Bund scarp. The top
profile is from the western extremity while the bottom one is from the eastern extremity of the
scarp. Vertical scale is highly exaggerated. The elevation data is based on the SOI topographical
maps (survey years-1960–66)
Numerical dating methods have opened a way for estimation of age of geomor-
phological features such as fault scarps and river terraces (Wallace 1977). This
method has been successfully used in determining the age of the fault scarps in
Southern Arava, Israel (Enzel et al. 1994), West Yellowstone, Montana (Nash
Geomorphic Characteristics and Morphologic Dating … 65
1984), Lost River fault in Custer County, Idaho (Hank and Schwartz 1987),
and terrace risers in Rhine Graben (Niviere and Marquis 2000). In tectonically
active areas, this method characterizes the degradation of a geomorphic marker
based on rate of vertical movement and erosion (Carretier et al. 2002). The evo-
lution of a cohesionless slope through time can be evaluated quantitatively by
assuming that sediment transport is a diffusion process (Colman and Watson 1983).
Generally, the word diffusion describes how chemicals in solution move from areas
of high concentration to areas of low concentration, and how heat moves from areas
of high temperature to areas of low temperature. The sediment diffusion indicates
the gravity transport of sediment from an area of high elevation (the top of the
slope) toward an area of low elevation (the base of the slope). Without fresh uplift
or down cutting, diffusion tends to make slopes smoother and less steep temporally.
Scarp evolution is controlled by climate and several other geological factors.
Measure of age dependent parameters such as maximum scarp slope or the cur-
vature of crest will point towards the age (Wallace 1977). Under conditions in
which the transport efficiency is uniform and the flux is dictated linearly by the local
slope, the combination of the mass flux and mass conservation equations results in
the diffusion equation (Colman and Watson 1983).
d2 1
jt ¼ ðColman and Watson 1983Þ ð1Þ
4p ðtan h tan aÞ2 ;
This solution and model (Eq. 1) is applied as discussed in Colman and Watson
(1983) is applied to determine the morphological age of the Allah Bund Fault
scarp. The parameters applied (Fig. 7) in the equation are given in Table 1.
Over a time, the morphology of a scarp is altered by gravitational and slope
processes (Enzel et al. 1994). For the most part of this time, it follows geomorphic
process of slope modification i.e. mass wasting and wash processes (Summerfield
1991). Fault scarps that cut unconsolidated sediment or soil are very promising for
Fig. 7 Schematic model of a fault scarp showing the various input parameters to be measured for
the solving the diffusion equation (after Colman and Watson 1983)
66 A. Padmalal et al.
For applying morphologic dating technique to the Allah Bund Fault scarp, it is
essential to appreciate its peculiar climato-geomorphic setting. The Kachchh basin
is located in the hyper-arid zone of western India that includes the Thar desert to its
north. The region consequently receives very little rainfall during the monsoon
season. This is testified by the rocky topography of other parts of Kachchh that are
literally free of any vegetation except the thorny bushes. The rivers too are strongly
ephemeral with water flow during monsoon seldom lasting more than a few days.
We therefore consider negligible role of rain splash, slope wash and mass wasting
in the degradation of the Allah Bund Fault scarp.
Apart from the extremely dry climatic regime, the location of Allah Bund Fault
scarp in the Great Rann with peculiar geomorphologic characteristics is also sig-
nificant as far scarp degradation is concerned. In the western part, the scarp marks
the high tide line to the north of the Kori creek. The tidal water flushes against the
scarp, which can lead to erosion especially during extreme tides. However, this
process is limited to the western part only that is also the zone of the maximum
Geomorphic Characteristics and Morphologic Dating … 67
recorded surface effects during 1819 earthquake. As a result, the scarp suffered
maximum degradation in this part as evidenced by its very low height. The highly
gullied nature of the scarp in this part also testifies to the relatively higher mag-
nitude of scarp degradation.
The scarp in the central and eastern parts is located above the high tide line.
Here, the scarp witnesses periodic submergence of the vast supratidal salt flat in
front of it. The submergence however is generally < a meter and rarely reaches the
base of the scarp. We therefore rule out erosion by the submerging waters of the
scarp in this part. The Bet zone also gets submerged by thin sheet of rain water that
flows along the general southward slope. These waters are responsible for the
formation of gullies on top of the scarp. However, as mentioned above, the scarp is
highly gullied in the western part, whereas the gullies are significantly less in the
central and eastern part. This is due to the presence of rivers from the north
dissecting through the scarp to drain into the intertidal flats of the Kori creek.
However, there are no such drainages in the central and eastern parts resulting in
significantly less number of gullies, which are also shallower and wide spaced.
Also, the height of the Allah Bund Fault scarp is the maximum in the central
part. In the eastern part, the scarp height gradually reduces before finally disap-
pearing in the flat Rann surface. Evidently therefore that the scarp has undergone
least degradation in the central part.
Taking into account the climatic and geomorphologic factors as discussed above,
*2 km length in the central part of the Allah Bund Fault scarp is most suitable for
morphologic dating. Moreover, the scarp in this part can be included in the transport
limited category of slope classification i.e. transportation process occurs much
slower than the weathering. For the same reasons, we assume negligible sediment
transportation rate and diffusion constant for calculating the age of the Allah Bund
Fault scarp. Previous studies show that faults that cut unconsolidated sediment are
the most promising ones for diffusion modeling because it forms instantaneously
and will systematically degrade (Colman and Watson 1983). This criterion is met in
the case of Allah Bund scarp, which makes it relatively easier to quantify. Further,
the scarp is not affected by any anthropogenic influences as the Great Rann is an
inhospitable terrain due to its difficult environmental conditions.
Three sites were selected from the central part of the Allah Bund Fault scarp that
shows maximum elevation, is sub-vertical and has clear-cut morphological
expression. The general shape of the scarp is determinable by a single profile
without the need of large sample sizes. The selected scarp face is away from the
vicinity of notches, gullies and local channels. The base of the selected scarp
locations are devoid of colluvium or small fan deposits, which can create irregu-
larities in the scarp profile. The thickness of post-Allah Bund aeolian deposit is
removed from the profile to get the original scarp height. The age of the scarp was
calculated for these three sites as outlined by Colman and Watson (1983) and Nash
(1987).
68 A. Padmalal et al.
6 Dating Results
In this study the application of diffusion equation is used to estimate the age of
Allah Bund scarp. Topographic profiles were prepared using the Survey of India
topographical maps (Nos. 40 H/4, 40 H/8) and the SRTM Digital Elevation Model
data (Fig. 8). The study site is located near the north of Rann, an area with extreme
arid condition. The climate of the region is consistent since last *2000 years after
the marked withdrawal of high sea (Merh and Patel 1988). The salt encrusted
sediment deposits are consistent with such conditions. The Late Holocene sedi-
ments, which make the scarp, consist of fine alluvial (silty clay and clayey silt).
Maximum slope angles measured along the free face at all three places were above
80°. This is consistent with the almost vertical nature of the Allah Bund Fault
scarp. Steeper slope angles further suggest a much younger age for the scarp
(Wallace 1977; Bucknam and Anderson 1979). Average far field (a) is calculated
Fig. 8 Topographic profiles drawn at three sites across the Allah Bund Fault in the central
part. Location of the profiles is marked in Fig. 1b. Vertical scale highly exaggerated. Profiles were
drawn from SRTM data (http://srtm.csi.cgiar.org). The thickness of aeolian deposit capping the
scarp is as measured in the field
Geomorphic Characteristics and Morphologic Dating … 69
by averaging the base and crest far field angles. Since there are no previous studies
available on the rates of sediment deposition, transportation and erosion, we used a
estimated regional diffusion constant generally applied to regions of high aridity
(e.g. Northern Negev, Israel, Begin 1992). The minimum diffusion constant value
of 1 104 m2 year1 matches with the geomorphological scenario of the scarp as
described above. The morphologic ages were calculated using Eq. 1.
Site 1
The topographic profile drawn over the Allah Bund Fault scarp along 24° 5′ 29″N,
69° 19′ 42″E to 24° 3′ 37.6″N, 69° 17′ 20.5″E is shown in Fig. 8. The 2.4 km long
N-S transect shows a near vertical (80°) scarp with measured height of 2.9 m above
the ground level. The thickness of the aeolian sediment cover is subtracted to get
the actual height of the scarp. As described previously the average far field value is
negligible (a = 0°). The scarp marks the abrupt change in the topography from the
supra-tidal flat to the elevated surface of the Bet zone. The morphologic age of the
scarp is determined using Eq. 1 as below.
8:41 1
1 104 t ¼
4 3:14159 ðtan 80Þ2
t ¼ 208 years:
Site 2
Figure 8 shows the topographic profile across the Allah Bund scarp at this site
(24° 5′ 40.7″N, 69° 19′ 31.7″E to 24° 3′ 9.94″N, 69° 17′ 3.2″E). The near vertical
scarp (80.5°) here gave a measured height of 3 m above the supra-tidal salt flat. The
thickness of the aeolian sediment cover was substracted to get the correct scarp
height. Taking the average far field value as negligible (a = 0°), the morphologic
age is determined using Eq. 1 as below.
9 1
1 104 t ¼
4 3:14159 ðtan 80:5Þ2
t ¼ 200 years:
Site 3
The topographic profile was drawn along a N-S transect along 24° 5′ 38.5″N, 69°
19′ 8.60″E to 24° 3′ 22.95″N, 69° 17′ 22.69″E. The profile shows a 2.5 km long
transect with well demarcated sharp contrast in the topography marked by the Allah
Bund Fault scarp (Fig. 8). The scarp marks the abrupt change in elevation of Bet
zone in the upthrown block and the flat monotonous supra-tidal salt flat in the
downthrown block to the south. The total height of the scarp is 4 m, however, the
correct scarp height is determined as 2.8 m after subtracting the thickness of
the aeolian sediment cover. The near vertical scarp (80°) and the negligible average
far field value is used to determine the morphological age using Eq. 1 as below.
70 A. Padmalal et al.
7:84 1
1 104 t ¼
4 3:14159 ðtan 80Þ2
t ¼ 193 years:
Here values of t obtained represent the morphological age i.e. the time since the
fault scarp was formed. The morphological ages of the Allah Bund Fault scarp
obtained at the above three sites are 208, 200 and 193 years before present. These
ages indicate that the scarp was formed in 1809, 1817 and 1824 A.D. respectively.
The results unequivocally suggest that the scarp was produced during the 1819
earthquake event.
7 Concluding Discussion
Watson 1983; Nash 1987). The basic parameters where measured in the field.
Topographic profiles across the scarp at the selected sites were prepared from
topographical map and SRTM data. The present geomorphological environment
along scarp summarized earlier in the article suggests that the central part is a zone
of minimum erosion and subsequently least amount of scarp degradation. The hyper
arid climate and the non influence of anthropogenic activity means the diffusion
constant for arid region can be applied to the Allah Bund scarp for determining its
morphologic ages. We therefore selected the value of diffusion constant as
1 10−4 m2 year−1, which has been applied successfully in the morphologic
dating of fault scarps located in the arid regions (Begin 1992). The morphological
ages of the Allah Bund scarp can be calculated as per Colman and Watson (1983).
The morphological ages of the Allah Bund fault scarp at the three selected sites
yielded an age of 208, 200 and 193 years B. P. Effectively these ages means that the
scarp was produced during the 1819 Allah Bund earthquake, since no other major
earthquake occurred in the Great Rann during the period indicated by the ages i.e.
between 1809 and 1825.
Our studies substantiate the description in the historical accounts regarding the
Allah Bund fault scarp. We therefore rule out the possibility of the scarp being a
product of multiple earthquake events. Our study also suggests that the 1819 Allah
Bund earthquake was the largest earthquake that occurred in the Great Rann after its
emergence from the shallow marginal sea. This is evident from the large-scale
landscape changes, that includes the formation of a *90 km long E-W trending
south facing scarp and also that no other comparable scarp occurs in the vast extent
of the Great Rann. Other earthquake that may have occurred during the time
possibly did not rupture the surface. However, the possibility of earthquakes of
comparable magnitude in prehistoric times in Holocene cannot be ruled out. It is
likely that some of them may have produced surface rupture, however these may be
buried under thick sediments, as the Great Rann was a shallow marginal gulf during
the Late Pleistocene to Holocene (Maurya et al. 2013; Khonde et al. 2017a, b).
It is pertinent to mention here that several large earthquakes have occurred in the
Kachchh rift basin, including the 1819 Allah bund earthquake and 2001 Bhuj
earthquake. The heightened level of seismicity is anomalous in view of the low
level seismic activity witnessed in the surrounding regions viz. Saurashtra in south,
Gujarat alluvial plain in the east and Barmer-Jaisalmer basins in the northeast
(Maurya et al. 2017). It has been suggested that the stress changes related to large
earthquakes in the Kachchh region affect the subsequent pattern of earthquakes. To
et al. (2004) demonstrated that the 1819 Allah Bund earthquake induced stress
changes which led to several moderate magnitude earthquake in the Great Rann.
Post-seismic relaxation following the 1819 earthquake resulted in the loading of
2001 Bhuj earthquake to a significant amount (To et al. 2004). The 2001 event was
followed by long aftershock activity and increased seismic activity along the South
Wagad Fault (SWF), Gedi Fault (GF) and the Island Belt Fault (IBF) (Mandal
2009). Earthquake swarm activity in different part of Saurashtra since 2001 has
been attributed to stress perturbation caused by 2001 Bhuj earthquake (Rastogi
et al. 2013). Low magnitude shocks are repeated from the Gujarat alluvial plain in
72 A. Padmalal et al.
the last 200 years (Maurya et al. 2000). Reactivation of pre-existing lineaments and
faults have resulted in moderate to low seismic activity in Barmer basin also in
recent times (Dasgupta and Mukherjee in press). However, none of the known
historical earthquakes are known to have produced surface rupture apart from the
1819 Allah Bund earthquake. Moreover, the co-seismic and post-seismic stress
changes due to pre-instrumental events still remain a topic of debate (Rajendran and
Rajendran 2001). Detailed palaeoseismic studies are necessary to unravel the
seismic history of the Great Rann for which the Allah Bund fault scarp can serve as
an analog.
References
Baker WE (1846) Remarks on the Allah Bund and on the drainage of the eastern part of the Sind
basin. Transactions Bombay Geographic Society 7, 186–188
Begin ZB (1992) Application of quantitative morphologic dating to paleo-seismicity of
northwestern Negev, Israel. Israel Journal of Earth Science 41, 95–103
Bilham R (1998) Slip parameters for the Rann of Kachchh, India, 16 June 1819, earthquake,
quantified from contemporary accounts. Geological Society Special Publication 146, 295–319
Biswas SK (1974) Landscape of Kutch—a morphotectonic analysis. Indian Journal of Earth
Science 1, 177–190
Biswas SK (1993) Geology of Kutch. K.D. Malaviya Institute of Petroleum Exploration,
Dehradun, p 450
Biswas SK (1987) Regional tectonic framework, structure and evolution of the Western Marginal
Basins of India. Tectonophysics 135, 307–327
Bucknam RL, Anderson RE (1979) Estimation of fault- scarp ages from a scarp-height-slope-angle
relationship. Geology 7, 11–14
Burnes A (1835) Memoir of the eastern branch of the River Indus, giving an account of the
alterations produced on it by an earthquake, also a theory of the formation of the Rann and
some conjunctures on the route of Alexander the Great; drawn up in the years 1827–1828.
Transactions of the Royal Asiatic Society of Great Britain and Ireland 3, 550–588
Carretier S, Ritz JF, Jackson J, Bayasgalan A (2002) Morphologic dating of cumulative reverse
fault scarp: examples from the GurvanBogd fault system, Mongolia. Geophysical Journal
International 148, 256–277
Chandra U (1977) Earthquakes of Peninsular India: a seismotectonic study. Bulletin of
Seismological Society of America 67, 1387–1413
Colman SM, Watson K (1983) Ages estimated from a diffusion-equation model for scarp
degradation. Science 221, 263–265
Dasgupta S, Mukherjee S (In press) Remote sensing in lineament identification: Examples from
western India. In: Billi A, Fagereng A (Eds) Problems and solutions in structural geology and
tectonics. Developments in structural geology and tectonics book series. Series Editor:
Mukherjee S. Elsevier. ISSN: 2542–9000
Geomorphic Characteristics and Morphologic Dating … 73
Enzel Y, Amit R, Bruce J, Harrison J, Porat N (1994) Morphologic dating of fault scarps and
terrace risers in the southern Arava, Israel: comparison to other age-dating techniques and
implications for paleoseismicity. Israel Journal of Earth Science 43, 91–103
Frere HBE (1870) Notes on the Runn of Cutch and neighboring region, J. R. Geograph. Soc.
London 40, 181–207
Glennie KW, Evans G (1976) A reconnaissance of the Great Rann of Kutch, India. Sedimentology
23, 625–647
Grindlay (1808) Diary reproduced in Burnes A., Travels into Bokhara, Volume 3 Appendix. 1834
Gutenberg B, Richer CF (1954) Seismicity of Earth. Princeton University Press, New Jersey, USA,
p 310p
Hanks TC, Schwartz DP (1987) Morphologic dating of the pre-1983 fault scarp on the Lost River
fault at Doublepring Pass Road, Custer County, Idaho, Bulletin of Seismological Society of
America 77, 837–846
Johnston AC (1989) The seismicity of stable continental interiors. In: Gregersen S, Bhasham PW
(eds), Earthquakes at North-Atlantic Passive Margins: Neotectonics and Postglacial Rebound,
Kluver Academic Publishers, pp 299–327
Johnston AC, Kanter LR (1990) Earthquakes in stable continental crust. Scientific American 262,
68–75
Kar A (1995) Geomorphology of the western India. Memoir of Geological Society of India 32,
168–190
Khonde N, Maurya DM, Singh AD, Das A, Chamyal LS (2013) Sediment characteristics and
foraminiferal distribution in the Bet Zone of the Great Rann of Kachchh, Western India.
Geological Society of India, Special Publications 1, 1–15
Khonde N, Maurya, DM, Chamyal LS (2017a) Late Pleistocene-Holocene clay mineral record
from the Great Rann of Kachchh, Western India: implication for palaeoenvironments and
sediment sources. Quaternary International 443, 86–98
Khonde N, Singh SK, Maurya DM, Rai VK, Chamyal LS, Giosan L (2017b)Tracing the Vedic
Saraswati River in the Great Rann of Kachchh. Scientific Reports 7, 5476
Lyell C (1855) A manual of elementary geology or, the ancient changes of the Earth and its
inhabitants as illustrated by geological monuments. John Murray, London, p 655
MacMurdo J (1824) Papers relating to the earthquake which occurred in India in 1819.
Philosophical Magazine 63, 105–177
Mandal P (2009) Estimation of static stress changes after the 2001 Bhuj earthquake: implications
towards the northward spatial migration of the seismic activity in Kachchh, Gujarat. Journal of
the Geological Society of India 74, 487–497
Maurya DM, Raj R, Chamyal LS (2000) History of tectonic evolution of Gujarat alluvial plains,
western India during Quaternary: a review. Journal of Geological Society of India 55, 343–366
Maurya DM, Khonde N, Archana Das, Chowksey V, Chamyal LS (2013) Subsurface sediment
characteristics of the Great Rann of Kachchh from prelimininary textural analysis of two
continuous cores. Current Science 104, 1071–1077
Maurya DM, Tiwari M, Rajawat AS, Kumar H, Khonde N, Chamyal LS (2016) Geomorphic
characterization of the Banni Plain, Kachchh, using orbital imaging radar (RISAT 1C) and
optical remote sensing data. In: Recent studies on the Geology of Kachchh, Thakkar MG (ed),
Geological Society of India Special Publication No. 6, pp. 168–178
Maurya DM, Chowksey V, Patidar AK, Chamyal LS (2017) A review and new data on
neotectonic evolution of active faults in the Kachchh Basin, Western India: legacy of
post-Deccan Trap tectonic inversion. Geological Society, London, Special Publications 445,
237–268
Merh SS (2005) The Great Rann of Kachchh: perceptions of a field geologist. Geological Society
of India 65, 9–25
Merh SS, Patel PP (1988) Quaternary geology and geomorphology of the Ranns of Kutch. In:
Proceedings seminar on recent quaternary studies in India. M.S. University, Baroda, India,
pp 377–393
74 A. Padmalal et al.
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Niviere B, Marquis G (2000) Evolution ofterrave risers along the upper Rhine Graben inferred
from morphologic dating methods: evidence of climatic and tectonic forcing. Geophysical
Journal International 141, 577–594
Nash DB (1984) Morphologic dating of alluvial terrace scarps and fault scarps near West
Yellowstone, Montana. Geological Society of America Bulletin 95, 1413–1424
Nash DB (1987) Reevaluation of the linear-diffusion model for morphologic dating of scarps. In:
Proceedings of conference XXXIX directions in paleoseismology, open-file report, pp 87–673
Oldham RD (1926) The Cutch Earthquake of 16th June 1819 with a revision of the great
earthquake of 12th June 1897. Memoir of Geological Survey of India 46, 71–146
Quittmeyer RC, Jacob KH (1979) Historical and modern seismicity of Pakistan, Afghanistan,
northern and southern Iran. Bulletin of Seismological Society of America 69, 773–823
Patidar AK (2010) Neotectonic studies in southern mainland Kachchh using GPR with special
reference to Katrol Hill Fault, Ph.D. thesis, The M S University of Baroda, Vadodara, India,
163p. Available online at www.shodhganga.com
Rajendran K, Rajendran CP (1999) Seismogenesis in the stable continental interiors: an appraisal
based on two examples from India. Tectonophysics 305, 355–370
Rajendran CP, Rajendran K (2001) Characteristics of deformation and past seismicity associated
with the 1819 Kutch Earthquakes, northwestern India. Bulletin of Seismological Society of
America 91, 407–426
Rastogi BK, Kumar S, Aggrawal SK, Mohan K, Rao N, Rao NP, Kothyari GC (2013) The October
20, 2011 Mw 5.1 Talala earthquake in the stable continental region of India. Natural Hazards.
65, 1197–1216
Roy B, Merh SS (1982) The Great Rann of Kutch: intriguing Quaternary terrain, in Recent
Researches in Geology, Series 9, Hindustan Publication Company, Delhi, pp. 100–108
Summerfield MA (1991) Global Geomorphology: an introduction to the study of landforms,
Essex: Longman, 537p
Thakkar MG, Ngangom M, Thakker PS, Juyal N (2012) Terrain response to the 1819 Allah Bund
earthquake in western Great Rann of Kachchh, Gujarat, India. Current Science 103, 208–212
To A, Bürgmann R, Pollitz F (2004) Postseismic deformation and stress changes following the
1819 Rann of Kachchh, India earthquake: was the 2001 Bhuj earthquake a triggered event?
Geophysical Research Letters 31(13)
Valdiya KS (2002) Saraswati: the river that disappeared, Universities Press, Hyderabad, 116pp
Wallace RE (1977) Profiles and ages of young fault scarps, north-central Nevada. Geological
Society of America Bulletin 88, 1267–1281
Wynne AB (1872) Memoir on the geology of Kutch. Memoir of Geological Survey of India 9, 1–
294
Interplay Between Tectonics & Eustacy
in a Proterozoic Epicratonic, Polyhistory
Basin, North Dharwar Craton
1 Introduction
Modern developments in sedimentary basin analysis (e.g. Einsele 2000; Miall 2000;
Allen and Allen 2013) have accepted the role of Wilson cycle in the formation,
growth and termination of sedimentary basins. The tectonic classifications of sed-
imentary basins (Dickinson 1974; Kingston et al. 1983; Allen et al. 2015) reflect
this perception. The sediment fill of a basin manifests the interplay of sediment
supply, eustacy and subsidence. Climate, topography and petrographic constitution
of the source region control the sediment supply. The geometry and volume of the
fill depends upon the available accommodation space, moderated by tectonics and
eustacy. In Precambrian basins, their deformation and metamorphism hampers the
elucidation of the contents, distribution patterns and sequences of the basin fills.
Sparse age controls in Precambrian sediments when compared to fossiliferous
Phanerozoic sediments further hampers their accurate elaboration (Eriksson et al.
2001). The Indian Peninsular Shield amalgamated into a singular continental block
towards the end of the Archean (2700–2500 Ma) with a series of K-rich granitic
batholiths representing the terminal events of this unification (Naqvi and Rogers
1987; Ramakrishnan and Vaidyanadhan 2010; Meert et al. 2010). Two contrasting
regimes punctuate its ensuing history (Kale 1995). The narrow, linear metamor-
phosed Early and Middle Proterozoic Mobile Belts with compressive tectonic
history (Radhakrishna and Naqvi, 1986; Ramakrishnan and Vaidyanadhan 2010)
are juxtaposed against the wide, extensional epicratonic platform basins that display
S. Patil Pillai
Department of Environmental Sciences, Savitribai Phule Pune University,
Ganeshkhind, Pune 411007, India
V. S. Kale (&)
Advanced Center of Water Resources Development and Management
(ACWADAM), Kshipra Society, Karvenagar, Pune 411052, India
e-mail: dr.vivekale@gmail.com
a paucity of pervasive metamorphism and/or deformation (Fig. 1). The latter are
known as the ‘Purana’ basins (Radhakrishna 1987; Kale and Phansalkar 1991).
They contain thick sequences of shallow marine sediments deposited in discrete
basins which are in tectonic contacts with the mobile belts on one side (Kale 2016).
The Kaladgi and Bhima basins on the northern exposed edge of the Dharwar craton
are the exceptional Purana basins that do not have any Proterozoic mobile belts
adjoining them. The Kaladgi basin is additionally unique amongst the Purana
basins in that it has suffered stronger deformation along the median sectors of the
basin, unlike all others which have deformation restricted to their margins (Kale
1991).
The three Purana basins surrounding the Dharwar craton (Fig. 1) display diverse
modes of evolution (see Kale and Phansalkar 1991; Kale 1991, 2016; Meert et al.
2010; Allen et al. 2015). The Cuddapah basin has a record of its early growth as an
Fig. 1 Precambrian terrains of Peninsular India (modified after Kale 1995; Sharma 2009;
Ramakrishnan and Vaidyanadhan 2010)
Interplay Between Tectonics & Eustacy in a Proterozoic … 77
2 Regional Geology
2.1 Location
The Kaladgi basin extends between the 73°E and 76°E longitudes and 15° 30′N to
17°N latitudes, from the Konkan coastal tract in Sindhudurg district; across the
Western Ghats and the Deccan Plateau in Kolhapur district of Maharashtra state
into the Belgaum, Bagalkot and Bijapur districts of northern Karnataka state.
Isolated outliers of these sediments are scattered south of the main basin around
Saundatti and Gajendragad. Inliers within Deccan Traps occur near Kallamavadi
and near Jamkhandi to the north of the main basin. The westernmost exposures are
found as patchy inliers (within the Deccan Traps) in the Konkan coastal strip,
around Phonda and Malvan. They are collectively referred to as the Konkan
Kaladgis (Sarkar and Soman 1985; Deshpande and Pitale 2014). These discon-
nected outcrops show that the main basin extended far beyond its currently exposed
limits. The present exposures of about 8000 km2 represent only a fraction of the
span of the basin, that can be interpreted to occupy more than 30,000 km2.
The sediments in this basin are subdivided into the Bagalkot and Badami Groups
(Viswanathiah 1979) separated by an angular and erosional unconformity. The
Bagalkot Group occupies an irregular ovoid shaped basin in the eastern parts (Inset
Fig. 2). The Badami Group covers more than 300 km in an E-W direction.
It is significant that while the Konkan Kaladgis are exposed at elevations of less
than 200 m above mean sea level, none of the remaining exposures (east of
Kallamavadi) occur at elevations below 550 m. This elevation difference may the
78
Fig. 2 Regional geological map of the Kaladgi basin and its surrounding areas (compiled from Foote 1876;
GSI 1981; Jayaprakash 2007). Inset is a sketch of the geographic spread of the exposures of the older
Bagalkot and younger Badami Groups showing the two subbasins that developed at different times
S. Patil Pillai and V. S. Kale
Interplay Between Tectonics & Eustacy in a Proterozoic … 79
Table 1 Lithostratigraphy of the Kaladgi basin (after Kale et al. 1999; Kale and Patil Pillai 2011)
The Mallapur Intrusive (sensu: Jayaprakash et al. 1987) should be separated from the Simikeri
Subgroup as the intrusive bodies post-date the Simikeri sedimentation and early deformation (Patil
Pillai et al. 2018) and is hence shown with an asterisk before it
result of the downfaulting of the coastal strip prior to the Late Cretaceous—Early
Paleogene eruption of the Deccan Traps since the trappean basalts do not display
any evidence of faulting of this magnitude. It is likely that this elevation difference
is an artifact of the pre-Trappean downfaulting related to the rifting and separation
of the Indian plate from Madagascar—Syechellis in the Late Cretaceous, but it
could also be a much earlier event. The significance of this is that the Kaladgi
sediments have been subjected to Phanerozoic upheavals along with the rest of the
Dharwar craton. Its local implications could be several, when trying to correlate the
terrain evolution of this basin.
Archaean cratonisation, Proterozoic sedimentation, and the Cretaceous-Tertiary
Deccan volcanism have contributed to the overall geology in this region that was
eventually subjected to emergent erosion, fluvio-colluvial sedimentation and loca-
lised lateritization during Quaternary times. Table 1 summarises the sequence of
strata exposed in the Kaladgi basin and its surroundings.
80 S. Patil Pillai and V. S. Kale
2.2 Basement
The pre-Kaladgi basement profile was essentially a peneplained one with local
undulations, such as one below the Chitrabhanukot Dome, that emerge from the
geophysical data (Ramakrishna and Chayanulu 1988; Mallick et al. 2012). Deeply
weathered basement (sometimes with a paleosol horizon on top) is exposed
immediately below the sedimentary cover, at several locations along the northern
and southern rims of the basin.
Linear ridges of quartzites present on the eastern, northern and southern rims of
the main basin host the basal terrigenous clastics of the Bagalkot Group resting on
the basement. This contact between the basement and the capping sediments is an
angular and erosional non-conformity (Fig. 3a). Shear zones and faults that affect
both the Kaladgi sediments and their basement are exposed at places where the
contact between them is a structural one (see: Fig. 8). The exposures east of Bilgi,
between Idgal and Suriban, southeast of Ramdurg (Fig. 3b) and near Saundatti are
excellent examples of such structural contacts.
The subhorizontal Badami sediments rest on the dipping Bagalkot Group sedi-
ments with an angular and erosional unconformity in the eastern parts of the
Interplay Between Tectonics & Eustacy in a Proterozoic … 81
Fig. 3 a Exposure of the Badami Group (on the top with subhorizontal disposition) resting
unconformably upon the steeply dipping Bagalkot Group sediments (appearing as barren linear
ridges) with a nonconformable contact with the basement at B. N. Jalihal looking eastwards. The
location is about 3 km NE of Badami. b Steeply dipping Saundatti Quartzite exposing the sheared
contact with the basement (on the right hand side—not in the picture). The photo is looking
northeastwards at a location about 4 km east of Ramdurg. Author standing is 1.7 m tall.
c Erosional and angular unconformity between the northwards dipping Manoli Argillite (Bagalkot
Group) and the capping Cave Temple Arenite (Badami Group) at Torgal Tanda located about
3 km south of Ramdurg. The conglomerate is 7 m thick. d Thin cap of subhorizontal Cave Temple
Arenites resting directly on the Archaean Basement at the Gokak hill forming flat-topped mesas
Kaladgi basin (Fig. 3a, c). This was first recorded at B. N. Jalihal (northeast of
Badami) and Torgal Tanda (south of Ramdurg) by Viswanathiah (1968). In the
western parts, the Badami sediments are exposed resting directly upon the basement
without the intervening Bagalkot Group sediments, forming flat topped mesas and
buttes (Fig. 3d).
Sediments in the Kaladgi basin are capped by the Deccan Trap basalts, which
conceal large tracts of the basin, making it difficult to establish their lateral conti-
nuity particularly in the western parts of the basin (Fig. 2). The Deccan Trap basalts
82 S. Patil Pillai and V. S. Kale
were erupted close to the Cretaceous—Tertiary Boundary and manifest one of the
largest continental flood basalt provinces in the world (Renne et al. 2015). At places
near Badami, Belgaum (where it is quarried) and along the western Konkan coastal
tract Late Tertiary—Quaternary lateritic caps are developed on top of the older
strata (Fig. 2). In some of the river valleys Quaternary fluvial sedimentary deposits
are developed capping the Kaladgi sediments. All of them contribute to masking of
the exposures of the Kaladgi rocks and making the lateral correlation difficult.
3 Kaladgi Supergroup
3.1 Stratigraphy
Fig. 4 a Map of the Bagalkot subbasin depicting the exposures of various stratigraphic units. c
Chitrabhanukot Dolomite and Chikshellikeri Limestones are Members of the Petlur Carbonate
Formation. The locations of the stratigraphic logs (Fig. 6a) are given. Note the outline of this basin
marked in this map is traced by connecting the outermost exposures of the Bagalkot Group as
exposed today. The detached exposures south of Manoli are separated from the others by structural
contacts and hence shown separately. They comprise of Saundatti Quartzite. They indicate a larger
southward expanse of this subbasin, which is detached by structural disruptions or has been
disconnected due to subsequent erosion. b Map of the Badami subbasin depicting the distribution
of various stratigraphic units. Locations of the measured logs (Fig. 6b) are given for reference.
West of the limits of the Bagalkot subbasin, the Badami sediments rest directly on the Archean
basement complex
Interplay Between Tectonics & Eustacy in a Proterozoic … 83
These horizons display a lateral facies change into the adjoining lithology. This
lateral facies changes have been earlier elaborated for a part of the basin by Kale
et al. (1996) and Patil Pillai (2004). Consequently there is a repetition of the argillite
—carbonate horizons encountered along several traverses across the basin.
Jayaprakash et al. (1987) and Jayaprakash (2007) recorded these as a number of
‘members’ that they clubbed into ‘formations’. This is not valid within the
framework of standard stratigraphic practices and Walther’s Law of facies suc-
cession (Brenner and McHargue 1988; Catuneanu 2006). A process responsive
lithostratigraphy (Table 1) of simplified, lithology-based, mappable formations was
erected for the Kaladgi basin (Kale et al. 1996, 1999; Patil Pillai 2004; Kale and
Patil Pillai 2011).
Direct onlap of the quartzitic sandstones of the Bagalkot Group without the
intervening conglomerates is seen at numerous locations located along the structural
contact with the basement. This indicates landward transgressive flooding beyond
the structural boundaries of the basin. While the Lokapur Subgroup is spread all
across the Bagalkot subbasin, the Simikeri Subgroup is restricted to the cores of the
synclines (Figs. 2 and 4). The Badami Group does not show any structural contacts
with its basement. In the western exposures of this subbasin, the Bagalkot Group is
not present at all between the Badami sediments and the Archean basement (see
Fig. 2). At no place does the Badami Group rest upon the Simikeri Subgroup. This
could be an indication of the limited development of the Simikeri Subroup, which
has no representatives south of the Badami subbasin.
3.2 Constitutents
Foote (1876) described the Kaladgi sediments within the framework of the con-
temporary formats. Since then, more than a century of research has been invested in
the understanding of the sediments of this basin. Viswanathiah (1979) recognized
four major lithofacies from this basin, namely the sandstones (/quartizites), shales (/
phyllites), carbonates (limestones and dolomites) and chertbreccias. The discovery
of stromatolites from the older carbonates (Viswanathiah and Chandrasekhara
Gowda 1970; Viswanathiah and Sreedhara Murthy 1979) was interpreted as
additional evidence of shallow marine environments.
Jayaprakash et al. (1987) showed that there were three cycles of shallow marine
epicratonic sedimentation within the basin, yielding three successions separated
from each other by unconformable contacts. Kale et al. (1996) further refined this
view using lithofacies classification to the southern parts of the basin (see: Figs. 7
and 8 of Kale et al. 1996) and later to the basin as a whole (Kale et al. 1999; Kale
and Patil Pillai 2011). A summary of the petrological characters of these facies
based on our studies is given below.
The Bagalkot Group is characterized by two sequences comprising of four
lithofacies, namely (a) detrital siliciclastic arenites and rudites (sandstones and
conglomerates), (b) argillites (siltstones, shales, mudstones and phyllites),
Interplay Between Tectonics & Eustacy in a Proterozoic … 85
JFig. 5 a Thickly bedded conglomerates with parallel stratified interlayers of sandstones occurring
at the base of the Badami Group (at Gokak falls). They grade upward into the sandstone beds. The
photographed section is about 5 m high. b Parallel bedded quartzitic sandstones of the Saundatti
Quartzite exposed in a quarry face near Bilgi. Beds dip gently towards the south (right of photo).
The quarry face is about 10 m high. c Current ripple marks on the bedding plane of Muchkundi
Quartzite exposed near the Muchkundi reservoir. Note the change in ripple axis orientations in
successive beds. d Trough cross bedded sandstones of Badami Group (exposed around Badami
town). The geometry of troughs, indicate small shifting channels similar to those in braided
channels of mature river systems close to their mouth (in estuaries/deltas). e Lenticular beds of
Yadhalli Argillite exposed near Manoli. Note the gently wavy nature of the bedding in the upper
part of the photograph. f Thin parallel bedding with alternations of grey (siliceous) and purple
(ferruginous) shales of Halkurki Formation exposed near Halkurki. They are extremely friable and
weather easily. Hammer for scale is 34 cm long. g Alternate laminations of calcareous (buff
colored) and pure crystalline limestones (dark grey colored) from the Chikshellikeri Limestone
exposed near Gaddankeri. Note the climbing ripple laminated beds interspersed within the parallel
laminated bedding. They indicate reworking of the lime-mud under influence of currents. Pinching
and swelling of the lamina thicknesses is due to uneven compaction. h Variegated parallel
stratified Kokankoppa Limestone with an interbedded limestone breccia horizon. The clasts of the
breccia are embedded in an impure calcareous muddy matrix. Such intraformational breccia
horizons testify to intermittent emergence (shallowing) or may manifest synsedimentary
deformation
influence during their deposition. The sandstones from the Bagalkot Group have
been altered to glassy orthoquartzites due to diagentic recrystallisation and further
during deformation.
The Badami sandstones are dominantly trough cross-bedded (Fig. 5d) with
occassional birectional and parallel stratification. This suggests their deposition in
braided fluvial to transitional marine environments with minor tidal influence.
Slope-base conglomerates and the sandbars, over-bank mud drapes on hummocky
sand beds and upward fining beds that are interrupted by gravel beds or local
erosional surfaces (overlain by coarser sands) are often seen. These characters and
the stratal patterns of dominantly wide amplitude trough-cross bedded channels
suggest the deposition in braided channels. They are therefore attributed to a wide,
mature braided river system that drained into a shore-face environment. Few of
these sandstones (e.g. those exposed north of Ramdurg and northwest of Badami)
display tidal silty interlayers. Their stratification with low amplitude shallow and
wide troughs stacked one above the other (Fig. 5d) is also suggestive of their being
estuarine or deltaic channel deposits.
3.2.2 Argillites
Siltstones, shales and mudstones represent this facies. The Bagalkot Group also
contains horizons of phyllitic shales (with chlorite neocrystallisation) having
well-developed slaty cleavages. This is the result of the localised, low grade
metamorphism in the strongly deformed northern sector of the Bagalkot subbasin.
88 S. Patil Pillai and V. S. Kale
3.2.3 Chertbreccias
They occur only in the Bagalkot Group but are not present in the Badami
Group. Two separate horizons are present in the Lokapur and Simikeri Subgroups
(Mahakut and Niralkeri Chertbreccia Formations respectively). In field exposures
they can be easily misidentified as fault breccias, but are unequivocally interbedded
horizons within the rest of the Bagalkot sediments. They have been deformed
coherently with the other enveloping sediments of this Group. They are interpreted
as washed fault-generated debris produced during the synsedimentary tectonic
movements along the basinal faults and subsequently silicified (Kale and Patil Pillai
2011).
The chertbreccia and argillaceous beds display a mutually exclusive develop-
ment in the Bagalkot sequence. In the sectors of the basin that suffered synsedi-
mentary tectonism, where chertbreccia deposits developed, the extent of the
mudflats was curtailed. This led to accumulation of thinner deposits of the
argillaceous sediments. In other sectors such as in the southern sector around
Manoli and west of Badami, where the basin floor subsided in a more gentle
continuum, wide tidal mudflats yielded thicker argillaceous deposits (see: Kale et al.
1996).
Interplay Between Tectonics & Eustacy in a Proterozoic … 89
3.2.4 Carbonates
This facies is represented by limestones and dolomites, with the latter being
restricted to the Bagalkot Group. The limestones display a variety of colours (pink,
white, green and grey) attributed to the impurities in them. The limestones of
Bagalkot Group display flat, parallel-laminated alternations of impure lime muds
and pure crystalline limestone. Occurrence of climbing ripple laminations and
cross-stratified lenses (Fig. 5g) indicate that they were deposited in a very shallow
shelf affected by currents. The dolomites or dolostones of the Bagalkot Group
display tabular parallel-laminated stratification. They are often diagenetically sili-
cified. They contain wavy algal-laminated dolomites with stromatolitic colonies
(Viswanathiah and Sreedhara Murthy 1979; Sharma et al. 1998) indicating water
depths of less than 20 m and a warm equitable climate. These carbonates are
interpreted to be deposited in a carbonate flat regime with occasional interruptions
of storm surges manifested as thin sandy lenses. Occurrence of intraformational
limestone breccia horizons (Kale et al. 1998) in them is indicative of their episodic
emergence or the manifestation of synsedimentary deformational events during
their accumulation.
The flaggy limestones of Badami Group are varigated (Fig. 5h) and contain thin
lensoid intercalations of siliciclastics. They display wave ripples, clay galls, rip up
clasts, tidal couplet laminae suggesting a tidally influenced depositional setting.
Presence of glauconites in the sandy lenses suggests locally developed reducing
environments during their deposition.
Basic dykes and quartz veins intruding discordantly into the sediments have been
recorded from the northern folded sector of the Bagalkot subbasin (Foote 1876;
Jayaprakash et al. 1987; Jayaprakash 2007). They are collectively named as
Mallapur Intrusives Formation of the Simikeri Subgroup. Their occurrence along
the axial planes of folds in the Simikeri Subgroup indicates that they were emplaced
after the cessation of sedimentation. Therefore they are separated as an independent
formation post-dating the Simikeri Subgroup as listed in Table 1 (Patil Pillai et al.
2018).
3.3 Lithologs
Besides the detailed lithological mapping (Fig. 4), logs of stratigraphic sections
were reconstructed based on sections measured during field studies (Fig. 6). They
include the sedimentary structures, stratal patterns and thicknesses of the various
lithotypes.
90 S. Patil Pillai and V. S. Kale
Interplay Between Tectonics & Eustacy in a Proterozoic … 91
Fig. 6 (continued)
92 S. Patil Pillai and V. S. Kale
Petrological studies of the strata from these sections, to establish their textural
and composition are embedded in these logs. These logs provide a synoptic view of
the various lithotypes occuring in different sections. In conjunction with the geo-
logical map they provide a clear understanding of the vertical and lateral distri-
bution of various horizons from the Kaladgi basin.
A combination of supply (types and rates) and accommodation space (as a function
of relative subsidence, driven by eustacy and tectonics) defines the sedimentation
patterns and contents in a sedimentary basin (Catuneanu et al. 2005; Allen and
Allen 2013). The motifs of the sediment piles that accumulate at different locations
within a basin manifest the patterns of relative subsidence in those parts. The
reconstruction of successive stages of sediment accumulation within a basin plotted
against time enables understand the relative subsidence of the basin floor at different
stages of its evolution (Gallagher and Lambeck 1989).
This analysis requires primary data of (a) the lithotypes and their compaction and
diagenetic history, (b) estimated paleobathymetry of the different types of sedi-
ments, (c) stratigraphic (plotted as vertical) succession of the constituent sediments,
(d) the ages of the different strata within the stratigraphic succession; and
(e) backstripping that enables quantify the removal or loss of sediments that were
accumulated but are not preserved (Sclater and Christie 1980; Bond and Kominz
1984).
In case of the Phanerozoic successions, in which accurate ages of the strata are
provided by their fossil assemblages, such subsidence analysis has led to the
development of a robust global eustatic sea-level curve (e.g. Vail—EXXON curve
—Vail et al. 1977; Haq et al. 1987; Miller et al. 2005). However, this is not possible
in the unfossiliferrous Precambrian sequences where age data is sparse
Interplay Between Tectonics & Eustacy in a Proterozoic … 93
(Catuneanu et al. 2005). In such cases, accurate determination of the net rate of
subsidence of the basin floor cannot be plotted and needs to be estimated using
other proxies. Kale (1991) used these principles to plot the ‘sediment accumulation
curves’ (‘SAC’s) of the Purana basins, using the total accumulated stratigraphic
thickness in the basin as a proxy of 100% time of the basin evolution. The maxi-
mum stratigraphic thickness of each stratigraphic unit computed as a percentile
value of the total thickness represents the relative time taken by it in the basin. This
yields a relative (rather than absolute—as in case of the Phanerozoic) scale for the
basin. We have used our observations (as enumerated in Figs. 4 and 6) and the
maximum thicknesses of strata as defined by Jayaprakash (2007) in these compu-
tations. The cumulative uncompacted maximum stratigraphic thickness for the
Bagalkot Group is 13,717 m and for the Badami Group is 1846 m. These values
match well with the geophysically determined maximum depth to basement (e.g.
Ramakrishna and Chayanulu 1988; Kale 1991) of the Kaladgi basin.
This analysis uses Cant’s (1989) equation that derives a relationship between
eustacy (E), subsidence (Sub), sediment accumulation (Sed) and the depth of the
depositional interface (D). It is postulated that the change in water depth (DD)
during the deposition of a single uninterrupted stratum with thickness (DSed), in a
basin where space is created due to subsidence (tectonic or gravity driven) of the
basin floor (DSub) and eustatic sea-level change (DE) will be as follows:
If the combined effects of subsidence and eustacy are clubbed together into a
single parameter (DDI), this equation can be simplified into the following form:
In other words, the sum of sediment thickness and its depositional depth of each
strata gives the net change in the depth of the depositional interface due to subsi-
dence and eustacy:
Table 2 Listing of compaction factors and depositional depths used in the subsidence analysis of
the Kaladgi sediments
Lithology Value
Compaction factor
1. Highly compacted and recrystallized sandstones (quartzites) 0.76
2. Well-indurated, cemented sandstones 0.86
3. Siltstones 0.53
4. Mudstones and shales 0.41
5. Carbonates (without significant recrystallisation) 0.75
6. Carbonates (recrystallized) 0.69
Depositional depth in meters
1. Carbonates (unaffected by influence of surface waves) -7
2. Carbonates (deposited within wave-influence depths) -5
3. Tidal muds and silts (=shales) -3
4. Supratidal and beach sandstones +3
5. Conglomerates +7
6. Chertbreccias (reworked tectonically generated debris) 0
Fig. 7 SACs of the Kaladgi Supergroup. The curves are numbered as per the stratigraphic section
(Fig. 6a, b) used to plot them. For the Bagalkot Group rows are arranged from east to west (right to
left) respectively to give a perception of the along-strike changes in subsidence patterns. The rows
are arranged north to south (top-bottom) respectively. For the Badami Group only two
representative SACs are given
history of the Badami Group. The steeper Bagalkot SACs suggest a faster subsi-
dence, which must have been enabled by tectonic deepening besides the eustatic
changes in the sea-level through its evolutionary history.
The SACs of the section numbers 2, 7, 9, 14, 15, 16 displays a consistent
gradient with little or no deflection; pointing to a rapid subsidence with a fairly
consistent rate. These sections are not located close to any of the basement shears.
On the other hand, the SACs of section numbers 1, 3, 6, 10, 11 that are located near
96 S. Patil Pillai and V. S. Kale
4 Structural Configuration
For understanding the evolution of the basin, it is necessary to try and separate the
diverse influences (basement undulations, supply rate variations, sea-level changes
and tectonics) and their imprints in the sedimentary record. The primary require-
ment for enabling this was to establish the three-dimensional geometry of various
horizons. This is relatively straight forward in case of the subhorizontal Badami
Group. However, the tightly folded sequence of the Bagalkot Group does not
permit the lateral continuity of the lithostratigraphic units to be established without
deciphering the structural patterns and ‘unfolding’ the strata.
The deformation patterns in the basin were earlier documented by Awati and
Kalaswad (1978), Nair and Raju (1987), Jayaprakash et al. (1987), Gokhale and
Pujar (1989), Jayaprakash (2007), Mukherjee et al. (2016), Jadhav and Kshirsagar
(2017) amongst others. Traverse mapping to validate earlier observations and field
observations on the orientations of the bedding planes, fractures and phyllitic
cleavage (in case of the argillites) have been used to understand the sense of
movement in different folds and faults in the basin.
Figure 8 compiles these observations in the Bagalkot subbasin that has under-
gone significant tectonic deformation. It is apparently divided into the northern and
southern sectors (Figs. 4a and 8) that are structurally different from each other,
separated by younger cover. The northern sector hosts tight eastwards or westwards
plunging folds. The southern sector exposes monoclinal, northward dipping strata
from the Lokapur Subgroup (Fig. 8). This stronger deformation of the Bagalkot
Group in the central parts than that along its fringes is akin to mobile belts. This
contrasts to that recorded from the other Purana basins, which display deformed
margins and relatively undeformed mid-basinal sectors.
The narrow linear Badami subbasin (Inset Figs. 2 and 4b) is almost undeformed.
The Badami Group is characterized by subhorizontal dips and small local faults
showing dragging along them. No other structural deformation is recorded from this
Group.
The Kaladgi basin displays several sets of inherent as well as superimposed faults
(with evidence of brittle deformation and displacement) and shears (zones of
quasi-ductile to ductile deformation). The sense of slip along them is recorded
based on slickenside surfaces, drag of the bedding and joint patterns along them.
They dominantly trend E-W to ESE-WNW with gentle swirls along their lengths
and occur close to the margins of the basin.
The Sirur Shear zone extends for >100 km across the central part of the basin
and further eastwards into the basement terrain. It displays a minor right-lateral
sense of movement and a larger dip-slip component with downthrows towards
98 S. Patil Pillai and V. S. Kale
Fig. 8 Structural map of the Bagalkot subbasin showing the major folds and faults. The dotted
outline marks the boundary of this subbasin that is surrounded on all sides by basement rocks.
Regional shear zones such as the Kolchi and Anagwadi Shears have been recorded in geophysical
studies as crustal scale lineaments. The Sirur Shear separates the northern folded sector and the
southern monoclinal (northward dipping) sector of this subbasin
north. This shear zone (Fig. 8) is largely concealed below younger cover (of the
Badami Group and Deccan Traps). It occurs along the boundary between the
northern and southern sectors of the Bagalkot subbasin. It also marks the northern
edge of the Badami subbasin in the eastern sector (Fig. 4b). This suggests that it
may have suffered a reversal of sense of movement, yielding southward down-
throws during the Badami sedimentation. Most of the other E-W trending faults
display dominantly dip-slip displacements with basin-ward downthrows, and neg-
ligible to no strike-slip components.
The NW-SE trending sinistral Anagvadi and Kolchi Shears extend into the
basement. The imprints of the former are exposed across the Bagalkot subbasin.
The northward continuity of the Kolchi Shear is concealed below younger cover
and is hence shown with a thinner line in Fig. 8. Both correspond to megascale
geophysical lineaments of Peninsular India (Sridhar et al. 2017; Rajaram et al.
2016) that extend far beyond the bounds of the Kaladgi basin.
Minor faults have been mapped in several parts of the basin with NNE-SSW,
N-S, NNW-SSE trends (Fig. 8). They appear related to folding as sympathetic
fractures that show brittle deformation when transecting the quartzitic and
Interplay Between Tectonics & Eustacy in a Proterozoic … 99
conglomeratic horizons, while the argillites and carbonates suffered ductile drag-
ging coherent with the sense of movement (see Mukherjee 2014).
The synsedimentary deformation from the Bagalkot Group is oriented parallel to
the adjoining shears and faults (Kale et al. 1998; Kale and Patil Pillai 2011; Patil
Pillai and Kale 2011). This indicates that they may be inherited structures from the
basement that were reactivated during and after the sedimentation in the Kaladgi
basin. Faulted contact of the basement and cover sediments when traced laterally,
assumes the nature of an unconformable relation, with the trace of the fault con-
cealed below the sedimentary cover. This suggests that although faulting played an
active role in the basin subsidence, transgressive onlap of sediments beyond the
faults did take place. The combination of the NW-SE trending regional shears and
the E-W trending Sirur Shear suggest that basinal growth took place in a
transtensional regime.
Several cross-faults that display quasi-ductile deformation in the argillaceous
and carbonate sediments cross the quarzitic ridges as brittle faults lined with fault
breccia. This aspect of competency contrast provides a better understanding of the
stress distribution during deformation and provides clues to the sense of slip along
them.
4.2 Folds
Folds in the Bagalkot subbasin are easily mappable from aerial photographs and
satellite imageries (Awati and Kalaswad 1978; Nair and Raju 1987) due to the rims
of the quartzite ridges along them (Fig. 9). These folds are large amplitude, small
wavelength ‘close folds’ with inter-limb angles of less than 30°. They display
plunging fold axes and curviplanar axial planes (as shown in Figs. 8 and 9).
This has been interpreted as a weaker event of cross folding (Mukherjee et al.
2016; Jadhav and Kshirsagar 2017). Alternatively, the fold geometry may have also
been influenced by inherited structures such as the basement shears and paleoto-
pographic undulations, yielding a perception of a second phase of coaxial folding.
Incipient greenschist facies metamorphism is manifested as development of
chlorite, sericite, albite (Govinda Rajulu and Chandrasekhar Gowda 1972) and
phyllitic cleavages in argillaceous sediments; and uralitization of pyroxenes and
albitisation of plagioclase in the Mallapur Intrusives (Patil Pillai et al. 2018). It is
pronounced at locations where the secondary faults cut across them and along fold
axes, but becomes indistinct away from the faults and on the fold limbs.
4.3 Synthesis
Three N-S geological cross-sections are reconstructed to depict the complex nature
of deformation and the stratigraphic relationships between various horizons in the
100 S. Patil Pillai and V. S. Kale
Fig. 9 Sketch maps of major folds in the Bagalkot subbasin. The fold axes are shown with their
plunge direction and anticlinal/synclinal symbol. Faults are shown with thin dotted red lines.
Strike-slip faults are depicted with the sense of slip along them. Faults with dominant dip-slip
component are shown without any marking of sense of movement along them
Kaladgi basin (Fig. 10). Besides enabling understanding of the lateral facies
changes, they also show the relationships between various folds and faults in the
deformed sector of the Bagalkot subbasin.
The E-W and NW-SE trending regional shears in the basement continue as faults
affecting the sediments. The parallel orientations of the synsedimentary deforma-
tional structures to them (Patil Pillai and Kale 2011) indicate that they were active
during the sedimentation in the basin (D0) and have influenced the basin geometry.
The sediment distribution patterns as lithofacies stripes parallel to these basement
structures reinforces this interpretation. They controlled its deepening by subsi-
dence of the continent-margin shallow marine platform.
The pattern of the SACs (Fig. 7), the shore-line—shelf depositional regimes of
the sediments and their stripping patterns collectively indicate that extensional
subsidence was responsible for the growth of the basin. Presence of synsedimentary
extensional faults as documented by Patil Pillai and Kale (2011; Fig. 5) in the lower
sequences of the Kaladgi basin and the onlap of sediments across bounding faults
are additional evidences in support of multiple transgressions that occurred in a
progressively deepening basin. In absence of systematic geophysical or drilling
data, it is difficult to validate the depth penetration of the structures recognized
during surface mapping as depicted in the cross-sections (Fig. 10).
Interplay Between Tectonics & Eustacy in a Proterozoic … 101
Fig. 10 Sketches of N-S geological cross sections across the eastern part of the Kaladgi basin at
locations given in Fig. 9. The sections are N-S sections along the longitudes of 75° 20′E, 75° 35′E
and 75 45′E from top to bottom respectively. Note that the subsurface extension of the dips of beds
and of the fault/shear zones have limited validation from field observation
The Badami sediments were deposited on the eroded surface of the deformed
Bagalkot Group (D3). These younger sediments do not display significant defor-
mation except for some local faults. Its depositional characters dominated by
continental deposition suggest that the younger Badami subbasin was nested in the
uplifted and eroded fold-mountains of the Bagalkot Group in the north and the
ridges of the Saundatti Quartzites in the south.
5 Basin Analysis
This basin had developed on the continental crust of the Dharwar craton (Kale
1991; Basu and Bickford 2015; Allen et al. 2015). The Dharwarian basement
served as the source of its sediments. The geometric relationship between the
regional basement shears that were active during the sedimentation and the sedi-
mentation patterns of the Bagalkot Group indicates that movements along them
facilitated the inception and growth of the basin. The tectonic influence on the
evolutionary history of this basin therefore precludes the possibility of this being an
intracratonic sag as suggested earlier (Jayaprakash et al. 1987; Jayaprakash 2007;
Dey et al. 2009; Dey 2015). A different model of the evolution of the Kaladgi basin
emerges based on the evidence and analyses documented above.
While integrating these observations, the ages of different events (Table 3) inferred
by Patil Pillai et al. (2018) are used. It is necessary to recognize that the ranges of
the ages inferred in the Table 3 are not the durations of the sedimentation. An
average sediment accumulation rate of 0.1 mm/year yields a duration of *130
million years to the Bagalkot Group and *19 million years for the Badami
Group. Modern accumulation rates in deltas range between 1 mm/year (=1 m/
k year) and 10 cm/year (=0.1 km/k year). They are slightly different, but in the
same magnitude range for shelf seas (Saddler 1999; Sommerfield 2006). Using
analogs of Phanerozoic sediments from the near-shore systems, the actual durations
of sedimentation in the Kaladgi basin are more likely to be a fraction of the
estimates based on average sedimentation rates (Kale 2015, 2016). This adds to
certain uncertainties when tagging the ages of the events in this basin.
Given the ages of the Mesoproterozoic dyke swarms, occurring in the adjoining
areas (Kumar et al. 2015), it is evident that the Kaladgi sedimentation cannot be
older than 1800 Ma (see also Basu and Bickford 2015). This cratonic block of the
Interplay Between Tectonics & Eustacy in a Proterozoic … 103
Table 3 Chronology of events in the Kaladgi basin modified after Patil Pillai et al. (2018)
Subbasin Sedimentological events Tectono-thermal events Estimated age
range/measured
age (Ma)
Badami Continental sedimentation Extension (D3) *900–800
with localised marine
transgression
Emergence and erosion of Hiatal break 1100–900
folded strata (+basement)
Bagalkot Post depositional events Low grade 1150–1100
metamorphism and
deformation (D2′)
Mafic intrusion 1154 ± 4
Folding and faulting
(D2)
Basin deepening with Extension *1200
synsedimentary intra-basinal
tectonics
Local intra-basinal Intra-basinal tectonics >1200
reorganisation (?) with co-axial
deformation (D1)
Peak transgression Extension and <1300 + 50
Basin inception subsidence (D0) *1400 + 50
Basement Emergence and Continental regime >1600 Ma
peneplanation
5.1.2 Growth
The growth of the Kaladgi basin between 1400–1300 Ma is consistent in the global
context of marginal break-up of the supercontinents. The upper crustal regime
would normally be expected to fail (under extensional stress) along brittle fault
planes. The brittle upper crustal levels of the continental crust would essentially
display failure during extension with fault complexes analogous to those found in
present day continent margin rifts (Gibbs 1984; Allen and Allen 2013) as postulated
in the model in Fig. 11a. The cross-sections across the Kaladgi basin (Fig. 10), and
fault-plane orientations, can be modeled into an analogous listric system.
A postulated listric geometry of the main faults/shear zones that governed the
growth of this basin is endorsed by the fact that at several locations, mesoscopic
listric and antithetic fault complexes of meter-scale dimensions are seen (Fig. 11b–
e). The field exposures not only endorse the basin-scale listric geometry of the
growth fault complex, but also show it’s replicability at smaller scales. The max-
imum recorded vertical thickness of sediments (of *5000–6500 m) suggests that
the depth of the detachment plane was located at those depths.
Given the element of horizontal movements recorded from various associated
faults (see Figs. 8 and 10); the extension was associated with a degree of horizontal
shear. It is therefore interpreted that the early growth of the Kaladgi basin was in a
transtensional regime rather than purely extension driven by gravity as previously
modeled by Mukherjee et al. (2016). The essentially extensional, fault-driven
subsidence (=D0) of the floor of the Bagalkot subbasin enabled accumulation of
thick sediments in it.
A minor, localized event of basin floor reorganization led to the intrabasinal
unconformity between the Lokapur and Simikeri Subgroups in the northern sector
of this subbasin (Kale et al. 1998; Jayaprakash 2007). The Simikeri sediments are
confined within the doubly plunging synclinorial areas. These areas are closely
aligned with the Anagvadi, Sirur and Gaddankeri Shears (see Fig. 8). We infer that
the localized restructuring of the basin floor in these narrow parts of the Bagalkot
subbasin may be linked to reactivation of these shears (D1). The incremental
occurrence of intraformational limestone breccia horizons towards the Gaddankeri
Shear (Kale et al. 1998) endorses it.
Sedimentation in other parts of the basin appears to be uninterrupted by this
localized event (D1). It may have also produced the first generation folds in the
central parts of this subbasin, leading to the exposure of some of the strata to
weathering and erosion. The emergent strata of the Lokapur Subgroup have con-
tributed sediments to the Simikeri Subgroup as evident in basal conglomerates from
the Muchkundi Quartzites. This event reflects as kinks in the SACs that cover both
the subgroups from the central parts of the basin (e.g. nos. 6 and 10 in Fig. 7).
Extension, consequent relative deepening of the basin and sedimentation continued
after that for a relatively shorter duration, during which the thinner Simikeri
Subgroup was deposited. Based on the chronological data available (Patil Pillai
et al. 2018) and the assumption of sedimentation rates ranging between *0.1 to
1.0 mm/year; the event D1 is estimated to have occurred slightly before 1200 Ma.
Interplay Between Tectonics & Eustacy in a Proterozoic … 105
Fig. 11 a Schematic N-S model of the fault/shear framework of the Bagalkot subbasin based on
models of continental rifting with a listric geometry predicted by Bott and Mithen Bott and Mithen
1983; Dula 1991). The inferred listric geometry is based on field evidence of movement along
various faults. The postulated northward verging listric geometry is endorsed by the occurrence of
similar mesoscopic arrays of faults, whose examples are given below (b and d). b Field photograph
of subvertical dipping Chikshellikeri Limestone near Gaddankeri. Uneven compaction of the
alternating impure and pure limestone strata have produced several synsedimentary deformation
features such as compaction lobes, flame structures and pinching and swelling of the parallel
bedded strata. Note the small-scale fault system transecting it, which has not affected the overlying
(top of the photo) strata indicating its intraformational nature. c Sketch of the photograph in
b. Note two sets of small listric faults and antithetic faults terminating against the substrate of
undisturbed strata can be marked in this exposure. The lack of penetration into the underlying
strata and termination against the overlying bed confirms their intraformational status. d Field
photograph of subhorizontal Kokankoppa Limestone at Chandargi village (see Fig. 4b). Weakly
stylolitic parallel bedded limestones are interspersed with thin muddy impure beds resulting in a
flaggy appearance to them. A small fault system has affected only the strata in between; that shows
a ‘saucer-like geometry’ analogous to listric fault systems. e Sketch of the photograph in d. Their
intraformational nature of the listric array of faults is established by the lack of penetration into the
underlying strata and termination against the overlying bed. Typical antithetic faults and roll-over
anticline is observed on the northern (right) side of the picture as marked
106 S. Patil Pillai and V. S. Kale
5.1.3 Deformation
The Bagalkot sediments were folded and faulted during the event recognized as D2.
The sub-parallel alignment of the fold-axes with the NW-SE to WNW-ESE regional
shear zones indicates that this event was related to another pulse of activity along
them or that these zones channelized N-S oriented compressive stress systems. The
D2 event did not affect the southern sector of the Bagalkot subbasin as much as it
did the northern sector.
The deformation of the Bagalkot Group was interspersed with the intrusion of
dykes (=Mallapur Intrusives) along the D2 fold axes. The dykes are dated at
1154 ± 4 Ma by 40Ar/39Ar method (Patil Pillai et al. 2018). The estimated age
(*1200 Ma: based on the dating of a dyke (Mallapur Intrusives) that is emplaced
along the D2 fold axes) of this event corresponds well with the period of super-
continental reassembly leading to the development of Rodinia (Condie and Kroner
2013; Condie et al. 2015).
Further deformation and low grade greenschist facies metamorphism of the
sediments along with the intrusives is recognized as the event D2’, since there is
insufficient evidence to say whether it represents the peak of the deformation event
(D2) or is a separate episode altogether. Cross faulting of the earlier folds and
co-axial second generation shear folds observed here are likely to have evolved at a
later stage than the main folding. Jadhav and Phadke (1989) recognize two separate
events of folding while Jayaprakash (2007) records only a single deformational
event for the Bagalkot Group. The sedimentation in this subbasin ceased following
the deformation and low grade metamorphism.
The subsequent period of quiescence coincides with the period of stability of the
Rodinia supercontinent estimated between *1100 and 900 Ma. This period was
marked by the emergence and erosion of the folded hills of the Bagalkot Group
along with its basement. The deposition in the Badami subbasin is essentially a
continental sedimentation with very little component of the marine interface. The
presence of ichnofossils from the Badami sediments (Kulkarni and Borkar 1997)
confirm their Neoproterozoic age. Given that some parts of the Badami Group rest
directly upon the basement, suggests a significant removal of the preceding
Bagalkot sediments prior to the inception of these younger sediments. We therefore
project the age of the extensional growth (D3) of the Badami subbasin to be
between 800 and 900 Ma corresponding to the age of the early breakup of Rodinia
before its reorganization into the Gondwanaland assembly (Cawood and
Hawkesworth 2014).
The thickly-bedded, poorly-sorted conglomerates at the base of the Badami
Group are analogous to debris-wash deposits commonly found at the base of slopes
in a continental regime. Fluvial influence in the form of braided channel sands is
overprinted on the sandstones of the Badami Group. Taken together with the
Interplay Between Tectonics & Eustacy in a Proterozoic … 107
narrow linear expanse of the Badami subbasin, a bulk of these sediments are
interpreted as continental fluvial deposits derived from the flanking hill ranges of
the folded Bagalkot Group and its basement. They are eventually overlain on the
eastern side by the mudflat siltstones and flaggy impure limestones. This indicates
that a transgressive event on a shallow shelf occurred during this sequence of
sedimentation. The adjoining sandstones in this tract between Yargatti–Chandargi
(on the west) and Badami (in the east) show evidence of deposition in a near-shore
environment. It is concluded that the Badami sedimentation occurred in a nested
basin, attended by limited extension and subsidence. The subsidence curves
(Fig. 7b) indicate this to be a very slow subsidence that was overtaken by a marine
transgression during its terminal phase.
Fig. 12 Depositional environments of the successive lithostratigraphic units from the Kaladgi
Supergroup. The relative sea-level curve plotted is based on the depth of deposition of
corresponding lithologies and their mutual interrelations (adjoining/overlapping) across different
parts of the basin. The tectonic events interpreted in this work are depicted in the next column to
demonstrate the relative temporal relation between the tectonic events and sea-level fluctuations as
inferred by us
Piper 2013; Rooney et al. 2015; Condie et al. 2015) that controlled Proterozoic
sedimentary basin evolution (Eriksson et al. 2006; Allen et al. 2015).
The Kaladgi basin was described as an intracratonic sag/rift basin by earlier workers
(Jayaprakash 2007; Dey 2015). Using the classification parameters of sedimentary
basins (see Allen et al. 2015), the Kaladgi basin qualifies as an intracratonic/
continent interior basin, bounded on all sides by the basement rocks of the Dharwar
craton. Miall et al. (2015) consider this to be a ‘fault-bounded basin related to far
field tectonism’, given that there is no contemporary tectonic zone in its vicinity.
Our work shows that the two constituent Groups in this basin display different
tectonic styles. A unified classification of this basin is therefore not appropriate. It
must therefore be recognized as a polyhistory basin (sensu Kingston et al. 1983).
Interplay Between Tectonics & Eustacy in a Proterozoic … 109
6 Conclusions
Acknowledgements This work is the result of more than 3 decades of research in the Kaladgi
basin that was funded by CSIR, DST and AMD. Different students and colleagues at different
times have helped in the data collection and we acknowledge their contributions to the work. We
are thankful to Prof. V. V. Peshwa, Prof. T. K. Biswal, Prof. V. G. Phansalkar, Dr. Amogh Chitrao,
Dr. Anand Kale, Dr. Himanshu Kulkarni and Dr. Pradeep Jadhav for their suggestions and dis-
cussions on this work. We thank Prof. Soumyajit Mukherjee for inviting us to contribute this
chapter and his patient support and encouragement. The comments and suggestions from the
reviewers helped improve the manuscript significantly.
References
Allen PA, Allen JR (2013) Basin analysis—principles and applications, 3rd edn. Wiley-Blackwell,
549pp. ISBN: 978-0-470-67377-5
Allen PA, Eriksson PG, Alkmim FF, Betts PG, Catunaeunu O, Mazumdar R, Meng Q, Yong GM
(2015) Precambrian basins of India: stratigraphic and tectonic context. In: Mazumdar R,
Eriksson PG (eds) Precambrian basins of India: stratigraphic and tectonic perspective, vol 43.
Geological Society, London, Special Publications, pp 5–28
Audet DM, Fowler AC (1992) A mathematical model for compaction in sedimentary basins.
Geophysical Journal International 110, 577–590
Awati AB, Kalaswad S (1978) Structure of the Kaladgis around Yadwad, Belgaum district
(A study based on Landsat-1). Bulletin of Earth Science 6, 43–47
Baldwin B, Butler CO (1985) Compaction curves. AAPG Bulletin 69, 622–626
Basu A, Bickford ME (2015) An alternate perspective on the opening and closing of the
intracratonic Purana Basins in Peninsular India. Journal of Geological Society of India 85,
5–25
Biswas SK (2008) Petroliferous basins of India—fifty years’ history and perspective. Memoir of
Geological Society of India 66, 159–201
Bjørlykke K (2014) Relationships between depositional environments, burial history and rock
properties. Some principal aspects of diagenetic process in sedimentary basins. Sedimentary
Geology 301, 1–14
Bond GC, Kominz MA (1984) Construction of tectonic subsidence curves for the early Palaeozoic
miogeocline, Southern Canadian Rocky mountains: implications for subsidence mechanisms,
age of breakup, and crustal thinning. Geological Society of America Bulletin 95, 155–173
Bose PK, Sarkar S, Mukhopadhyay S, Saha B, Eriksson P (2008) Precambrian basin-margin fan
deposits: Mesoproterozoic Bagalkot Group, India. Precambrian Research 162, 264–283
Bott MHP, Mithen DP (1983) Mechanisms of graben formation: the Wedge-subsidence
Hypothesis. Tectonophysics 94, 11–22
Bradley DC (2011) Secular trends in the geological record and supercontinental cycles. Earth
Science Reviews 108, 16–33
Brenner RL, McHargue TR (1988) Integrative stratigraphy—concepts and applications. Prentice
Hall, New Jersey, p 419
Cant DJ (1989) Simple equations of sedimentation: applications to sequence stratigraphy. Basin
Research 2, 73–81
Catuneanu O (2006) Principles of sequence stratigraphy. Elsevier, Amsterdam, p 375
Catuneanu O, Martins-Neto MA, Eriksson PG (2005) Precambrian sequence stratigraphy.
Sedimentary Geology 176, 67–95
Cawood PA, Hawkesworth CJ (2014) Earth’s middle age. Geology 42, 503–506
Condie KC, Kroner A (2008) When did plate tectonics begin? evidence from the geologic record.
Geological Society of America, Special Papers 440, 281–291
Interplay Between Tectonics & Eustacy in a Proterozoic … 111
Condie KC, Kroner A (2013) The building blocks of continental crust: evidence from a major
change in tectonic setting of continental growth at the end of the Archean. Gondwana Research
23, 394–402
Condie KC, Pisarevsky SA, Korenga J, Gardoll S (2015) Is the rate of supercontinental assembly
changing with time? Precambrian Research 259, 278–289
Deshpande GG, Pitale UL (2014) Geology of Maharashtra, 2nd edn. Geological Society of India,
Bangalore, p 266
Dey S (2015) Geological history of the Kaladgi–Badami and Bhima basins, south India:
sedimentation in a Proterozoic intracratonic setup. In: Mazumder R, Eriksson PG
(eds) Precambrian basins of India: stratigraphic and tectonic context. Geological Society,
London, Memoir, 43, 283–296
Dey S, Rai AK, Chaki A (2008) Widespread arkose along the northern margin of the Proterozoic
Kaladgi basin, Karnataka: product of uplifted granitic source of K-metasomatism? Journal of
Geological Society of India 71, 79–88
Dey S, Rai AK, Chaki A (2009) Palaeoweathering, composition and tectonics of provenance of the
Proterozoic Intracratonic Kaladgi–Badami basin, Karnataka, southern India: evidence from
sandstone petrography and geochemistry. Journal of Asian Earth Sciences 34, 703–715
Dickinson WR (1974) Plate tectonics and sedimentation. In: WR Dickinson WR (ed) Tectonics
and sedimentation: Society of Economic Paleontologists and Mineralogists Special Publication
22, 1–27
Dula WF Jr (1991) Geometric models of listric normal faults and rollover folds. AAPG Bulletin
75, 1609–1625
Einsele G (2000) Sedimentary basins: evolution, facies and sedimentary budget. Springer-Verlag,
Berlin, p 792
Eriksson PG, Martins-Neto MA, Nelson DR, Aspler LB, Chiarenzelli JR, Catuneanu O, Sarkar S,
Altermann W, de W Rautenback CJ. (2001) An introduction to Precambrian basins: their
characters and genesis. Sedimentary Geology 141–142:1–35
Eriksson PG, Mazumder R, Catuneanu O, Bumby AJ, Ilondo BO (2006) Precambrian continental
freeboard and geological evolution: a time perspective. Earth Science Reviews 79, 165–204
Evans DAD (2013) Reconstructing pre-pangean supercontinents. GSA. Bulletin 125, 1735–1751
Foote RB (1876) Geological features of the south Mharatta country and adjoining districts.
Memoir of Geological Survey of India 12, 70–138
French JE, Heaman LM, Chacko T, Srivastava RK (2008) 1891–1883 Ma southern Bastar–
Cuddapah mafic igneous events, India: a newly recognized large igneous province.
Precambrian Research 160, 308–322
Gallagher K, Lambeck K (1989) Subsidence, sedimentation and sea-level changes in the
Eromanga Basin, Australia. Basin Research 2, 115–131
Geological Survey of India (1981) Geological and mineral map of Karnataka and Goa on 1:0.5
million scale. GSI, Kolkatta
Gibbs AD (1984) Structural evolution of extensional basin margins. Journal of Geological Society,
London 141, 609–620
Gokhale NW, Pujar GS (1989) Bedding plane faults in the Kaladgi rocks, Basidoni, Belgaum
district, Karnataka state. Current Science 58, 1088–1089
Govinda Rajulu BV, Chandrasekhar Gowda MJ (1972) Albitised slates from the Upper Kaladgi
(Precambrian) formations, Lokapur, Bijapur district, Mysore state. Journal of Geological
Society of India 13, 247–261
Haq BU, Hardenbol J, Vail PR (1987) Chronology of fluctuating sea levels since the Triassic (250
million years ago to present). Science 235, 1156–1167
Jadhav PB (1987) Microstructures of quartzites of Yadwad–Lokapur–Bagalkot area, Bijapur
district, North Karanataka. Unpubl. Ph.D. thesis, Pune University, 204pp
Jadhav PB, Kshirsagar LK (2017) Analysis of folding in the Proterozoic sedimentary sequence of
bagalkot group exposed around Yadwad, N Karnataka, India: a study based on remote sensing
techniques and field investigations. Journal of Indian Society of Remote Sensing 45, 247–258
112 S. Patil Pillai and V. S. Kale
Miall AD, Catunenu O, Eiksson PG, Mazumder R (2015) A brief synthesis of Indian Precambrian
Basins: classification and genesis of basin fills. In: Mazumder R, Eriksson PG
(eds) Precambrian basins of India: stratigraphic and tectonic context, vol 43. Geological
Society of London Memoir, pp 339–347
Miller KG, Kominz MA, Browning JV, Wright JD, Mountain GS, Katz ME, Sugarman PJ,
Cramer BS, Christie-Blick N, Pekar SF (2005) The Phanerozoic record of global sea-level
change. Science 310, 1293–1298
Mukherjee S (2014) Atlas of shear zone structures in meso-scale. Springer, Cham, p 124
Mukherjee MK, Das S, Modak K (2016) Basement-cover structural relationships in the Kaladgi
Basin, southwestern India: indications towards a Mesoproterozoic gravity gliding of the cover
along a detached unconformity. Precambrian Research 281, 495–520
Mukhopadhyay S, Choudhuri A, Samanta P, Sarkar S, Bose PK (2014) Were the hydraulic
parameters of Precambrian rivers different? Journal of Asian Earth Sciences 91, 289–297
Nair MM, Raju AV (1987) A remote sensing approach to basinal mapping of the Kaladgi and
Badami sediments of Karnataka state. In: Radhakrishna BP (ed) Purana Basins of Peninsular
India (Middle to Late Proterozoic), vol 6. Memoir of Geological Society of India, pp 375–381
Naqvi SM, Rogers JJW (1987) Precambrian Geology of India Oxford Monographs Geology &
Geophysics. Clarendon Oxford Press, Oxford, p 223
Patil Pillai S (2004) Testimony of sedimentation and structural patterns of the Bagalkot-Simikeri
area on the evolution of the Proterozoic Kaladgi Basin. Unpublished Ph.D. thesis, Pune
University, Pune
Patil Pillai S, Kale VS (2011) Seismites in the Lokapur subgroup of the Proterozoic Kaladgi Basin,
South India: A testimony to syn-sedimentary tectonism. Sedimentary Geology 240, 1–13
Patil Pillai S, Peshwa VV, Nair S, Sharma M, Shukla M, Kale VS (1999) Occurrence of a
manganese-bearing horizon in the Kaladgi basin. Journal of Geological Society of India 53,
201–204
Patil Pillai S, Pande K, Kale VS (2018) Implications of new 40Ar/39Ar age of Mallapur Intrusives
on the chronology and evolution of the Kaladgi Basin, Dharwar Craton, India. Journal of Earth
System Science 127, 32, p 18
Piper JDA (2013) A planetary perspective on Earth’s evolution: lid tectonics before plate tectonics.
Tectonophysics 589, 44–56
Piper JDA (2018) Dominant lid tectonics behavior of continental lithosphere in Precambrian times:
paleomagnetism confirms prolonged quasi-integrity and absence of supercontinental cycles.
Geoscience Frontiers 9, 61–89
Pirajno F, Santosh M (2015) Mantle plumes, supercontinents, intracontinental rifting and mineral
systems. Precambrian Research 259, 243–261
Pratap M, Sharma R, Das SK, Tripathi B (1999) Tectono-stratigraphic styles and Hydrocarbon
prospectivity of Bhima-Kaladgi Basins vis-à-vis Vindhyan and other coeval Proterozoic
Basins. Abstract Field workshop on Integrated evaluation of the Kaladgi & Bhima Basins.
Geological Society of India, Bangalore, 61–63
Radhakrishna BP (1987) Introduction: Purana Basins of Peninsular India (Middle to Late
Proterozoic). Memoir of Geological Society of India 6, i–xv
Radhakrishna BP, Naqvi SM (1986) Precambrian continental crust of India and its evolution.
Journal of Geology, 94, 145–166
Radhakrishna BP, Ramakrishnan M (1988) Archaean_Proterozoic Boundary in India. Journal of
Geological Society of India 32, 263–278
Rajaram M, Anand SP, Erram VC, Shinde BN (2016) Insight into the structure below the Deccan
Trap covered region of Maharashtra, India from geopotential data. In: Mukherjee S, Misra AA,
Calvès G, Nemčok M (eds) Tectonics of the Deccan Large Igneous Province. Geological
Society, London, Special Publications, 445, 219–236
Ramakrishna TS, Chayanulu AYSR (1988) A geophysical reappraisal of the Purana basins of
India. Journal of Geological Society of India 32, 48–60
Ramakrishnan M, Vaidyanadhan R (2010) Geology of India, vol 1. Geological Society of India,
Bangalore, p 556
114 S. Patil Pillai and V. S. Kale
Renne PR, Sprain CJ, Richards MA, Self S, Vanderkluysen L, Pande K (2015) State shift in
Deccan volcanism at the Cretaceous-Paleogene boundary possibly induced by impact. Science
350, 76–78
Roberts NMW, Slagstad T, Viola G (2015) The structural, metamorphic and magmatic evolution
of Mesoproterozoic orogens. Precambrian Research 265, 1–9
Rogers JJW, Santosh M (2004) Continents and Supercontinents. Oxford University Press, New
York, p 292
Rooney AD, Strauss JV, Brandon AD, Macdonald FA (2015) A Cryogenian chronology: two
long-lasting synchronous Neoproterozoic glaciations. Geology 43, 459–462
Saddler PM (1999) The influence of hiatuses on sediment accumulation rates. GeoResearch Forum
5, 15–40
Sambasiva Rao VV, Sreenivas B, Balaram V, Govil PK, Srinivasan R (1999) The nature of the
Archean upper crust as revealed by the geochemistry of the Proterozoic shales of the Kaladgi
basin, Karnataka, South India. Precambrian Research 98, 53–65
Sarkar PK, Soman GR (1985) Heavy mineral assemblages from the Konkan Kaladgi sediments.
In: Bhaskara Rao B (ed) Proceedings of 5th Indian Geological Congress, IIT, Bombay,
pp 19–25
Sclater JG, Christie PAF (1980) Continental stretching: an explanation of the post-Mid Cretaceous
subsidence of the central North Sea Basin. Journal of Geophysical Research 85(B7),
3711–3739
Sharma RS (2009) Cratons and fold belts of India. Springer-Verlag, Berlin, p 304
Sharma M, Nair S, Patil S, Shukla M, Kale VS (1998) Occurrence of tiny digitate stromatolite
[Yelma Digitata Grey, 1984], Yargatti Formation, Bagalkot Group, Kaladgi Basin, Karnataka,
India. Current Science 75, 360–365
Sommerfield CK (2006) On sediment accumulation rates and stratigraphic completeness: lessons
from Holocene ocean margins. Continental Shelf Research 26, 2225–2240
Sridhar M, Chaturvedi AK, Rai AK (2014) Locating new Uranium occurrences by integrated
weighted analysis in Kaladgi basin, Karanataka. Journal of the Geological Society of India 84,
509–512
Sridhar M, Markandeyulu A, Chaturvedi AK (2017) Mapping subtrappean sediments and
delineating structure with the aid of heliborne time domain electromagnetics: Case study from
Kaladgi basin, Karanataka. Journal of Applied Geophysics 136, 9–18
Stephen George T (1999) Sedimentology of the Kaladgi Basin. Abstract volume. Field workshop
on integrated Evaluation of the Kaladgi & Bhima Basins. Geological Society of India,
Bangalore, 13–17
Vail PR, Mitchum RM Jr, Thompson SIII (1977) Seismic stratigraphy and global changes of sea
level, part 3: relative changes of sea level from coastal onlap. In: Payton CE (ed.) Seismic
stratigraphy-Applications to hydrocarbon exploration: American Association of Petroleum
Geologists Memoir 26, 63–81
Valdiya KS (2016) The making of India: geodynamic evolution, 2nd edn. Springer International,
Switzerland, 924pp
Viswanathiah MN (1968) Badami series: a new post-Kaladgi formation of Karnatak state. Bulletin
of Geological Society of India 5, 94–97
Viswanathiah MN (1979) Lithostratigraphy of the Kaladgi and Badami Groups, Karnataka. Indian
Mineralogist 18, 122–132
Viswanathiah MN, Chandrasekhara Gowda MN (1970) Algal stromatolites from Kaladgi
(Precambrian) formations near Alagundi, Bijapur district, Mysore state. Journal of
Geological Society of India 11, 378–385
Viswanathiah MN, Sreedhara Murthy TR (1979) Algal stromatolites from Kaladgi Group around
Bilgi, Bijapur district, Karnataka. Journal of Geological Society of India 20, 1–6
NE-SW Strike-Slip Fault
in the Granitoid from the Margin
of the South East Dharwar Craton,
Degloor, Nanded District, Maharashtra,
India
1 Introduction
Present study emphasizes deformations in the basement granite from the ‘South
East Deccan Volcanic Province’ (SEDVP) around Degloor/Diglur. W dipping
thrusts (Kaplay et al. 2013), steep normal faults, fault planes dipping towards S,
from basalts and deformations from Quaternary deposits (Kaplay et al. 2017a) were
reported from the area around Nanded. Similarly, Kaplay et al. (2017b) studied the
margin of SEDVP around Kinwat lineament, Maharashtra, India, (northern part of
Nanded district). The E-W strike-slip faults (with maximum net-slip of 24 cm) were
identified in the basement granite from the Kinwat region. The zone was designated
as the ‘Western Boundary East Dharwar Craton Strike-Slip Zone’ (WBEDCSZ).
This prompted us to study the structural aspects of another margin of SEDVP with
basement granite near the Degloor region. The study area (Fig. 1) is loca-
ted *135 km SSW of Kaddam fault, the area from where strike-slip faults were
reported from the basement granite (Sangode et al. 2013).
Basement structures affected Deccan tectonics. For example, *N-S Dharwar
trend in the basement was inherited as weak planes in Deccan trap in subsequent
strike-slip faulting around Mumbai (Misra et al. 2014).
Fig. 1 Location map showing deformation zones along SEDVP contact with EDC. SEDVP:
South East Deccan Volcanic Province, EDC: East Dharwar Craton, EDMDZ: East Dharwar
Margin Deformation Zone, and WBEDCDZ: ‘Western Boundary East Dharwar Craton Deformed
Zone’. Right corner inset: Rose diagram- dominant trend of the strike-slip faults from the regions
1–5. Data population: N = 129. (Geology after Banerjee et al. 2012)
Reverse faults, normal faults (step faults), strike-slip faults, brittle shear along the
boundary of veins, minor folds and boudins are observed in the basement rock. In
Fig. 2, a thin quartz vein is faulted in normal manner at as many as 12 locations.
These faults are (sub) parallel to each other and constitute a step-like array with
small throw. The fault plane dips 80° due W. The slip of fault plain varies from 1.6
to 6.5 cm.
Brittle strike-slip faults are numerous in the study area. In some cases, single
veins are strike-slip faulted (Fig. 3a). In Fig. 3b, a pegmatite vein is dextrally
faulted. In Fig. 3c, slightly folded pegmatite vein is slipped possibly with a “reverse
drag” (Mukherjee and Koyi 2009). A medium size pegmatite vein (A) in granitic
country rock (B), at the contact of basic dyke (C) shows reverse slip but no drag
close to the fault plane (Fig. 3d). A pegmatite vein is strike-slip sinistrally faulted
twice in two directions. Sigmoid P-planes decode the former slip, and sharp offset
the later one (Fig. 3e). Brittle shear fractures and fault planes were also observed;
Y-planes and the fault planes in Fig. 4a–d and also in Fig. 5 trend NE, showing
both sinistral and dextral slip. A thin vein, observed on a horizontal plane, is reverse
dragged only at one side of the fault plane. At the other side, no drag is noted
(Fig. 3f). The slip is 17 cm and fault strikes *NE. A rather irregular basic intru-
sion in granitic rock shows faulting. This is a strike-slip fault (Fig. 6a).
In other cases, single veins are multiply faulted (Fig. 6b). Fault ‘A’ strikes N20°
E and slips 9.0 cm; ‘B’ strikes N50°E and slips 5.0 cm; ‘C’ strikes N75°W and
slips 24 cm; and ‘D’ strikes N80°E and slips 28 cm. In Fig. 6c, the
quartzo-feldspathic vein is dextral faulted at three places. The en-echelon joints
Fig. 2 Step faulting in quartz vein (12 faults in vertical section). We tend to think them as a
manifestation of a major fault, since our observations repeat
118 Md. Babar et al.
Fig. 3 a Normal step faults exposed on steeply inclined exposure, at ‘A’ and ‘B’ dip towards E at
25°. Net slip: 5.5 cm both at ‘A’ and ‘B’. b Dextral sheared pegmatite vein, 2.1 cm net slip along
NE trending fault plane. c Joints acted locally as brittle-ductile fault planes (strike N20°E, 3.2 cm
slip), at horizontal outcrop. d Dextral brittle slip. Net slip 12 cm; N45°E (horizontal exposure).
e Sinstral shear (at horizontal outcrop). Strike of the fault is N20°E; net slip is 5.0 cm. f A thin vein
shows bending before it is deformed in an brittle manner (horizontal exposure)
locally acted as fault planes. In Fig. 6d as well, the quartzo-feldspathic vein faulted
along joint planes at three places both dextrally and sinsitrally.
In Fig. 7a, the vein is faulted at four places. At ‘A’ the sense of movement is
sinistral, while at ‘B’, ‘C’ and ‘D’ the sense is dextral. At ‘B’ the vein has not lost
the continuity indicating local ductile deformation. At ‘A’, the vein is bent close to
the fault plane indicating drag.
A quartz vein (Fig. 7b) slips at three places along parallel set of N6°W-S6°E
trending faults. The displacement at ‘A’ is 7 cm, at ‘B’ it is 4.2 cm and at ‘C’ it is
4 cm. In Fig. 7c, curved joints act as fault planes slipping the vein at five places.
One more slip is noted at ‘F’. Some of these fault planes are curved. In some of the
cases, one of the veins is strike-slip faulted (Fig. 7d) while the other parallel vein is
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 119
Fig. 4 Brittle strike slip shear observed on horizontal exposures. Y-plane trends * NE. a, c,
d Sinistral slip. b Dextral slip. Additionally: d secondary quartz along Y-plane
Fig. 5 a Quartzo-feldspathic vein dextral slipped along N40°E striking fault. b Three fault planes
marked on the diagram are vertical, striking N69°E, N85°E and N54°E from bottom to top of the
image also show dextral slip
120 Md. Babar et al.
Fig. 6 a Basic intrusion faulted with 2.2 cm slip. Fault strikes NE. b Multiple faults (at horizontal
outcrop). c Joints acting as fault planes, at horizontal outcrop. Fault strikes N65°E at ‘A’ with 3 cm
slip, at ‘B’: strike: N65°E, slip: 2.0 cm; at ‘C’, strike: N65°E, slip: 3.5 cm. d ‘A’ and ‘B’ are the
joints which acted as fault planes in displacing pegmatite vein (horizontal outcrop). At ‘A’ the fault
strikes N48°E, with 4.0 cm slip, at ‘C’, strike: N78°W, slip: 9.0 cm
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 121
Fig. 7 a Vein displaced at four places, at horizontal outcrop. Strike of the fault at ‘A’: N30°E,
slip: 7.0 cm, at ‘B’ strike: N30°E, slip: 10 cm, at ‘C’ strike: N65°E, slip: 6.5 cm; at ‘D’ strike:
N50°E, slip: 13 cm. b Quartz vein faulted at three places: ‘A’, ‘B’ and ‘C’. Sense of slip: dextral.
c Curvilinear en-echelon joints acting as fault planes, at horizontal outcrop. Slip at ‘A’ is 8.0 cm, at
‘B’ 0.2 cm, at ‘C’ 4 cm, at ‘D’ 2 cm, at ‘E’ 14 cm; and at ‘F’: 3.8 cm. The strike all these small
faults is N35°E. d At ‘A’ the vein shows deformation with 8 cm strike slip. Fault strikes N60°E. At
‘B’ the vein is unaffected by faulting and is open folded (at horizontal outcrop)
undeformed. In other words, in meso-scale one can see termination of fault planes.
In Fig. 8a, the topmost vein (A) faults at four places (locations ‘D’ to ‘G’) along
parallel en-echelon joints while the lower vein (‘B’) is not. The host rock (‘C’)
above and below the vein ‘A’ is also faulted. In some other cases, parallel veins
faulted (Fig. 8b). Here, left vein shears sinistrally and the right vein oppositely and
dextrally. In Fig. 8c, however, both the veins shear dextrally.
122 Md. Babar et al.
Fig. 8 a En-echelon joints (marked as ‘D’, ‘E’, ‘F’ and ‘G’) displaced the vein in brittle and
ductile-brittle manner (at horizontal outcrop). At ‘D’ slip: 2.0 cm, at ‘E’: 0.6 cm, at ‘F’ the vein is
ductile deformed, at ‘G’ slip is 0.5 cm. These faults strike *NE. b brittle deformation (at
horizontal outcrop). Strike of the fault at ‘A’ is N40°W with 26 cm slip; at ‘B’ strike: N70°E, slip
14.0 cm. c Two parallel veins brittle deformed. At ‘a’ 2 cm strike slip, at ‘B’ 1.5 cm slip. Fault
strikes N40°E. d Multilayered faulting
Several layers in the rock got faulted (Fig. 8d). Two conjugate joints, i and ii, cut
through the layers. Joint i does not slip the layer, however, joint ii does dextrally.
Therefore joint ii can be called more accurately a fault. The strike of the fault is
N30°E. The displacement for layer ‘A’ is 0.8 cm, for ‘B’ 0.9 cm, for ‘C’ 1.1 cm, for
‘D’ 1.2 cm, and for ‘E’ 1.4 cm. Slip increases locally towards SW. Commonly fault
planes demonstrate maximum slip near their central part and at the margins mini-
mum or zero (review in Mukherjee 2014a). Figure 8d is an exception.
Conjugate strike-slip faults at *30° are observed in this study at places
(Figs. 6d, 9a, b and 10). Fault ‘A’ strikes N50°E and slips 7 cm. Fault ‘B’ strikes
N80°E and slips 6 cm. Spectacular conjugate multilayer strike-slip faulting is
observed in the ancient granitoid rock (Fig. 10). On a *N striking fault,
brittle-ductile slip of 3 cm is noted at ‘A’, 5 cm at ‘B’, and 1.2 cm at ‘C’. On a
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 123
N30°E striking plane, slip at ‘A/’ is 1 cm, at ‘B/’ is 1.1 cm, and at ‘C/’ is 1 cm. The
thickest pegmatite layer is sheared sinistrally. Note fracture patterns inside the vein.
That shear acted prior to the conjugate faulting. The left hand side geographic
direction in this snap is N30°W/330°. A pegmatite vein faults along two
sub-parallel planes. Since the angle between the two fault planes is rather low, they
are not conjugate faults. Fault ‘A’ strikes N25°W and slips 2 cm. Fault ‘B’ strikes
N6°W and slips 4 cm.
‘Intrusion-in-intrusion’ feature is ductile sheared (Fig. 11). Basic intrusion (B),
with quartz vein (C) within, inside host granitic rock (A), is ductile sheared along
N20°W-S20°E. Both the basic intrusion and the quartz vein thin significantly inside
the shear zone (Fig. 11). Outside and inside the shear zone, ‘B’ is 22 and 6 cm
thick, respectively. Likewise, ‘C’ is 6 and 1 cm thick.
Though major/regional folds are absent, small (ptygmatic) folds are found in
vertical and horizontal sections. Synclinal type of fold is observed at one place
(Fig. 12a). The axial plane is slightly inclined. A folded structure, at the horizontal
outcrop, is exposed in granitic gneiss (Fig. 12b).
Folding in quartzo-feldspathic vein/pegmatite veins is also a common feature
observed in the study area (Fig. 12c). At many places, pegmatite veins intruded in
host granitoids is ptygmatic folded (Fig. 12d; also see Mukherjee et al. 2015).
Drawing tectonic inferences from ptygmatic folding would be difficult because of
their irregular and sometimes inconsistent geometries. ‘S’ and M-type folding in
Fig. 11 Quartz vein (C) within basic intrusion (B) in host granitoid (A) ductile deformed. Shear
zone is marked by red lines (at horizontal outcrop)
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 125
Fig. 12 a Asymmetric folds at vertical outcrop. b Folding in granitic gneiss (horizontal outcrop).
Axial plane trends N. c Folding in pegmatite vein (horizontal outcrop). d Ptygmyatic folding.
(E-W trending photo). e Pegmatite vein folded (W-E trending photo). f Folding in pegmatite vein
(horizontal outcrop)
pegmatite dyke in granitoid (Figs. 12e, f and 13a, b) do exist though we do not
decipher any regional folding. ‘Z’-type tight fold with hinge thicker than the limbs
is observed in one of the pegmatite veins. Multiple (faint) faults also occur in the
close vicinity (Fig. 13c). Pinch and swell structures of pegmatite layer indicates
local brittle-ductile NW-SE extension. Not all swells resemble geometrically, while
few are sub-elliptical.
There are two major trends of vertical joint planes observed. Trend-1 varies
N10° to 340°NW and trend-2 260°SW to 280°NW (Fig. 14).
126 Md. Babar et al.
Fig. 13 a Folding in pegmatite vein. b ‘M’ type of folded pegmatite vein. c ‘Z’ type folding with
multiple faulting in its close vicinity (W-E trending photo)
Fig. 14 Mechanism of strike-slip conjugate faulting. Strike of the fault is N50°E. Red dash lines:
actual fault lines, black solid lines intersecting shear axes. r1: short solid black arrow, r3,
compressive stress axis: open arrow
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 127
3 Discussions
Table 1 (continued)
Structure No. Deformation regime Fault type Shear sense Fault trend Net slip (cm)
F41 Brittle Strike slip Sinistral N70°E 1.1
F42 Brittle Strike slip Sinistral N60°E 08
F43 Brittle Strike slip Dextral N65°E 03
F44 Brittle Strike slip Dextral N65°E 02
F45 Brittle Strike slip Dextral N65°E 3.5
F46 Brittle Strike slip Dextral N45°E 09
F47 Brittle Strike slip Dextral N85°E 3.0
F48 Brittle Strike slip Dextral N85°E 3.5
F49 Brittle Strike slip Dextral N20°E 7.5
F50 Brittle Strike slip Sinistral N35°W 03
F51 Brittle Strike slip Sinistral N40°E 04
F52 Brittle Strike slip Dextral N20°E 3.2
F53 Brittle Strike slip Dextral N47°E 02
F54 Brittle Strike slip Dextral N48°E 04
F55 Brittle Strike slip Sinistral N80°W 09
F56 Brittle Strike slip Dextral N30°E 07
F57 Brittle Strike slip Dextral N30°E 10
F58 Brittle Strike slip Dextral N65°E 6.5
F59 Brittle Strike slip Dextral N50°E 13
F60 Brittle Strike slip Dextral N45°E 14
F61 Brittle Strike slip Dextral N70°E 26
F62 Brittle Strike slip Sinistral N40°W 09
F63 Brittle Normal – – 12.5
F64 Brittle Strike slip Dextral N20°E 02
F65 Brittle Strike slip Dextral N50°E 02
F66 Brittle Strike slip Dextral N50°E 05
F67 Brittle Strike slip Sinistral N75°W 24
F68 Brittle Strike slip Sinistral N80°E 28
F69 Brittle Strike slip Sinistral N39°E 02
F70 Brittle Strike slip Dextral N37°E 05
F71 Brittle Strike slip Dextral N40°E 2.2
F72 Brittle Strike slip Dextral N40°E 6.0
F73 Brittle Strike slip Sinistral N40°E 1.0
F74 Brittle Strike slip Sinistral N40°E 1.5
F75 Brittle Strike slip Dextral N20°E 18
F76 Brittle Strike slip Dextral N45°E 11
F77 Brittle Strike slip Dextral N45°E 08
F78 Brittle Strike slip Dextral N43°E 6.5
F79 Brittle Normal – – 4.5
F80 Brittle Normal – – 1.6
(continued)
130 Md. Babar et al.
Table 1 (continued)
Structure No. Deformation regime Fault type Shear sense Fault trend Net slip (cm)
F81 Brittle Normal – – 2.0
F82 Brittle Normal – – 6.5
F83 Brittle Normal – – 3.0
F84 Brittle Normal – – 6.0
F85 Brittle Normal – – 4.0
F86 Brittle Normal – – 1.6
F87 Brittle Normal – – 2.0
F88 Brittle Normal – – 4.0
F89 Brittle Normal – – 3.5
F90 Brittle Strike slip Dextral N35°E 08
F91 Brittle Strike slip Dextral N35°E 04
F92 Brittle Strike slip Dextral N35°E 02
F93 Brittle Strike slip Dextral N35°E 14
F94 Brittle Strike slip Sinistral N80°E 12
F95 Brittle Strike slip Sinistral N80°E 19
F96 Brittle Strike slip Sinistral N20°E 15
F97 Brittle Strike slip Sinistral N20°E 22
F98 Brittle Strike slip Dextral N20°E 22
F99 Brittle Strike slip Dextral N24°E 04
F100 Brittle Strike slip Dextral N24°E 04
F101 Brittle Strike slip Dextral N30°E 12
F102 Brittle Strike slip Dextral N30°E 11
F103 Brittle Strike slip Dextral N30°E 06
F104 Brittle Strike slip Sinistral E-W 07
F105 Brittle Strike slip Dextral N35°E 6.5
F106 Brittle Strike slip Dextral N35°E 09
F107 Brittle Strike slip Dextral N26°E 05
F108 Brittle Strike slip Sinistral N20°E 05
F109 Brittle Thrust – – 13
F110 Brittle Strike slip Dextral N40°E 1.5
F111 Brittle Strike slip Dextral N40°E 2.0
F112 Brittle Strike slip Dextral N05°W 10
F113 Brittle Strike slip Dextral N20°W 1.0
F114 Brittle Strike slip Dextral N60°E 07
F115 Brittle Strike slip Sinistral N30°E 06
F116 Brittle Strike slip Sinistral N20°W 82
F117 Brittle Strike slip Dextral N40°W 10
F118 Brittle Strike slip Dextral E-W 05
F119 Brittle Strike slip Dextral E-W 05
F120 Brittle Strike slip Dextral E-W 03
(continued)
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 131
Table 1 (continued)
Structure No. Deformation regime Fault type Shear sense Fault trend Net slip (cm)
F121 Brittle Strike slip Dextral N25°E 04
F122 Brittle Strike slip Dextral N40°E 05
F123 Brittle Strike slip Dextral N30°E 2.5
F124 Brittle Strike slip Dextral N30°E 07
F125 Brittle Strike slip Dextral N30°E 4.0
F126 Brittle Strike slip Dextral N30°E 2.0
F127 Brittle Strike slip Dextral N30°E 1.5
F128 Brittle Strike slip Dextral N30°E 1.0
F129 Brittle Strike slip Dextral N30°E 3.5
which is located along the SE of Kinwat region. This NE-SW trend of strike-slip
faults from Dharwar Craton near Degloor mismatches with the trend of the N-S
trending strike-slip faults near Mumbai reported by Misra et al (2014) (also see
Misra and Mukherjee 2015). The easternmost fault reported at Malshej Ghat (Misra
et al. 2014) is *500 km west of Degloor. These faults cannot be related with the
study area presumably because they are far away from the study area.
The Degloor region is bound at SW by seismically active Killari region, NW by
the present day microseismically active Nanded and NE by seismic event nearer to
Kaddam fault/lineament reported in 2015 (Kaplay et al. 2017a, b). The study area
locates exactly at the edge of one of the side of a seismically active triangle i.e.,
Killari-Kaddam-Nanded. Killari seismically activated since 1993, Nanded since
2006 and a small-scale seismicity is reported recently in 2015 nearer to Kaddam
fault/lineament (Kaplay et al. 2017a, b). So far no (micro)seismic events are
reported from Degloor. The region Killari and Nanded in Maharashtra and Kaddam
in the Indian state Telangana lies in the zone III of Indian map of seismic zonation
(Subhadra et al. 2015). Killari is a typical example of intracratonic seismicity and a
‘slow-deforming non-rifted zone’ (Rajendran 2016). Nanded and Kaddam too
exemplify intracratonic microseismicity. The studies in Killari region suggested that
Killari had experienced similar earthquakes in past (Rajendran et al. 1996). The
Nanded region also experienced earthquakes in 1942 (review in Valdiya 2015).
Killari and Nanded seismicity occurred in the SEDVP, which was previously
considered tectonically stable. The Kaddam event occurred in granitic province.
The E-W strike-slip deformation zone is reported at the SEDVP contact around
Kinwat (Kaplay et al. 2017a). A stretch of 60 km (Adampur-Bhaisa) SW of Kinwat
along SEDVP contact has been recently designated as the “East Dharwar Margin
Deformation Zone” (EDMDZ; (Kaplay et al., submitted). The reporting of
strike-slip faults from Degloor, 15 km SW of Adampur-Bhaisa stretch, is marked as
the EDMDZ in Fig. 1 confirms that the stretch EDMDZ continued further SW up to
Degloor and hence this zone is designated as ‘Western Boundary East Dharwar
Craton Deformed Zone’(WBEDCDZ) which includes entire EDMDZ. The zone
132 Md. Babar et al.
might continue SW right down up to Killari and also towards NE from Bhisa to
Kinwat, but is to be checked by fieldwork. The study area is bound towards SE by
seismically active Killari, towards NW by microseismically active Nanded and
towards NE by recently reported microseismic event near Kaddam.
4 Conclusions
A. Deformation style, *NE-SW structures, at Degloor does not match with those
at the South East Deccan Volcanic Province (SEDVP), with Dharwar craton at
Kinwat with *E-W trend and the NW-SE strike-slip faults at Kaddam.
Therefore, the stretch of *105 km from Degloor to Bhainsa is designated
separately as the ‘Western Boundary East Dharwar Craton Deformed Zone’
(WBEDCDZ).
B. Orientation of stress axes deciphered from this study: Fig. 6d: r3 (N83°E), r1
(N8°W/352°). Figure 9a: r3 (N65°E), r1 (N25°W/335°). Figure 9b: r3 (N20°
W/340°), r1 (N66°E/66°). Figure 10: r3 (N°55E/°), r1 (N40°W/320°). This
shows the extent of local variation of stress regime. This work presents the first
detail information on orientation of stress axes from the remobilized EDC.
Future structural geological works in the same line can provide comparison of
how stress pattern had varied in the past. A statistical model highlighting the
probability of earthquake occurrence can better justify the suggestion of
installing a seismic network along the SEDVP.
Acknowledgements D. Kaplay thanks the School of Earth Sciences, S.R.T.M. University. Dr.
Babar thanks the Principal, Dnyanopasak College, Parbhani, Maharashtra for support and
encouragement. S. Mukherjee was supported by the CPDA grant of IIT Bombay. A research
sabbatical provided by IIT Bombay helped him to finish this work. Fieldwork by S. Mahato was
supported partially by IIT Bombay. The anonymous reviewers provided several positive inputs in
two rounds. Mukherjee (2019) summarizes this work.
References
Babar MD, Kaplay RD, Mukherjee S and Kulkarni PS (2017) Evidences of deformation of dykes
from Central Deccan Volcanic Province, Aurangabad, Maharashtra, India. In: Mukherjee S,
Misra AA, Calvès G, Nemčok M (eds) Tectonics of the Deccan large igneous province.
Geological Society, London, pp 337–353 (Special Publications 445)
Banerjee R, Veena K, Pandey BK, Parthasarathy TN (1993) Rb-Sr geochronology of the
radioactive granites of Nanded area, Maharashtra, India. Journal of Atomic Mineral Science 1,
111–117
Banerjee R, Jain SK, Shivkumar K (2008) Geochemistry and petrogenesis of uraninite bearing
granitoids and radioactive phosphatic cherty cataclasite of Thadisaoli Area, Nanded District,
Maharashtra. Memoir of Geological Society of India 73, 55–84
NE-SW Strike-Slip Fault in the Granitoid from the Margin … 133
Rajendran CP, Ranjedran K, John B (1996) The 1993 Killari (Latur), Central India, Earthquake: an
example of fault reactivation in the Precambrian Crust. Geology 24, 651–654
Sangode SJ, Meshram DC, Kulkarni YR, Gudadhe SS, Malpe DB, Herlekar MA (2013)
Neotectonic response of the Godavari and Kaddam Rivers in Andhra Pradesh, India:
implications to quaternary reactivation of old fracture system. Journal Geological Society of
India 81, 459–471
Subhadra N Padhy S, Prabhakara PP, Seshunarayana T (2015) Site-specific ground motion
simulation and seismic response analysis for microzonation of Nanded City, India. Natural
Hazards 78. https://doi.org/10.1007/s11069-015-1749-z
Valdiya KS (2015) The making of India: geodynamic evolution. Springer. ISBN 978-3-319-
25029-8
Wesanekar PR, Patil RR (2000) Rb-Sr dating of pink Granites of Deglur, Nanded district,
Maharashtra, India. In: Proceedings of National seminar on tectonomagmatism, geochemistry
and metamorphism of Precambrian Terrain, pp 181–187
Synthesis of the Tectonic and Structural
Elements of the Bengal Basin and Its
Surroundings
1 Introduction
The Bengal Basin is one of the largest peripheral collisional foreland basins in
South Asia (Mukherjee et al. 2009; DeCelles 2012; DeCelles et al. 2014) consisting
of Permo-carboniferous to Mesozoic and Tertiary deposits covered by the Recent
alluvium. This is one of the thickest sedimentary basins of the world consisting of
*21 km thick Early Cretaceous–Holocene sedimentary succession (Curray 1991;
Curray and Munasinghe 1991). Geographically, the Bengal Basin lies approxi-
mately between 20° 34′ to 26° 40′N and 87° 00′ to 92° 45′E with its major portion
being constituted by Bangladesh and also covering parts of the Indian States of
West Bengal, Tripura and Assam (Fig. 1). To the north, the basin passes into the
Tertiary Shelf of the Assam Basin, and to the south, it continues as the Bengal Fan
(Acharyya 2007). The Bengal Basin is bordered on the west by the Indian Shield
and the Shillong Plateau to the north, the Indo-Burman Ranges to the east and the
Bay of Bengal to the south (Uddin and Lundberg 2004).
The basin occupied dominantly by the Ganges-Brahmaputra-Meghna
(GBM) delta named after the Ganges (Padma in Bangladesh), the Brahmaputra
(Jamuna in Bangladesh) and the Meghna rivers, constituting the largest
fluvio-deltaic to shallow marine sedimentary basin of the world. The Ganges enters
the basin from the northwest, the Brahmaputra from the northeast, whereas the
Meghna drains the Sylhet Trough and part of the Tripura hills before flowing into
the Brahmaputra and finally into the Bay of Bengal. Cumulative sediment dispersal
through the Bengal Basin into the Bay of Bengal per year is estimated to about 1.1
Gigatonne (GT), results the formation of the biggest submarine fan of the world, the
Fig. 1 Simplified map of the geotectonic provinces of the Bengal Basin and its surroundings.
Solid blue lines indicate the cross-sections position. The boundaries of the three geotectonic
provinces are taken from Bakhtine (1966), Guha (1978) and Matin et al. (1983), and the domain
boundaries are taken from Wang et al. (2014). Faults and lineaments are mostly taken from Kayal
(2008) and Wang et al. (2014). MBT: Main Boundary Thrust; MFT: Main Frontal Thrust; MT:
Mishmi Thrust and T.: Thrust
Bengal Fan (Kuehl et al. 1989; Milliman et al. 1995; Curray et al. 2003; Goodbred
et al. 2003; Syvitski et al. 2005; Curray 2014; Reitz et al. 2015). The GBM delta
receives most of the sediments shed from the Himalaya and is subjected to tectonic
Synthesis of the Tectonic and Structural Elements … 137
influence from the Himalayan front and the Shillong Plateau uplift in the north and
the accretionary fold belt in the east (Reitz et al. 2015). The Bengal Basin traps and
accumulates less than half of the total sediment budget through flexural subsidence
over a broad area, in addition to faulting, folding and localized compaction
(Goodbred and Kuehl 1998, 1999). This accumulated sediment has caused the
entire Bengal Basin margin to prograde more than 300 km since the Eocene
(35 Ma) (Najman et al. 2008; Reitz et al. 2015).
The geotectonics of the eastern part of the Indian Plate is dominantly influenced
by its collision with the Eurasian Plate (Mukherjee 2013a, 2015; Mukherjee et al.
2013, 2015) to the north and the Burmese Plate to the east, uplifting the Himalaya
and the Indo-Burman Ranges (IBR) in the north and east, respectively (Alam et al.
2003; Vernant et al. 2014; Wang et al. 2014; Hossain et al. 2016). The Himalayan
mountain range started to form in the Mid-Late Eocene (Garzanti et al. 1987;
Klootwijk et al. 1992; Searle et al. 1997; Ali and Aitchison 2008; Baxter et al.
2016), whereas the IBR began to form approximately in the Middle Miocene
(Acharyya 2007; Steckler et al. 2008). The Bengal Basin at the eastern part of the
Indian Plate is tectonically active since the collision began (Eocene). Although the
direction of subduction of the Indian Plate below the Eurasian Plate is *N-S,
whereas below the Burmese Plate, it is NE–SW (Goodbred et al. 2003; Reitz et al.
2015). In the north of the Bengal-Assam Basin, *E trending Himalaya takes a
southward turn towards Myanmar and connects with the IBR (Hossain et al. 2016).
The geodetic measurements show that the motion of the Indian Plate relative to the
Burmese Plate is *36 mm/year (Socquet et al. 2006; Steckler et al. 2016) and the
convergence between the Shillong Plateau and the Bengal Basin across the Dauki
Fault varies from 3 mm/year in the west to >8 mm/year in the east (Socquet et al.
2006; Vernant et al. 2014). Plate boundary deformation along these collisional belts
is broadly distributed over a series of reverse and strike-slip structures, including
the Dauki and Haflong-Disang Faults to the north and Chittagong-Teknaf, Kaladan
and Sagaing Faults to the east of the Bengal Basin, respectively (Hossain et al.
2014, 2016).
The Bengal Basin rests on lithosphere that is transitional (along the
paleo-continental slope also known as the Eocene Hinge Zone) between thick,
buoyant Indian continental lithosphere in the west and north and dense Indian
oceanic lithosphere in the east (Bender 1983; Curray 1991; Reimann 1993; Uddin
and Lundberg 2004; Singh et al. 2016). Along its northern edge, it is subjected to
continent-continent collision whereas along eastern and southeastern edge, it is
subjected to ocean-continent subduction. According to Ingersoll et al. (1995), due
to continued oblique subduction of the Indian Plate beneath the Burmese Plate to
the SE, the Bengal Basin turned into a Remnant Ocean Basin at the beginning of the
Miocene. Due to eastward component of subduction of the oceanic part of the
Indian Plate, the thick pile of sediments of the Bengal Basin has been deformed into
a fold belt and a huge flat accretionary prism (Alam et al. 2003; Steckler et al.
2008). In general, by considering the overall regional tectonic setting, the Bengal
Basin as a remnant ocean basin can be divided into three major geotectonic pro-
vinces: (i) the stable shelf to the northwest—passive to extensional cratonic margin
138 M. S. Hossain et al.
in the west, (ii) the deeper foredeep basin at the center—remnant ocean basin, and
(iii) fold belt to the east—the Chittagong Tripura Fold Belt (CTFB) (Bakhtine 1966;
Alam 1972; Khan and Rahman 1992; Khan and Agarwal 1993; Reimann 1993;
Shamsuddin and Abdullah 1997; Alam et al. 2003).
The northwest Stable Shelf of the Bengal Basin consists of an easterly dipping
shelf that is separated from the Precambrian Indian Shield to the west by a
prominent fault zone (Basin Margin Fault Zone: more than 350 km in length and
width is variable in hundreds of metres) with dislocation and cataclasis (Raman
et al. 1986; Matin and Misra 2009; Roy 2014). To the southeast it is bounded by the
Eocene Hinge Zone that forms the paleo-continental slope (Sengupta 1966; Guha
1978; Bender 1983; Rahman et al. 1990a; Reimann 1993; Uddin and Lundberg
2004). Only a few studies have been performed on the basement rocks and their
possible origin in the Stable Shelf part of the Bengal Basin (e.g., Ameen et al. 2007;
Hossain et al. 2007, 2017; Ameen et al. 2016). The deeper Foredeep Basin lies in
between the Eocene Hinge Zone to the west and CTFB to the east. The northern
Boundary of the Foredeep Basin is marked by the Shillong Plateau consisting of
Garo-Khasi-Janitia-Mikir Hills of Assam, India with *2000 m elevation. In con-
trast, immediate south of the plateau, the Foredeep Basin contains up to 20 km
thick pile of sediments. The Shillong Plateau and the Foredeep Basin are separated
by the E-W trending Dauki Fault (Bilham and England 2001; Biswas and
Grasemann 2005a; Najman et al. 2016). This Foredeep Basin has been studied by
Holtrop and Keizer (1970), Woodside (1983), Hiller and Elahi (1984), Murphy and
Staff BOGMC (1988), Shamsuddin (1989), Johnson and Alam (1991), Mannan
(2002), Alam et al. (2003), Rahman and Suzuki (2007), Najman et al. (2012),
Farhaduzzaman et al. (2014) and Rahman and Worden (2016) to understand tec-
tonic evolution of sedimentary basin architecture and provenance of the Neogene
sediments.
Since 1914, the CTFB has been drawing attention to both structural and
exploration geologists because of numerous earthquake epicenters and structural
complexities including major thrust fault, and mud volcanoes (Bakhtine 1966;
Hossain and Akhter 1983; Hossain 1985; Sikder 1998; Chowdhury et al. 2003,
Steckler et al. 2008; Maurin and Rangin 2009a; Das et al. 2010; Wang et al. 2014;
Steckler et al. 2016). Moreover, presence of several anticlinal folds, numerous
faults and associated gas seepages attracted several national and international oil &
gas companies to conduct seismic survey and drilling in the CTFB area
(Shamsuddin and Abdullah 1997; Racey and Ridd 2015). Orogenic development
and morphotectonic elements of the CTFB and IBR have been studied extensively
by Bakhtine (1966), Gilbert (2001), Acharyya (2007), Richards et al. (2007),
Bhattacharya et al. (2008), Das et al. (2011a) and Hossain et al. (2014). In addition,
numerous investigations have been carried out in this area focusing mainly on 2D
structural modeling (Sikder and Alam 2003; Mandal et al. 2004; Kabir and Hossain
2009; Maurin and Rangin 2009a; Najman et al. 2012; Abdullah et al. 2015),
geochemistry and provenance studies (Najman et al. 2008; Dina et al. 2016;
Rahman et al. 2017), active tectonics (Goodbred and Kuehl 1999; Khan et al. 2005;
Wang et al. 2014; Khan et al. 2015), and earthquake and landslide hazard studies
Synthesis of the Tectonic and Structural Elements … 139
(Ansary et al. 2000; Cummins 2007; Steckler et al. 2008; Wang et al. 2014;
Steckler et al. 2016). Based on spatial deformation intensity and geometry of the
folds, Bakhtine (1966) subdivided the CTFB into three zones: eastern highly
compressed disturbed zone, middle asymmetric thrust faulted zone, and western
quiet zone. Chowdhury (1970), Hossain and Akhter (1983), and Hossain (1985)
performed single-fold geometric analysis on a few anticlines in the area. No
comprehensive analysis of the fold geometric features has been carried out for a
large population of folds till date. As a result, there are still many uncertainties
concerning the original geometric shape as well as spatial distribution of the fold
deformation intensity within the CTFB.
The first stratigraphic scheme of the Bengal Basin was originally proposed by
Evans (1932) based purely on the exposures found in the CTFB and their simple
long distance lithostratigrahic correlation with type sections in Assam. Over the
years, several researchers have attempted to improve this scheme based on
micro-paleontological studies (Ahmed 1968; Ismail 1978), palynological studies
(e.g., Chowdhury 1982; Uddin and Ahmed 1989; Reimann 1993), and
seismo-stratigraphic studies (Lietz and Kabir 1982; Salt et al. 1986; Lindsay et al.
1991; Alam et al. 2003), and consequently, abetted in solving some of the strati-
graphic problems of the Bengal Basin. Three stratigraphic schemes for the three
respective geotectonic provinces of the Bengal Basin have been proposed (Bakhtine
1966; Guha 1978; Khan 1991a, b; Alam et al. 2003) and these are: (i) Geotectonic
Province 1—Stable Shelf (Zaher and Rahman 1980; Lindsay et al. 1991; Reimann
1993; Alam 1997; BOGMC 1997); (ii) Geotectonic Province 2—Central Deep
Foredeep Basin (Hiller and Elahi 1988); and (iii) Geotectonic Province 3—CTFB
(Fig. 1) (Gani and Alam 1999).
Deposition of sediments within the Bengal Basin is believed to be controlled by
tectonic cycles or stages related to the interaction and collision patterns of major
plates, and has taken place in five distinct phases (Alam et al. 2003): (1) Syn-rift
stage: Permo-Carboniferous to Early Cretaceous (Sclater and Fisher 1974; Molnar
and Tapponnier 1975; Uddin and Lundberg 2004); (2) Drifting stage: Cretaceous–
Mid-Eocene (Banerji 1981; Ramana et al. 1994); (3) Early collision stage:
Mid-Eocene–Early Miocene (Bender 1983; Uddin and Lundberg 1998, 2004);
(4) Late Collision stage: Early Miocene–Mid-Pliocene (Ingersoll et al. 1995); and
(5) Latest collision stage: Mid-Pliocene–Quaternary (Alam et al. 2003; Najman
et al. 2016). However, proper correlation of the stratigraphic succession in the three
tectonic provinces and with other adjacent foreland basins is yet to be done.
Based on the above discussion, it can be concluded that there is no compre-
hensive study that synthesizes the tectonic evolution and structural elements of the
Bengal Basin as a whole including its adjacent areas. Therefore, this chapter is
aimed at understanding the tectonic evolution, structural framework, and strati-
graphic succession of the Bengal Basin. The work will not only enhance the
understanding of present day structural and tectonic elements of the Bengal Basin
but also help to comprehend their evolution with time.
140 M. S. Hossain et al.
The Bengal Basin is bounded on all sides by different tectonics and structural
controlled features which evolved at different geological times between the Late
Cretaceous and Late Neogene, except in the south where the basin is open to the
Bay of Bengal (Fig. 1) (Alam et al. 2003; Ali and Aitchison 2008; Roy and
Chatterjee 2015; Baxter et al. 2016). The southeastern boundary of the Bengal
Basin coincides with the eastern margin of the Chittagong Tripura Fold Belt
(CTFB) bordering the western fringe of the Indo-Burman Ranges (IBR) (Acharyya
2007; Hossain et al. 2014; Wang et al. 2014). In the north, the basin is bounded by
the Himalayan Foredeep, the Shillong Plateau and the Assam Basin (Roy and
Chatterjee 2015; Najman et al. 2016). In the NNW corner, the Bengal Basin extends
up to the Major/Main Frontal Thrust (MFT), which separates the Sub-Himalayan
Zone from the Indo-Gangetic Alluvial Plain (Roy and Chatterjee 2015). The
western and southwestern parts of the Bengal Basin consist of an easterly inclined
shelf, separated from the Singhbhum Craton of the northeastern part of the Indian
Shield (Acharyya 2003; Hossain et al. 2007, 2017; Kumar et al. 2017). The
southern boundary of the basin is marked by the northern edge of the Bengal Fan
(Fig. 2) (Curray et al. 2003). A generalized description of these bordered tectonic
features is summarized below.
Fig. 2 Gravity mosaic of the Bengal Fan (Bay of Bengal) obtained from the satellite derived data
(modified after Desa et al. 2013). Thick dashed lines mark the extent of the subsurface 85°E Ridge
inferred using the satellite gravity mosaic. Blue dashed lines indicate the likely hotspots tracks of
the 85°E and 90ºE Ridges. Fracture zones are marked by the thin dashed black lines. Thin blue line
outlines the large igneous provinces. ANI: Andaman-Nicobar Islands; BD: Bangladesh; IN: India;
MN: Myanmar; SL: Sri Lanka; SoNG: Swatch of No Ground; ST: Sunda Trench
142 M. S. Hossain et al.
Wang et al. (2014) divide the IBR and the CTFB into three domains along
north-south direction, and these are the Ramree domain to the south, the Dhaka
domain in the middle, and the Naga domain to the north (Fig. 1). The Ramree
domain sustained oblique subduction to nearly orthogonal subduction and accretion
along the 450 km long plate margin. This subduction produced a belt of defor-
mation that increased not only the width from 170 km in the south to 250 km in the
north, but also the height of the IBR from <1000 m in the south to >2000 m in the
north. In the middle Dhaka domain, plate margin sustained collision, resulting in
the formation and rapid westward propagation of a great fold belt and low angle
subduction mega thrust within the sediments of the eastern side of the Ganges–
Brahmaputra-Meghna (GBM) Delta (Maurin and Rangin 2009a). The total length
of domain is *500 km and the maximum width along 23°N is about 400 km. The
height of the IBR in this domain increases towards N, and attains *3000 m at
21.3°N.
The CTFB lies to the western margin of the Dhaka domain swing eastward to its
northern end. The approximately east-west oriented Dauki Fault (Biswas et al.
2007; Hossain et al. 2016) marks the northern end of the Dhaka domain (beyond
which CTFB is absent), and the beginning of the Naga domain. A major change in
the collisional characteristic occurs at the beginning of the Naga domain approxi-
mately at 25°N. The Naga domain consisting of the Naga Hills (northern edge of
the Burmese Plate) overrides the narrow continental shelf of the Assam Basin to the
northeast (Wang et al. 2014). The significant geomorphologic difference between
the Dhaka and the Naga domains seems to reflect the contrast in collision at shallow
crustal level rather deep-seated ones. The length of the Naga domain is *430 km
and the width of the domain narrows from *170 km in the southwest to *90 km
to the northeast, suggesting a northeastward depletion in total shortening across the
hill range (Maurin and Rangin 2009a; Wang et al. 2014).
The Bengal Basin in the northeast is bounded by the Shillong Plateau, which was
uplifted during the Pliocene (Acharyya et al. 1986; Biswas et al. 2007), and sep-
arated the Assam Basin from the Bengal Basin at the juncture of the southwestern
part of the Naga domain (Fig. 1). Approximately E-W stretching Shillong Plateau
raised *2000 m elevation as continental basement pop-up structure due to two
seismically active steep reverse faults along its southern and northern edges. These
two prominent faults are the E-W trending Dauki Fault in the south and the
WNW-ESE trending Oldham Fault in the north (Fig. 1) (Bilham and England 2001;
Biswas and Grasemann 2005a; Kayal et al. 2006; Kayal 2008). South-dipping
Oldham Fault is thought to be the backthrust of the north dipping master Dauki
Fault (Yin et al. 2010; see Mukherjee 2013b for backthrust kinematics in general,
Synthesis of the Tectonic and Structural Elements … 143
also see Bose and Mukherjee, submitted-a, b). The importance of the Oldham Fault
is considered to be minor compared to the Dauki Fault for the plateau uplift (Biswas
et al. 2007). However, another school of thought believes the Shillong Plateau to be
a large anticlinorium floored by Archean and Proterozoic basement rocks with
remnants of the passive margin strata on its steep southern face adjacent to the
Surma basin (Singh et al. 2016).
The plateau is separated from the Mikir Massif to the east by the NW-SE
trending Kopili Fault (also known as the Kopali Fracture Zone) (Evans 1964;
Biswas and Grasemann 2005a; Biswas et al. 2007; Angelier and Baruah 2009). In
the west, the plateau abuts against the Bengal Basin along a prominent fault, which
continues northward up to the Himalayan front. Up to the point of inflection of the
sharply bent Brahmaputra River course from westward to southward, the fault is
known as the Jamuna Fault and to the further north as the Dhubri Fault (Desikachar
1974; Nandy 2001). According to Johnson and Alam (1991) and Uddin and
Lundberg (1999), the paleo-Brahmaputra River flowed towards SW between the
current day juncture of the Shillong Plateau and the Naga Domain (accretionary
prism of the IBR) until the Miocene ended. The uplift of the Shillong Plateau as
well as the westward encroachment and the final abutment of the CTFB accre-
tionary prism (Neogene Outer IBR) against the already-uplifted Shillong Plateau
led to a *300 km westward shift of the Brahmaputra River to its present course
during the Pliocene (Najman et al. 2012; Bracciali et al. 2015; Najman et al. 2016).
Towards the east, the Shillong Plateau comprises the Garo, Khasi and Jaintia hills,
which stretch for *97 km in N-S and *240 km in E-W (Morgan and McIntire
1959). The basement of the plateau is primarily composed of intensely stressed
Precambrian and Early Paleozoic metamorphic and intrusive rocks overlain by the
Cretaceous and Cenozoic sediments in its steep southern face as well as east and
west margins (Wadia 1953; Nag et al. 2001; Biswas et al. 2007; Najman et al.
2016).
The Shillong Plateau presumably has had a major influence on strain partitioning
and consequently tectonics of the Eastern Himalaya that influenced the seismic risk
in the surrounding regions (Bilham and England 2001; Banerjee et al. 2008;
Najman et al. 2016). According to Mitra et al. (2005) and Vernant et al. (2014), the
southern edge of the Shillong Plateau is elevated by stresses arising from the master
thrust (Dauki Fault), and the plateau is tilted northward at approximately 2°–4° as
indicated by a 1–2 km increase in depth of the Moho between the northern edge of
the plateau and the Himalayan Front. Under the Shillong Plateau and the Surma
basin, the Moho has been observed at similar depths of *37 km despite the
contrast of craton vs. thick sedimentary basin. This possibly indicates the uplift of
the overthrusting Shillong Plateau and the concurrent downward flexure of the
Surma basin (Singh et al. 2016). The thickness of the crust of the Indian Plate
beneath the Shillong Plateau is *38–40 km (Singh et al. 2015, 2016). To the north
and northwest, the Bengal Basin extends up to the Himalayan front where it is
separated by the Main Frontal Thrust (MFT) or Himalayan Frontal Thrust
(HFT) (Acharyya 1994; Ganguly 1997). The Assam Basin, situated in the north of
the Shillong Plateau and northwest of the Naga domain, was connected to the
144 M. S. Hossain et al.
Bengal Basin up to the end of the Miocene. During the Pliocene, the upliftment of
the Shillong Plateau separated the Assam Basin from the Bengal Basin except to the
western part, across the N-S oriented Dhubri Fault.
Peninsular India lying south of the Indo-Gangetic Alluvial Plain commonly referred
to as the Indian Shield reportedly occupied a much wider area than the present
triangular shaped region, made up of a diverse mosaic of igneous and metamorphic
terrains that has undergone deformation and metamorphism (Roy 2014). These
terrains, constituting the continental crust have attained tectonic stability since the
Precambrian and are designated as cratons (Valdiya 2010, 2016). The cratons are
flanked by fold belts, with or without a discernible suture or shear zone, suggesting
that the cratons, as crustal blocks or micro-plates, moved against each other (Naqvi
2005). Alternatively, these cratons could be the result of fragmentation of a large
craton that constituted the Indian Shield. Welded fold belts in between the neigh-
bouring cratons indicate either rifting or splitting of cratons. The crustal blocks of
Indian shield are divided into six cratons (Sharma 2010; Valdiya 2010, 2016).
Among these cartons, northeastern Chhotanagpur Granite-Gneiss Complex
(CGGC) and Singhbhum Craton (SC) together with Satpura Mobile Belt or Central
Indian Tectonic Zone (CITZ) and Eastern Ghat Mobile Belt are lying along the
western margin of the Bengal Basin (Misra 2006). The thickness of the crust of the
Indian Plate near the western part of the Bengal Basin is *38 km (Singh et al.
2015). The approximately N-S running Malda-Kishanganj Fault (Nandy 2001;
Vaccari et al. 2011; Mohanty et al. 2013, 2014; Prasad and Pundir 2017) reaches up
to the eastern and northeastern margins of the Rajmahal Hills which is considered
as the northwestern margin of the Bengal Basin.
The Bengal Basin constitutes the eastern continuation of the Indo-Gangetic
Alluvial Plain (IGAP), which separates the Extra Peninsular India to the north and
the Peninsular India to the south (Oldham 1893; Roy 2014). The Indo-Gangetic
plains are stretching across northern India between Punjab to the west and Assam to
the east through Bangladesh. The Extra Peninsular India, which is the mountainous
region formed by the Himalayan ranges and their extensions into the Baluchistan to
the west and Myanmar to the east (Valdiya 2010).
To the north northwest, the Bengal Basin is separated from the Rajmahal Hills by
the approximately N-S running Rajmahal Fault (Ghose et al. 2017). The Rajmahal
Hills of West Bengal and Bihar is a fault bounded small tectonic element situated in
the western edge of the Stable Shelf of the Bengal Basin (Fig. 1). The Early
Cretaceous (100–118 Ma; McDougall and McElhinny 1970; Baksi 1995; Kent et al.
2002) Rajmahal Traps (RT), a sequence of tholeiitic basalts, crop out in the
Rajmahal Hills (McDougall and McElhinny 1970). Geologically, the Rajmahal Hills
is located at the juncture of the Singhbhum Craton (SC) towards west and the
Rangpur Saddle of the Bengal Basin to the east (Bage et al. 2014; Singh et al. 2016).
Synthesis of the Tectonic and Structural Elements … 145
Rajmahal Hills is bounded by the Rajmahal Fault and Saithia-Brahmani Fault in the
east and the west sides, respectively (Ghose et al. 2017). The RT is the result of the
Kerguelen Plume upwelling activity in the Early Cretaceous (Ghatak and Basu 2011,
Roy and Chatterjee 2015), which resulted in a massive amount of mantle material
covering the entire surface (Baksi et al. 1987).
According to Nandy (2001), Ghatak and Basu (2011), Sager et al. (2013) and Roy
and Chatterjee (2015), the N–S running Jamuna Fault is coincident with the Ninety
East Ridge along the path of the Kerguelen Plume in the Bay of Bengal as evidenced
by the hotspot-related magma upwelling in the region. Roy and Chatterjee (2015)
suggest that the Tista Fault, the Dauki Fault and N30°E–S30°W oriented Basin
Margin Fault form a triple-point intersection of lineaments/fractures because of local
doming up of the crust during upwelling of the mantle plume. Mantle upwelling
extruded tholeiite lava through fractures over a wide region in and around the Bengal
Basin. The Rajmahal Hills evolved along the then eastern continental margin of the
Indian Plate, following rifting from the Gondwana. Over time, the upper part of this
tectonic element subjected to extensional brittle deformation and formed graben
structures (Singh et al. 2004). The basalts of the RT above the Rajmahal Hills are
commonly vesicular that are filled with secondary minerals e.g., calcite, analsite,
chalcedony, and agate (Ball 1877). The intertrappean beds are composed of sedi-
mentary rocks like siltstone, claystone, and shale (Valdiya 2010). The central part
carries the impression of up to 28 flows, which has been identified earlier by the
Geological Survey of India (Pascoe 1975; Mukhopadhyay et al. 1986). On the other
hand some other authors refer that the maximum exposed thickness of the lava is
about 230 m or approximately ten flows (Kent et al. 1997). The Sylhet Traps are
equivalent to the Rajmahal Traps, and are reported to be found in the northern edge
of the Surma basin (Roy and Chatterjee 2015). Talwani et al. (2016) suggest that the
Rajmahal and the Sylhet Traps are not separate eruptions as they are connected by a
prominent magnetic doublet (Rahman et al. 1990a). According to them, both the
Rajmahal and the Sylhet Traps are originated from same igneous activity.
The Bengal Fan (Fig. 2), which marks the southern end of the Bengal Basin and
northeastern lobe of the Indian Ocean, floors the entire Bay of Bengal, forming the
largest submarine fan in the world (Curray et al. 2003; Curray 2014). Continental
slope of eastern Sri Lanka and India, southern Bangladesh, and the Sunda Trench
extending from Myanmar to Andaman-Nicobar Islands marks the western, northern
and eastern margins of the fan, respectively (Fig. 2). The southern end of the fan
reaches up to the south distal end of the Bay of Bengal at *7°S. The fan was first
apparently recognized by Dietz (1953), and delineated and named by Curray and
Moore (1971). The length of the fan is *3000 km along N-S, the width is
*1430 km along E-W, the area is *3 106 km2 and the maximum sediment
thickness of the fan is *16.5 km (Curray et al. 2003; Shanmugam 2016).
146 M. S. Hossain et al.
The development of the Bengal Fan is related to the creation of the Bay of Bengal,
which was the direct result of the India–Asia collision and uplift of the Himalaya and
thickening of the Tibetan Plateau. The initial Late Paleocene–Eocene collision of the
Indian Plate with the subduction zone of the north side of the Tethys Ocean initiated a
small fan in the northern Bay of Bengal, now lying beneath the younger deltaic and
shelf sediments of the Bengal Basin (Curray and Moore 1974; Alam et al. 2003;
Curray 2014). From the Eocene onward, collision continued resulting rapid clastic
sedimentation in the Bengal Basin primarily from erosion of the high standing and
rising Himalaya and the Tibetan Plateau, and formation of fan on top of this conti-
nental rise (Curray et al. 2003; Curray 2014). The GBM delta formed by the river
Ganges, Brahmaputra, and Meghna gradually filled the Bengal Basin. The surplus
sediment, which has passed through has been distributed across the entire Bay of
Bengal to eventually form the world’s biggest submarine fan. The fan gradually
prograded southward during the Tertiary (Curray et al. 2003). The Swatch of No
Ground (SoNG) is the main avenue that funneled down sediment from the Bengal
Basin to the Bengal Fan (Curray et al. 2003). It is a shelf-incising aggrading canyon
(Fournier et al. 2016) located 250 km *S of the confluence of the Ganges and the
Brahmaputra rivers and only 30 km south from the coastal delta plain of Bangladesh.
Curray and Moore (1974) assumed that the formation of the modern Bengal Fan
commenced during the Oligocene to the Early Miocene. Two lobes of the fan
separated by the Ninety East Ridge were active until the Middle Pleistocene,
namely the Bengal Fan proper in the west and Nicobar Fan in the east (Curray and
Moore 1974). The SoNG, which was the main supplier of sediments to the Bengal
Fan is largely cut off from its supply of sediment as the modern river sediments
trapped on the Bengal shelf due to the Holocene transgression (Emmel and Curray
1985; Kuehl et al. 1989). The supply of sediments to the Nicobar Fan was cut off by
the convergence of the Ninety East Ridge and the Sunda Trench during the Early
Pleistocene. At present, only the axial area of the Sunda Trench is still receiving
sediments from the fan in the Bay of Bengal (Curray and Moore 1974; Bender
1983). The surface features of the Bengal Fan are mainly active and abandoned
turbidity current channels, and the Ninety East Ridge and Eighty Five East Ridge
(Curray et al. 2003; Curray 2014).
The Bengal Basin, a complex collisional foreland basin, exhibits intense variability
in Neogene sediment thickness that reflects a complicated depositional environment
and tectonics (Curray et al. 2003; Uddin and Lundberg 2004; DeCelles 2012). With
broad attributes of basin geometry, tectonic evolution and stratigraphy, the dis-
cussion begins with the basin geometry, followed by the tectonic evolution. Finally,
the stratigraphy in context to the basin geometry and tectonic evolution has been
discussed in the later part.
Synthesis of the Tectonic and Structural Elements … 147
(Alam 1972; Das Gupta 1977; Guha 1978; Reimann 1993). In general, the
Foredeep Basin shows small-amplitude, isometric or geographically equant gravity
anomalies (Bakhtine 1966). The transitional zone of the continent–ocean crust of
the Indian Plate occurs in this sub-basin between the Eocene shelf-break in the west
and the Barisal–Chandpur Gravity High in the east (Shamsuddin and Abdullah
1997; Uddin and Lundberg 2004; Mukherjee et al. 2009).
The folded flank of the Bengal Basin or Geotectonic Province 3 comprises the
eastern part of the Bengal Basin and is one of the active collisional orogenic belts in
the world. Approximately NNW to SSE trending folds that have developed in the
upper part of the thick deltaic sequence of the CTFB generally exhibit linear or
elongated gravity anomalies with large amplitudes (Figs. 5 and 6) (Bakhtine 1966).
Although standard traditional models suggest that the India–Eurasia (Tibet part)
collision event started 50–55 Ma ago (Rowley 1996; Hodges 2000; DeCelles et al.
2002; Zhu et al. 2005; Najman 2006), Ali and Aitchison (2005, 2008), Aitchison et al.
(2007), Metcalfe (2013), Misra et al. (2015), Baxter et al. (2016) and Mukherjee et al.
(2017) proposed new model with much younger age (Tables 1 and 2). Ali and
Aitchison (2008) discourse the revised plate tectonic reconstructions model into two
phases: (i) Break-up and dispersal of Gondwana—Middle Jurassic through
end-Paleocene, and (ii) Convergence and collision of Indian Plate with the Eurasia
Plate—Eocene. However, before discussing the revised plate tectonic reconstruction
model, it is necessary to consider the palaeogeography of the Neotethys Ocean.
The commonly accepted standard models involved the present-day Indian Plate
with some form of extension to the north as a passive margin, north of which lay
oceanic crust of the Neotethys. As the Indian Plate moved towards the Eurasian
Plate, the Neotethys was consumed beneath the Lhasa Block in a setting similar to
the present-day Sunda Arc (Hall 2002). This palaeogeographic scenario is con-
sidered to be the ‘one ocean–two continent convergence model’. However, fol-
lowing seismic tomography studies of the mantle beneath the Indian Ocean–South
Asia region, geological investigations of the India–Asia suture zone, and paleo-
magnetic and magnetic anomaly based India and Asia motion models, paleogeo-
graphic scenario of Neotethys are considered as the ‘two ocean–two continent
Synthesis of the Tectonic and Structural Elements … 149
Table 1 Major tectonic events related to the Indian and Eurasian plate collision and subsequent
development of the Bengal Basin
Geologic time Age in Ma Major tectonic events
Convergence and collision of the Indian plate with the Eurasian plate—Eocene
Mid- to Late Pliocene *3.5–2.5 The major thrust related uplift of the Shillong Plateau along
the Dauki Fault in the south and the Oldham Fault in the
north. Centre of the Bengal Basin started evolving as a
foreland basin
Late Oligocene *25 Remnant ocean basin took shape due to the collisional
orogeny of the Barail–Cachar Hills at the northeastern corner
of the Indian Plate
Eocene-Oligocene *35 The continental crust of the Indian Plate impacted with the
transition Tibet part of the Eurasian Plate, resulted in subduction of the
northern part of the Indian Plate beneath the southern Tibet,
therefore, initiated the continent–continent collision
Middle Eocene *45 The northern edge of the Indian Plate was still at some
distance south of the Lhasa Block, i.e. southern front of the
Eurasian Plate
Early Eocene *55 The collision of Indian Plate with a Neotethyan intra-oceanic
arc (Dazhuqu Arc/Kohistan–Ladakh Arc), but not with the
Eurasian Plate
Break-up and dispersal of the Gondwana supercontinent—middle Jurassic through Late Paleocene
Late Paleocene *55.9 Indian Plate appears to have reached its maximum level of
isolation after being jettisoned by the Seychelles block
Cretaceous–Paleogene *66 Indian Plate started breaking from the Seychelles
transition
Late cretaceous *90–85 India–Seychelles separated from the Madagascar
Mid-cretaceous *118 The Kerguelen large igneous province began forming on the
floor of the SE Indian Ocean, resulting the Rajmahal Trap
eruptions, and provided a stepping-stone for migrations
between the India–Seychelles–Madagascar and Australia–
Antarctica for much of the Mid-Cretaceous
Mid-early cretaceous *125 East Gondwana moved along the Davie Fracture Zone on a
trail roughly parallel to the coast of the eastern Africa, which
led to formation of true ocean floor (Somalia and
Mozambique Basin)
Early cretaceous *132 Australia–Antarctica began to drift away from the India–
Seychelles–Madagascar with a fan-shaped expanse of ocean
floor that separated the two elements by gradually widening to
the east
Middle Jurassic *170–175 Gondwana supercontinent started breaking following the
rifting of South America–Africa (West Gondwana) from the
India–Seychelles–Madagascar–Australia–Antarctica (East
Gondwana)
Paleozoic—Mesozoic *180 Indian Plate presumably occupied a central location in the
(Early—Middle Jurassic) Gondwana supercontinent
Compiled from Smith and Hallam (1970), Norton and Sclater (1979), Ingersoll et al. (1995), Storey et al.
(1995), Acton (1999), Torsvik et al. (2000), Bilham and England (2001), Biswas and Grasemann (2005a),
Schettino and Scotese (2005), Rabinowitz and Woods (2006), Aitchison et al. (2007), Ali and Aitchison (2008),
Yin et al. (2010), Metcalfe (2013), Misra et al. (2014, 2015), Misra and Mukherjee (2015), Baxter et al. (2016),
Najman et al. (2016, 2017), Mukherjee et al. (2017)
150 M. S. Hossain et al.
Table 2 Major tectonic events related to the Indian and Burmese plate collision and subsequent
development of the Folded Flank (CTFB) of Bengal Basin
Geologic time Age in Major events
Ma
Late *0.075 The fold has been progressively growing westward, and some
Pleistocene - of the folds in the westernmost part have been actively
Recent developing
Early *2 Sediments in the upper part (*up to 5 km) within the major
Pleistocene individual thrust started to deform above a de’collement level
by combined thin-skinned and thick-skinned tectonic
processes. Many of the folds in the westernmost part have been
actively developing
Late Pliocene *3.5–3 Oblique subduction of the Indian Plate beneath the Burmese
Plate in an arc-trench setting developed accretionary prism and
the CTFB started to develop within the upper parts of the thick
deltaic sequence in the eastern margin of the Bengal Basin
Early Miocene *20 Sediment contributions to the Bengal Basin began to arrive
from the Indo-Burman Ranges (IBR) due to subduction of the
Indian Plate beneath the Burmese Plate
Early Eocene *45 The northeast corner of the Indian Plate making a glancing
contact with the Sumatra Block, followed by the Burmese Plate
Compiled from Curray et al. (2003), Uddin and Lundberg (2004), Acharyya (2007), Ali and
Aitchison (2008), Maurin and Rangin (2009a), Wang et al. (2014) and Najman et al. (2016)
convergence model’ (Fig. 3) (Allégre et al. 1984; Abrajevitch et al. 2005; Aitchison
et al. 2007; Ali and Aitchison 2008; Baxter et al. 2016). According to this model,
during the Late Cretaceous–Cenozoic passage of the Indian Plate towards the Tibet
part of the Eurasian Plate, the Neotethys comprises two oceanic plates separated by
north-dipping subduction zone (Ali and Aitchison 2008). Considering the
present-day island arcs as a guide, Ali and Aitchison (2008) and Baxter et al. (2016)
suggest that there might have some volcanic islands (Neotethyan arc) spaced every
30–90 km apart present in the Neotethys ocean. If the ‘two ocean–two continent
convergence model’ is correct, one might, therefore, predict that remnants of the
two separate ocean slabs are present beneath the subduction system in the India–
Eurasia collision belt in southern Tibet (Yarlung Tsangpo Suture Zone- Indus
River) and northern India (Himalayan Foredeep) (Ali and Aitchison 2008).
resulting the Rajmahal Trap eruptions. This could have provided a stepping-stone for
migrations between India–Seychelles–Madagascar and Australia–Antarctica for
much of the mid-Cretaceous. India–Seychelles–Madagascar was connected until the
beginning of the Late Cretaceous (recent reviews in Misra et al. 2014, 2015; Misra and
Mukherjee 2015; Mukherjee et al. 2017). In the Late Cretaceous (90–85 Ma), India–
Seychelles separated from Madagascar (Storey et al. 1995; Torsvik et al. 2000;
Aitchison et al. 2007). India started breaking from Seychelles around the Cretaceous–
Paleogene boundary (*66 Ma), and India appears to have reached its maximum level
of isolation after being jettisoned by Seychelles block at the end of the Paleocene
(55.9 Ma). It is believed that the jettisoned Seychelles block coincided with the main
Deccan Trap eruptions. The centre of the Indian Plate, which now sits on *23.5°N
(Tropic of Cancer), was at *30°S in the Late Cretaceous, some 6000 km to the south
from where it is today (Ali and Aitchison 2008).
The Indian Plate migrated rapidly northwards during the Late Cretaceous and the
Paleocene (Acton 1999). According to Acton’s (1999) model, northwards advance
of the Indian Plate in between 120 and 73 Ma was *6.6 cm/year, which increased
to *21.1 cm/year between 73 and 57 Ma. At *57 Ma, the plate motion abruptly
decreased to 9.5 cm/year, which continued until 20–30 Ma when there was a fur-
ther major slowdown of the rate of movement. According to Ali and Aitchison
(2008) and Baxter et al. (2016), the ‘57 Ma event’ mark the collision of the Indian
Plate with a Neotethyan intra-oceanic arc, but not with the Eurasian Plate. Ali and
Aitchison (2004, 2006) and Aitchison et al. (2007) calculate the position of the new
pole that sits 5°–9° further from the present-day North Pole. This new pole places
the southern Lhasa Block at *28°N at 55 Ma. Since then, this part of the Eurasian
Plate has rotated *21° clockwise relative to the spin axis and migrated 1100 km
east. Ali and Aitchison (2008) suggest that evidence related to the collision of the
oceanic part of the Indian Plate with the Dazhuqu Arc in the earliest Eocene
(*55 Ma) has been found along the 2500-km-long Indus–Yarlung Tsangpo suture
zone (trending *E-W in the eastern part and *NW-SE in the western part;
Aitchison et al. 2007). The Dazhuqu Arc, also known as the Kohistan–Ladakh Arc
(Metcalfe 2013) may have been formed as early as *135 Ma, i.e. in the Early
Cretaceous (Bosch et al. 2011). A number of island arc fragments that had obducted
onto India after it collided with the Dazhuqu Arc at *55 Ma are present in this
suture zone. At the Middle Eocene (*45 Ma) the northern edge of the Indian Plate
was still at some distance south of the Lhasa Block, i.e. southern front of the
Eurasian Plate. Isotopic studies of igneous rocks and the biostratigraphic investi-
gations of sedimentary rocks from the suture zone, led to the conclusion that the
hard collision between continental part of the Indian Plate and the Eurasian Plate
(southern Tibet, i.e. Lhasa Block), i.e. continent-continent collision happened at the
very end of the Eocene, *35 Ma (Ali and Aitchison 2008; Baxter et al. 2016).
Therefore, Aitchison et al. (2007), Ali and Aitchison (2008) and Baxter et al. (2016)
Synthesis of the Tectonic and Structural Elements … 153
reasonably argued that oceanic crust of the Indian Plate collided into a
sub-equatorially located intra-oceanic arc (Neotethyan arc) at *55 Ma, and later
the continental crust of the Indian Plate impacted with the Tibet part of the Eurasian
Plate starting around 35 Ma (Table 1). Approximately at the Eocene-Oligocene
boundary (*35 Ma), northern Indian Plate entered the north-dipping subduction
zone beneath southern Tibet, thus marking the initiation of continent–continent
collision (Baxter et al. 2016). To the east, the northeast corner of the Indian Plate
making a glancing contact with Sumatra Block, followed by the Burmese Plate
from *45 Ma, i.e. Early Eocene onwards (Ali and Aitchison 2008).
The basin evolved through two major tectonic episodes. First, it initiated as an
intra-cratonic rift basin within Gondwana landmass during Late Paleozoic–Mid
Mesozoic and received the continental Gondwana sediments. This episode of basin
development ended with widespread volcanism as continental flood basalts known
as the Rajmahal Trap covered the Gondwana sediments. The second episode of
basin development began in the Late Mesozoic with the break-up of Gondwana and
is still going on (Alam 1989). At this stage, the tectonic evolution of the greater
Bengal Basin is fundamentally related to the collision pattern of the Indian Plate
with the Eurasian Plate to the north and the Burmese Plate to the east. The first
uplift in the Himalayan and Indo-Burman region commenced in the Oligocene and
the Early Miocene, respectively (Bender 1983). In its northern journey, the Indian
Plate collided with the Eurasian Plate and caused folding and thrusting forming the
Himalayan orogenic belt. Further movement of the Indian Plate was in the
north-easterly direction, resulted in collision of northeastern end of the Indian Plate
with the Burmese Plate, the former subducting below the latter, and the sedimentary
accretionary prism gave rise to the formation of the CTFB and the IBR (Alam et al.
2003; Steckler et al. 2008; Wang et al. 2014).
The peri-cratonic part on the eastern margin subsided continuously and received
voluminous sediments from the Late Mesozoic through Tertiary to the Recent
times. This foreland basin containing a succession of dominantly deltaic sediments
derived primarily from the erosion of the Himalaya and the IBR. Sediments
accumulate in the basin through the Ganges, Brahmaputra and Meghna
(GBM) River systems, and are dispersed into the Bay of Bengal, forming the largest
submarine fan in the world (Curray et al. 2003). According to Curiale et al. (2002)
and Curray et al. (2003), sediment contributions to the basin began to arrive pri-
marily from the Himalaya and the Indo-Burman Ranges around the Early Oligocene
(*35 Ma), and the Early Miocene (*20), respectively, and have been prograding
southward presently. To keep the isostatic equilibrium (Mukherjee 2017), the
arriving mass of these huge sediments loaded and depressed the underlying litho-
sphere further, producing additional accommodation space for deltaic sediments.
The additional lithospheric depression and accommodation space have resulted
from the southward thrusting of the Shillong Plateau over the basin through the past
154 M. S. Hossain et al.
5 Ma in the north, and from the westward thrusting of the Indo-Burman Ranges
toward the basin in the east (Johnson and Alam 1991).
Geodynamic development of the previously mentioned three geotectonic pro-
vinces and their successive tectonic evolution are now discussed under the light of
the India and Asia collision to the north, and the India and Burma collision to the
east. Among the three geotectonic provinces (Alam et al. 2003), the development of
the Geotectonic Province 1—Stable Shelf and Geotectonic Province 2—central
Foredeep Basin are related to the collision between the Indian Plate to the south and
the Eurasian Plate to the north, whereas development of Geotectonic Province 3—
CTFB is related to the collision of the Indian Plate to the west and the Burmese
Plate to the east.
The Early Cretaceous rifting of the Indian Plate from Gondwana and con-
comitant volcanic eruptions along the northeastern margin of the Indian Plate ini-
tiated the tectonic evolution of the Geotectonic Province 1 and 2 (Alam et al. 2003).
In Geotectonic Province 1, pre-rift Permo-Carboniferous sediments are encountered
within the subsurface graben basins on top of the Precambrian Basement rock
(Khan 1991a, b; Reimann 1993). SW–NE trending lineaments are observed in both
the aeromagnetic anomaly map (Rahman et al. 1990a) and in the Bouguer gravity
anomaly map (Rahman et al. 1990b; Khan and Rahman 1992). This reflects grabens
on the continental crust part of the Geoteconic Province 2 (Surma basin)
beneath the thick Late Cenozoic sediment cover, which may also hold the Permo-
Carboniferous sediments at depth. The basement rock beneath the Geotectonic
Province 1 is characterized by tensional tectonics. Tension created a complex
graben system with tilted and/or downthrown blocks, which were intersected by
many normal faults (Khan and Agarwal 1993).
Along the Eocene Hinge Zone and its immediate east, alternating high and low
magnetic values could represent (*750–950 nT; Rahman et al. 1990a) a probable
transition from continental to oceanic crust (Curray 2014). The Barisal–Chandpur
Gravity High probably indicates the rift valley formed during the initial break-up of
Gondwana and formation of the Indian Plate along which transitional zone between
continent-ocean crust of the Indian Plate has been marked by pulses of basalt flow
(Desikachar 1974; Lohmann 1995; Alam et al. 2003; Curray 2014). Further to the
east, seismic signature on top of the basement is characterized by numerous
cross-cutting and down-going reflectors associated with basalt flows. They indicate
oceanic crust below. Alam et al. (2003) also suggest that the continent–ocean crust
boundary bends eastward beneath the Surma basin, and probably continues towards
the juncture of the Dauki-Haflong Thrust. Hence, north of this Hinge Zone, at least
northern part of the Surma basin is floored by the continental crust, whereas
southeast of this line, basement rocks of the Bengal Basin are oceanic rather than of
continental origin (BOGMC 1997). In addition, it is believed that position of the
Eocene Hinge Zone is within Geotectonic Province 1, and it truncates against the
Dauki Fault in the northeast boundary of the deeper foredeep centre.
Up to the Oligocene, the Geotectonic Province 1 and 2 are believed to be
undergone similar tectonic and sedimentary evolution (Alam et al. 2003). From the
Late Oligocene onward, when the remnant ocean basin (Ingersoll et al. 1995) took
Synthesis of the Tectonic and Structural Elements … 155
shape due to the collisional orogeny of the Barail–Cachar Hills at the northeastern
corner of India (Nandy 1986), Geotectonic Province 2 has undergone its own
tectonic evolution. A major change in sedimentation pattern here probably occurred
in the Mid Pliocene by the major thrust related uplift (upthrust) of the Shillong
Plateau (Najman et al. 2016) along the Dauki Fault (Bilham and England 2001;
Biswas and Grasemann 2005a) in the south and Oldham Fault (Yin et al. 2010) in
the north. In the Shillong Plateau, the Archaean Basement rocks are found *2 km
high, whereas equivalent rocks occur *16–18 and 4–5 km below sea level, in the
south and north of the plateau, respectively (Alam et al. 2003; Najman et al. 2016).
The Geotectonic Province 3—CTFB started to develop during the Late Pliocene
(Maurin and Rangin 2009a) within the upper parts of the thick deltaic sequence in
the eastern margin of the Bengal Basin. At the same time, the Geotectonic Province
2 started evolving as a foreland basin at the centre; whereas in the Geotectonic
Province 1, the authors of this chapter believes foredeep sediments will thrust
northwestward in near future. Oblique subduction of the Indian Plate beneath the
Burmese Plate in an arc-trench setting developed accretionary prism as well as
major east-dipping thrusts (Kaladan Fault to the east, and Chittagong Coastal Fault
to the west), which largely controlled the structural evolution of the Geotectonic
Province 3 (Gani and Alam 1999; Acharyya 2007; Maurin and Rangin 2009a;
Wang et al. 2014). In this province, the sediments in the upper part (*up to 5 km)
within the major individual thrust deformed above a de’collement level by com-
bined thin-skinned and thick-skinned tectonic processes during the last 2 Ma
(Uddin and Lundberg 2004; Maurin and Rangin 2009a), giving rise to a series of
elongate, N–S trending curvilinear folds (Sikder and Alam 2003). Through a bal-
anced cross section across this geotectonic province, Maurin and Rangin (2009a)
estimated a total E-W shortening of about 11 km in the past 2 Ma, which suggests a
shortening rate of about 0.5 cm/year. In this regard, recent investigations suggest
that the fold belt has grown progressively westward and many of the folds in the
westernmost part of this province have been active only from the Late Pliocene or
even later (Table 2) (Johnson and Alam 1991; Khan et al. 2005; Steckler et al.
2008; Maurin and Rangin 2009a; Wang et al. 2014; Khan et al. 2015, 2018).
4 Stratigraphy
The foremost depositional phase in the Bengal Basin began following the separa-
tion of the Indian Plate from Antarctica at about the beginning of the Late
Cretaceous. This is in exception of minor Carboniferous coal that preserved at least
locally on Precambrian continental crust (Sclater and Fisher 1974; Molnar and
Tapponnier 1975; Uddin and Lundberg 2004). In the beginning of the Late Eocene,
thick Tertiary clastic sediments accumulated in the basin with deposition acceler-
ating with the arrival of clearly orogenic sediments in the earliest Miocene (Bender
1983; Uddin and Lundberg 1998, 2004). According to Alam et al. (2003), the
156 M. S. Hossain et al.
present day active sediment depocentre is the Hatia Trough in the Geotectonic
Province 2 as well as its extension to the south into the Bengal Fan.
The sediments preserved in the Stable Shelf part of the Bengal Basin, i.e.
Geotectonic Province 1 (Tables 3 and 4) show a progressive thickening towards
north and south from the shallowest part of the basement at the Rangpur Saddle
from which it continues to slope progressively in both directions. Sediments attain a
thickness of 3–4.5 km, and *3.5 km to the northern and southern edge of the
Stable Shelf, respectively (Uddin and Lundberg 2004; Guha et al. 2010). The
stratigraphic unit on top of the Stable Shelf as well as in both southern and northern
slopes varies in thickness, spatial extent and lithology over geologic time (Reimann
1993). The basement grabens are filled up by the Gondwana sediments comprising
sandstone, shale and coal. The Rajmahal Trap lies above the Gondwana sediments
and are overlain by the Tura Sandstone, followed by the fossliferous Eocene Sylhet
Limestone and then the post Eocene sediments comprising sands, gravels and clay.
The facies evolution ranges from continental and volcanics through brackish and
lagoonal to shallow open marine to estuarine and fresh water environments
(Reimann 1993; Alam et al. 2003).
The sediments preserved within the Foredeep part, i.e., Geotectonic Province 2
(Table 5) and 3 (Table 6) are ca.16–22 km thick sequence of Cenozoic sediments
(Curray and Moore 1971; Curray 1994; Gani and Alam 1999; Uddin and Lundberg
2004), where facies evolution ranges from marine through deltaic to fluvial envi-
ronments. From oldest to youngest, the stratigraphic units are the Tura Sandstone,
Sylhet Limestone, Kopili Shale, Barail, Bhuban, Bokabil, Tipam Sandstone,
Girujan Clay, Dupi Tila Sandstone, and Dihing Formations. Although Rahman
et al. (2017) have tried to correlate sediment provenance with different Himalayan
tectonic units, the relative contributions of the sediments to the basin from the
Himalaya, CTFB-IBR, and Indian Shield and Shillong Plateau are much debated
(Johnson and Alam 1991; Gani and Alam 1999; Uddin and Lundberg 2004;
Najman et al. 2008, 2016).
The India-Eurasia and India-Burma (Myanmar) collisions have affected the basin since
the Tertiary (Bender 1983). Conventionally, the basin can be divided into two large
divisions based on overall geotectonic settings: Stable Continental Shelf to the west
and Bengal Foredeep to the east separated by the Continental Slope also known as the
Eocene Hinge Zone (Bakhtine 1966; Guha 1978; Bender 1983; Reimann 1993;
Uddin and Lundberg 2004). The Bengal Foredeep occupies the enormous area between
the Eocene Hinge Zone to the west and the Indo-Burman Ranges to the east and plays
important role in the tectonic evolution of the Bengal Basin. In the foredeep part of the
Bengal Basin, a thick pile of sedimentary strata overlies a deeply subsided basement.
Table 3 Stratigraphic succession of the Geotectonic Province-1A: Southern Slope of the Rangpur Saddle, Stable Shelf of the Bengal Basin in Bangladesh
Modified after Zaher and Rahman (1980), Khan (1991a, b), Lindsay et al. (1991), Reimann (1993), Alam (1997), BOGMC (1997) and Alam et al. (2003)
157
158 M. S. Hossain et al.
Table 4 Stratigraphic succession of the Geotectonic Province-1B: Southern Slope of the Rangpur
Saddle, Stable Shelf of the Bengal Basin in West Bengal, India
Environment
Age Lithological Description
Shelf Facies Basin Facies
Bengal Alluvium Bengal Alluvium
Recent - Pleistocene
Debagram Formation Ranaghat Formation
Silty sandstone with clay
Pliocene Brakish to marshy Brakish to marshy
Lagoonal to littoral Lagoonal to littoral
Late
Pandua Formation Matla Formation
Siltstone with sandstone and
claystone
Brakish to marshy Brakish to marshy
Miocene Middle Lagoonal to littoral Lagoonal to littoral
Shallow marine Shallow marine
Memari/Burdwan
Formation
Siltstone sandstone and
Early carbonaceous shale
Brakish to marshy
Lagoonal to littoral
Late
Kopili Shale
Brakish to marshy
Calcareous shale
Late Lagoonal to littoral
Eocene Shallow marine
Sylhet Limestone
Middle
Foraminiferal and algal limestone Brakish to marshy
Early with sandstone Lagoonal to littoral
Shallow marine
Late Jalangi Formation
Paleocene
Continental
Early Coarse to medium grained
sandstone with lignite and shale
Brakish to marshy
Lagoonal to littoral
Late
Bholpur/Ghatal
Formation
Cretaceous
Kaolinitic sandstone with shale
Estuarine
Early
Continetal
Basalt Rajmahal Trap
Sandstone, shale, carbonaceous Pre-Trappean/Gondwana
Permo-Carboniferous
shale, coal and sandstone
Modified after Biswas (1963), Sengupta (1966) and National Data Repository (2015)
Based on the gravity studies, the foredeep part is divided into a northeast-southwest
deeper Foredeep Basin or Western Platform Flank of the foredeep, just east of the
Continental Slope, and an Eastern Folded Flank of the foredeep that comprises the
CTFB (Guha 1978; Khandoker 1989; Khan 1991a, b; Reimann 1993; Uddin and
Table 5 Stratigraphic succession of the Surma basin (northern part of the Geotectonic Province-2), central deeper Foredeep Basin, Bengal Basin
Thickness Depositional
Age (approx) Group Formation Lithology Tectonic Events
Max (m) Environment
Holocene Alluvium Fluvial
Poorly consolidated sandstone
Pleistocene Dihing
Dihing and clayey sandstone.
Upper 3350 Fluvial
Medium to coarse ferruginous
Dupi Tila sandstone with layers of quartz
Dupi Tila Folding in the eastern
Late Pliocene pebbles and siltstone with
Lower lignitic fragments and petrified Bengal Basin
Dupi Tila wood.
Modified after Evans (1964), Holtrop and Keizer (1970), Khan et al. (1988), Hiller and Elahi (1988), Khan (1991a, b), Reimann (1993), Shamsuddin et al.
(2001) and Alam et al. (2003)
159
160
Table 6 Stratigraphic succession of the Chittagong-Tripura Fold Belt (Geotectonic Province-3), Folded Flank, Bengal Basin
Holocene Fluvial
Alluvium
Surma
Sandstone and pebbly sandstone at Shallow marine
Miocene
Middle the top and sandy shale at the
Bhuban bottom 1500
Miocene
Modified after Evans (1964), Khan (1991a, b), Reimann (1993), Gani and Alam (1999, 2003), Alam et al. (2003)
M. S. Hossain et al.
Synthesis of the Tectonic and Structural Elements … 161
The Stable Shelf (Fig. 4) to the northwest, also known as the Stable Pre-Cambrian
Platform or Western Flank of the Bengal Basin, comprises two major tectonic
elements: (i) the Foreland Shelf to the west, and (ii) the Eocene Hinge Zone or shelf
break to the east (Bakhtine 1966; Guha 1978; Matin et al. 1983). The Bouguer
anomaly map of the Stable Foreland Shelf is characterized by large negative
anomalies (0 to −140 mGal), mostly related to the basement structural features
(Fig. 5) (Rahman et al. 1990b; Khan and Rahman 1992). Khan and Rahman (1992)
identified four tectonic zones on the basis of the trend, shape and magnitude of the
Bouguer gravity anomaly contours and the basement faults from the aeromagnetic
data. Except the Eocene Hinge Zone, alternating gravity highs and lows with closed
contours indicate the presence of numerous graben and half-graben structures in
most of the Foreland Shelf, where Gondwana coal/sediments have been deposited
(Khan 1978; Khan and Rahman 1992; Khan and Chouhan 1996).
Fig. 4 Simplified tectonic map of the Geotectonic Province 1 of the Bengal Basin and its
surroundings. The tectonic element boundaries are based on Bakhtine (1966), Guha (1978) and
Matin et al. (1983). Faults and lineaments are mostly taken from Kayal (2008). MBT: Main
Boundary Thrust; MFT: Main Frontal Thrust
Synthesis of the Tectonic and Structural Elements … 163
Fig. 5 Bouguer gravity anomaly map of Bangladesh, which covered major portion of the Bengal
Basin. Projection used in this map is the Lambert Conformal Conic with Everest 1969 spheroid
(modified after Rahman et al. 1990b)
164 M. S. Hossain et al.
the Bengal Basin (Reitz et al. 2015). Starting from the northernmost tectonic ele-
ment, the sequential description of these tectonic elements of the foreland shelf is
given with their tectonic evolution history. Note as the N-S oriented Rajmahal Hills
is situated outside the Basin Margin Fault Zone (Raman et al. 1986; Roy and
Chatterjee 2015) and at the western edge of the Foreland Shelf (Raman et al. 1986),
it is not a part of the Bengal Basin.
Himalayan Foredeep
This tectonic element forms the northwestern most part of the Bengal Basin, lies
south of the Main Frontal Thrust (MFT) and is covered by the Recent to sub-Recent
piedmont plain deposits. This element is marked by the high negative Bouguer
gravity anomaly from −110 to −150 mGal within a short distance in the extreme
northwest of Bangladesh, and suggests noticeable thickening of basinal strata
northward into the Siwalik foreland basin of the northeastern Himalaya (Khan and
Rahman 1992; Rabbani et al. 2000; Uddin and Lundberg 2004). The −110 mGal
contour near Panchagarh represents the approximate southern boundary of the
Himalayan Foredeep with the Dinajpur Slope. The width of this tectonic element at
the centre is <50 km, but gradually increases towards both east and west.
The Neogene sediments consisting of sandstones, subordinate shales and clays,
and gravel beds are well developed in this tectonic element and attain a thickness of
3–4.5 km (Rabbani et al. 2000; Guha et al. 2010). The only well drilled in this
tectonic element is located on the north-western most tip of Bangladesh at
Salbanhat in Tetulia by Shell Oil Co. in 1988. The well touched the basement at
2518 m depth penetrating the Mio-Pliocene sequence but did not encounter the
Eocene Limestone (Rabbani et al. 2000). The northern edge of this tectonic element
along the MFT shows clear evidence of an active deformation zone as indicated by
active faulting, uplift and surface-rupture, and earthquakes (Thakur 2004). The
main deformation front related to the India-Eurasia collision slowly migrated
southward in this zone. Two major lineaments transect the Himalayan Foredeep of
the Bengal Basin along NW-SE direction, namely Tista and Gangtok lineaments
(Chopra et al. 2013; Baruah et al. 2016).
The Dinajpur Slope is characterized by almost E-W trending linear Bouguer gravity
contours, with values ranging between −50 and −110 mGal and sloping NNW
between Panchagarh to the north and Nilphamari to the south (Khan and Rahman
1992; Rabbani et al. 2000). The Neogene sediments are deposited in this tectoni-
cally down-warped part of the Foreland Shelf, which is known as the northern slope
of the Rangpur Saddle, or Dinajpur Slope or Northern Foreland Shelf (Reimann
1993) of the Bengal Basin. Although Reimann (1993) included Dinajpur Slope into
the Himalayan Foredeep, Guha et al. (2010) placed Dinajpur Slope as a separate
Synthesis of the Tectonic and Structural Elements … 165
tectonic element in between Himalayan Foredeep to the north and Rangpur Saddle
to the south. The width of this zone ranges between 100 and 70 km from WSW to
ENE, and gently plunges northward to the Himalyan Foredeep approximately at 1°–
3°. The Dinajpur Slope turns *E-W north of the Garo Hills of the Shillong Plateau
and then connects to the Assam Basin. The Dhubri Fault with approximately N-S
orientation separates the Dinajpur Slope to the west and the Assam Basin to the
east. One major fault and two major lineaments transect this tectonic element in
which the Katihar-Nilphamari Fault along *E-W, and the Tista and the Gangtok
lineaments along NW-SE (Chopra et al. 2013; Baruah et al. 2016). Moreover, N-S
graben structures with Gondwana fill are also present (Khan and Rahman 1992;
Reimann 1993).
Rangpur Saddle
The northern edge of the Bogra slope is characterized by the elongate close-spaced
gravity contours of 0 to −30 mGa1, extending from Nawabganj to Jamalganj up to
Gaibandha. From south of the Nawabganj-Gaibandha up to Singra area, several
prominent closed anomaly contours are observed reflecting gravity highs and lows
within −20 to −50 mGal range representing intra-cratonic basins, i.e. graben
structures where Gondwana sediments deposited (Khan and Rahman 1992). This
tectonic element is also known as Southern slope of the Rangpur Saddle or the
Western Foreland Shelf of the Bengal Basin, which contains various Permian to
Recent rocks, laid down on the Precambrian Basement rocks (Khan and Rahman
1992; Reimann 1993). The width of this zone ranges 60–125 km from NE to SW,
and gently plunges southeast to the shelf edge at 1°–3° is clearly marked by the
Eocene Sylhet Limestone on the seismic section (Salt et al. 1986; Reimann 1993).
Seismic contours on top of the Eocene Limestone reveal a number of NE-SW
Synthesis of the Tectonic and Structural Elements … 167
trending faults of which the Debagram-Bogra Fault is the most prominent (Fig. 4).
Another two major faults are also present in this tectonic element and these are the
NW-SE oriented Ganges-Padma Fault to the west and the N-S oriented Jamuna
Fault to the east. Recent GPS derived geodetic data suggest an overall very low
subsidence rate of less than 1 mm/year in the area (Reitz et al. 2015). The Bogra
Slope (Fig. 4) turns approximately E-W near the Garo Hills of the Shillong Plateau
and then traverses along the southern slope of the plateau towards the Khasi and
Jaintia Hills and reaches up to the southern fringe of the Mikir Hills. This tectonic
element started developing at the time of rifting of the Gondwana during the
Carboniferous-Permian when graben and half-graben structures were formed, fol-
lowed by the drifting of the East Gondwana from the remainder during the
Jurassic-Cretaceous. The later event results the Rajmahal lava flows and formed the
Rajmahal Trap, which crops out in the Rajmahal Hills. From the Late Cretaceous to
Eocene, carbonate sediments deposited at the lower edge of the shelf, whereas
arenaceous sediment were deposited at the upper edge (Reimann 1993).
This zone is also known as shelf break or paleo-continental slope or trace of the
Eocene shelf edge (Bakhtine 1966; Sengupta 1966; Bender 1983; Matin et al. 1983;
Rahman et al. 1990a; Reimann 1993; Uddin and Lundberg 2004; Singh et al. 2016)
and lies between the Stable Shelf/Foreland Shelf to the west and the Foredeep Basin
to the east. This tectonic element is characterized by almost ENE-WSW trending
linear Bouguer gravity contours, with values ranging approximately between −30 to
−15 mGal and sloping towards southeast (Khan and Rahman 1992). Precambrian
basement rock dips southeast abruptly from 2°–3° to 6°–12° at the contact of the
stable platform and the Eocene Hinge Zone to the west, and then dips more gently
1°–2° again in the southeast at the contact of the Hinge Zone and deeper foredeep
basin to the east (Uddin and Lundberg 2004). At its upper northwestern edge, the
recorded seismic depth on top of the Eocene Sylhet Limestone is 3500 m, whereas
at its lower southeastern edge, the recorded seismic depth on top of the same
limestone is 5000 m. Although Sengupta (1966) and Khandoker (1989) named this
NNE-SSW running narrow 25–100 km Hinge zone as the ‘Calcutta–Mymensingh
gravity high’, more recent data (Khan and Agarwal 1993) suggest that this term is
to some extent confusing and named more appropriately as Shelf-Break (Uddin and
Lundberg 2004; Roy and Chatterjee 2015). Although it is usually shown to truncate
against the Dauki Fault in the northeast, the other school of thoughts suggest that
the Hinge Zone progressively convexes basin-ward to the northeast and then passes
somewhere through the northern end of the Surma basin and possibly continues
towards the Haflong Thrust at the northeastern corner of the Bengal Basin (Alam
et al. 2003).
The Eocene Hinge Zone marks the structural as well as depositional transition
between central foredeep basin to the southeast and stable shelf to the northwest.
This tectonic element possibly marks the transition from the thick continental crust
168 M. S. Hossain et al.
(west) to extended thinned crust of the continental margin (east) (Singh et al. 2016).
Eocene Sylhet Limestone, the most prominent seismic reflector of the Bengal Basin
has been interpreted to imply this shelf break (Reimann 1993). However, results of
the recent seismic investigations suggest NE alignment of this Hinge Zone up to
90° 40′E and 25° 15′N, where the E-W oriented Dauki Fault presumably truncates
this structural element. More towards east, this paleo-continental slope is not rec-
ognizable even in the deeper seismic section (6 s TWT). This probably indicates
that the southern portion of the shelf as well as slope are downthrown by the Dauki
Fault and now hidden under the thick pile of the Neogene sediments deposited in
the northern part of the Surma basin (Salt et al. 1986). Although significant increase
of sedimentary thickness from *3 to *17 km is noticed across the Hinge Zone
(west to east), the crust thins from 38 km at the Indian Craton to 34 km at the Hinge
Zone to 16–19 km at the central deeper foredeep basin (Singh et al. 2016). Such
thickness of crust at the central deeper foredeep basin may result from igneous
activity related to the Kerguelen Plume at the time of rifting (Ray et al. 2005; Singh
et al. 2016).
The central deeper Foredeep Basin or the Western Platform Flank occupies the
huge area between the Eocene Hinge Zone to the west, the Shillong Plateau to the
north, and the CTFB to the east. This deeper foredeep basin is approximately
200 km wide to the north, narrowed at the middle, and then gradually widens to
about 500 km to the south, shows an overall NE trend (Fig. 6) (Reimann 1993).
The basin shows small-amplitude, isometric or geographically equant Bouguer
gravity anomalies (Bakhtine 1966; Uddin and Lundberg 2004). The deeper fore-
deep basin is a mosaic of few sub-basins (lows) and buried highs/ridges. This
deeper foredeep basin narrowed north-eastward by the Madhupur High in west and
the Tripura Uplift in the east, and broadly divides the basin into two parts. The oval
shaped Surma basin occupies the northern part of the deeper foredeep basin
whereas southern part is known as southern sub-basin. A NE-SW trending gravity
and magnetic anomaly known as the Barisal-Chandpur High separates the southern
part of the deeper foredeep basin into another two sub-basins, the Faridpur Trough
in the west and the Hatia Trough in the east (Alam 1972; Das Gupta 1977; Guha
1978; Reimann 1993). The deeper foredeep basin, therefore, can be divided into six
tectonic elements, and from the north to the south these are the Surma basin/Sylhet
Trough, the Madhupur-Tripura High/Tangail-Tripura High, the Faridpur Trough,
the Barisal-Chandpur High, the Hatia Trough, and the Bay of Bengal (Fig. 6)
(Bakhtine 1966; Guha 1978; Mirkhamidov and Mannan 1981; Matin et al. 1983;
Uddin and Lundberg 1999; Mukherjee et al. 2009).
Synthesis of the Tectonic and Structural Elements … 169
Fig. 6 Simplified tectonic map of the Geotectonic Province 2 of the Bengal Basin and its
surroundings. The tectonic elements boundaries are based on Bakhtine (1966), Guha (1978) and
Matin et al. (1983). Faults and lineaments are mostly taken from Biswas and Grasemann (2005a),
Kayal (2008), Khan et al. (2011), and Wang et al. (2014). A: Anticline; Atg: Atgram; Bakh:
Bakhrabad; Bang: Bangora; Beani: Beanibazar; Begum: Begumganj; Bi: Bibiyana; CCF:
Chittagong Coastal Fault; Ch: Chhatak; Com: Comilla; Daud: Daudkandi; F: Fault; Fench:
Fenchuganj; Go: Gowainghat; Ha: Habiganj; Haka: Hakaluki; Ja: Jalalabad; Ka: Kamta; Kac:
Kachua; Kai: Kailas Tila; Ku: Kutubdia; Kus: Kushiara; Lalm: Lalmai; Magn: Magnama; Me:
Meghna; Moul: Moulvibazar; Na: Narsingdi; Path: Patharia; Ra: Rashidpur; Rajn: Rajnagar; S:
Syncline; San: Sangu; Sand: Sandwip; Sh: Shahbazpur; Sri: Srikail; Srim: Srimangal; Sun:
Sunamganj; Sundal: Sundalpur; Sy: Sylhet and Ti: Titas
Fig. 7 Schematic cross-section across the Bengal Basin and its surroundings (drawn approximately N-S along the line AB as shown in Fig. 1)
(modified after Murphy and Staff of BOGMC 1988; BOGMC 1997 and Webb et al. 2013)
M. S. Hossain et al.
Synthesis of the Tectonic and Structural Elements … 171
Staff BOGMC 1988; Curray 1994; Uddin and Lundberg 2004), thickness <18 km
is also reported (Paul and Lian 1975; Banerji 1979; Guha 1978; Matin et al. 1986;
Singh et al. 2016). Beneath this thick sedimentary pile, type and thickness of the
crust of the Indian Plate is still debated (Singh et al. 2016). However, it is more or
less accepted that beneath the northwestern half of the Surma basin, the crust is
continental to transitional type; whereas beneath southeastern half of the basin, the
crust is oceanic. The Bouguer gravity map of Bangladesh suggests a minimum
Bouguer anomaly (−80 mgl) for the Surma basin (Ali and Raghava 1985; Uddin
and Lundberg 2004). Along N-S section, more or less flat and tectonically less
disturbed nature of the Moho has been observed at similar depth of *37 km
beneath this tectonic element as well as the Shillong Plateau (Singh et al. 2016).
Except for the western border made by the Eocene Hinge Zone, all other structural
features framing the Surma basin in north, east and south are the results of mostly
the Pliocene-Recent compressional tectonics (Reimann 1993). It is assumed that the
major subsidence of the Surma basin did not take place during the Miocene, rather
the Pliocene uplift of the Shillong Plateau and gradual westward encroachment of
the IBR might have subsided the Surma basin during the Pliocene (Johnson and
Alam 1991; Uddin and Lundberg 2004). Along the southern edge of the basin, a
major NE-SW oriented fault with a length of >150 km has been reported as the
Sylhet Fault (Kayal 1998; Bhattacharya et al. 2008; Kayal 2008; Angelier and
Baruah 2009; Vaccari et al. 2011; Mohanty et al. 2013, 2014), which probably
controls the course of the Surma River (Ovi et al. 2014).
Several investigations have noticed that the Surma basin, a complex conspicuous
sub- basin of the Bengal Basin with thick sedimentary fill (Hiller and Elahi 1988;
Murphy and Staff BOGMC 1988; Uddin and Lundberg 2004; Singh et al. 2016),
has evolved from a passive continental margin with sedimentary packages thick-
ening south, to a flexural basin with sedimentary packages thickening north
(Johnson and Alam 1991; Uddin and Lundberg 2004; Bracciali et al. 2015). These
changes of thickness of sedimentary packages are interpreted to be related to
loading from the adjacent uplifting Shillong Plateau. According to Najman et al.
(2016), passive margin configuration of the Surma basin with southward thickening
of the sedimentary packages ceased at the end of the Bokabil Formation deposition
(3.5 Ma). Transition of the Surma basin from passive margin to flexural basin
occurred during 3.5 to *2 Ma. At this time, the Tipam Formation deposited, which
does not show any lateral variation in thickness, and has sub-horizontal dips in the
basin centre. This also marks the onset of flexural loading by the Shillong Plateau.
The IBR also propagated westward in this area at this time, as evidenced by
thinning of Tipam Sandstone over the IBR anticlines. The combined influence of
the uplifting Shillong Plateau and westward encroachment of the IBR together
resulted the palaeo-Brahmaputra diversion away from the Surma basin, i.e. from the
east to west of the Shillong Plateau by the end of the Tipam Formation deposition
*2 Ma (Najman et al. 2012, 2016). After the drainage diversion, sedimentation in
the flexural Surma basin continued with the meandering facies of the Dupi Tila
Formation, sourced primarily by recycling of Himalayan-derived materials from the
sedimentary cover of the rising Shillong Plateau to the north (Evans 1932; Biswas
172 M. S. Hossain et al.
1961; Hiller and Elahi 1984; Najman et al. 2016). From the Dupi Tila Formation
onward (*2 Ma–Recent), sedimentary packages in the flexural Surma basin has
been thickening as well as dipping to the north. On the other hand, the
pre-Oligocene rock within the Surma basin is the Jaintia Group consisting of
Paleocene Tura Sandstone, Middle Eocene Sylhet Limestone and Upper Eocene
Kopili Shale, and all of which crop out in the northern margin of the Surma basin
(Reimann 1993; Alam et al. 2003). The Tura Sandstone, Sylhet Limestone and
Kopili Shale are interpreted to be deposited in shallow-marine to marine,
shallow-marine carbonate, and deltaic to slope depositional environments, respec-
tively (Alam et al. 2003). Moreover, Oligocene Barial Group (also exposed along
the northern fringe) is believed to be deposited in a predominantly tide-dominated
shelf environment (Alam 1991). In addition, the Surma Group consisting of the
Lower Bhuban and the Upper Boka Bil Formations (Holtrop and Keizer 1970;
Hiller and Elahi 1988; Khan et al. 1988) crop out along the northeastern margin of
the basin. Those deposited in a large, mud-rich delta system that might have drained
a significant portion of the eastern Himalaya through the palaeo-Brahmaputra River
(Johnson and Alam 1991). Sultana and Alam (2001) and Alam et al. (2003)
mentioned that the Surma Group was deposited within a cyclic transgressive–
regressive regime ranging from shallow marine to tide-dominated coastal settings.
The Tipam Sandstone also crops out in the same area and is interpreted to have
deposited in bed-load dominated braided-fluvial systems (Johnson and Alam 1991).
According to Uddin and Lundberg (2004), paleo-Brahmaputra, paleo-Meghna,
paleo-Karnafuli are most likely responsible for the consistent increase in the
thickness of the Miocene sediments in this part of the Bengal Basin as these rivers
delivered detritus from the eastern Himalaya and IBR straight to the Bengal Fan
through the Surma basin and forming a major delta complex at the northeastern part
of the Bengal Basin.
According to Holtrop and Keizer (1970) and Hiller and Elahi (1984), the Surma
basin was structurally engraved in almost sub-Recent geological time, i.e. since 3–
6 Ma and the tectonic movements are presumably still going on, as indicated by
shallow lakes between the anticlinal ridges, especially during the monsoon period.
Recent geodetic measurements (e.g., Nielsen et al. 2004; Vernant et al. 2014;
Steckler et al. 2016) and geochronology studies (e.g., Biswas and Grasemann
2005a; Khan et al. 2006) also suggest that oblique subduction of the Indian Plate to
the NE direction is going on and the basin is still tectonically active. This tectonic
convergence results in complex anticlinal and synclinal systems as well as over-
lapping thrust systems in and around the Surma basin due to overall E-W and N-S
shortening of the basin at the rate of 7 mm/year and 18 mm/year to the northern and
eastern margins, respectively (Akhter et al. 2010; Steckler et al. 2012; Bulbul
2015). Moreover, the northern margin of the basin is subjected to flexural loading
due to the Shillong Plateau. The above tectonic activities also result overall sub-
sidence of the Surma basin currently at the rate of approximately 7–12 mm/year
(Reitz et al. 2015). Primary structures of the basin is dominantly anticlines whose
orientation changes from N-S (e.g., Rashidpur, Maulvi Bazar, Fenchuganj,
Habiganj, Kailas Tila) at the southern part to the ENE-WSW (e.g., Patharia, Sylhet,
Synthesis of the Tectonic and Structural Elements … 173
Jalalabad, Atgram) to the middle-eastern part (Fig. 6). Most of these anticlines are
plunging to the north. To the northern part of the basin, the anticlines show
approximately E-W oriented Type 1 folding (e.g., Chhatak, Gobamura, Dupi Tila)
probably due to two sets of compressional forces, i.e. shortening direction one is
N-S and the other is E-W, acting simultaneously (Biswas and Grasemann 2005b;
Ovi et al. 2014). Some of these anticlines are fault controlled (Sikder and Alam
2003; Steckler et al. 2008; Najman et al. 2012), and few of them are dissected by
the approximately N-S and E-W oriented faults (Hiller and Elahi 1984). The Surma
basin has been studied extensively as a result of successful hydrocarbon explo-
rations (Holtrop and Keizer 1970; Woodside 1983; Shamsuddin 1989; Chowdhury
et al. 1996; Khanam et al. 2017). Anticlinal structures are the principal type of
hydrocarbon trap in the Surma basin (Lietz and Kabir 1982). The Miocene to
early-Pliocene shales up to a depth of 4.5 km are the source of gas/condensate
accumulated in the sandstone of the Bhuban and Bokabil Formations (Shamsuddin
1989).
The Madhupur High to west and the Tripura Uplift to the east are connected by a
threshold known as the Madhupur-Tripura High/Tangail-Trupura Uplift. This sepa-
rates the Surma basin to the north and the Faridpur Trough to the southwest
(Mirkhamidov and Mannan 1981; Reimann 1993; Guha et al. 2010). The Madhupur
High, aligned approximately NW-SE, forms the south-western margin of the Surma
basin which is the northernmost sub-basin of the Bengal Foredeep. The Madhupur
High is a Pleistocene Terrace (Morgan and McIntire 1959), which represents slightly
undulated and elevated topography from the adjacent flood plains, and is an exposed
Quaternary interfluve between two pathways for the Brahmaputra River. Gravity and
aero-magnetic data suggest that the basement is relatively uplifted beneath this area
(Reimann 1993; Guha et al. 2010). Dissection of the initial Madhupur High is resulted
from the Late Pleistocene monsoon climatic episodes that caused a tremendous cur-
rent flow over the plain (Monsur 1995). According to Guha et al. (2010), the Nagarpur
dome and the Nandina high are considered as protrusions from the basement without
deforming the overlying sedimentary packages above this tectonic element. Although
its upper part is heavily eroded, the crests of the hill keep flat surfaces which are gently
tilted to the east (CDMP II 2013). A series of en-echelon faults, known as the
Madhupur Fault flanks the western side of the uplifted Madhupur High/Madhupur
Tract. These en-echelon faulting resulted either from torsion of the region, possibly as
rotational faulting (Mukherjee and Khonsari 2017) or from the effect of shear (e.g.,
Mukherjee 2014a) along an assumed blind fault or possibly a combination of both
(Morgan and McIntire 1959). The upliftment of the Madhupur High may have exerted
a significant control over the avulsion history of the Brahmaputra/Jamuna River. This
avulsion history is cyclic, with a periodicity of *1800 years, and the position of the
Brahmaputra has fluctuated between east and west of the Madhupur High (Pickering
et al. 2014). It is assumed that the Holocene Brahmaputra avulsion history, to some
174 M. S. Hossain et al.
sequence. To the immediate southeast of the Eocene Hinge Zone near the con-
fluence of the Padma and the Jamuna rivers, the depth of the Sylhet Limestone is
6500 m (Guha et al. 2010). This tectonic element is bounded to the north by the
Madhupur-Tripura High and to the southeast by the Barisal-Chandpur High
(Fig. 8). Within this tectonic element, Chalna and Bagerhat are the two notable
structural highs of very low amplitude. While Khan and Agarwal (1993) and Khan
and Chouhan (1996) suggest that this tectonic element is underlain by oceanic crust.
The recent data reveal that oceanic crust is situated more to the east below the
Barisal–Chandpur Gravity High (Alam et al. 2003; Curray 2014; Singh et al. 2016),
and consequently, this tectonic element is underlain most likely by transitional
crust. Sediment thickness in the Faridpur Trough is more than 16 km, and beneath
this sediment, thickness of the transitional extended crust of the Indian Plate is
*16–19 km (Singh et al. 2016). The more or less flat and tectonically less dis-
turbed nature of the Moho has been observed at 31–32 km depth beneath this
tectonic element (Fig. 8) (Khan and Agarwal 1993; Singh et al. 2016). However,
GPS derived geodetic data from Khulna area indicate that the overall subsidence
rate is high (*8 mm/year) here (Reitz et al. 2015). Higgins et al. (2014) assumed
that this high rate appears to be related to the sediment properties as the Holocene
organic-rich muds: notably compaction (Mukherjee and Kumar in press; Dasgupta
and Mukherjee submitted).
It is a tectonically uplifted zone of the Bengal Foredeep Basin that separates the
Faridpur Trough to the northwest and the Hatia Trough to the southeast (Fig. 8)
(Bakhtine 1966; Guha 1978; Uddin and Lundberg 1999). Reimann (1993) mentioned
that the Barisal-Chandpur High can be interpreted as a gravity and magnetic anomaly
caused by a magnetic body at a great depth. The gravity maxima of the
Barisal-Chandpur High shows NE-SW trend. The average width of this tectonic
element is *70 km, and corresponds to a slightly uplift of the overlying sedimentary
cover. Muladi, Kamta and Daudkandi are the three main gentle anticlinal structures
located in this tectonic element. The arc shape deformation front (Maurin and Rangin
2009a; Steckler et al. 2016) of the Bengal Basin passes approximately N-S through
this tectonic element with slight concavity to the E. It is assumed that the continent-
oceanic crust boundary of the Indian Plate in the Bengal Basin is located beneath the
Barisal–Chandpur High (Curray 2014; Singh et al. 2016). Recent GPS derived
geodetic data suggest variable subsidence rate of this High (Reitz et al. 2015). The
northern part of this tectonic element (in Dhaka) shows high subsidence rate
(>10 mm/year) compared to the southern part near Patuakhali (*3 mm/year). The
very high subsidence around Dhaka city, which is situated at the southern edge of the
Madhupur High may be related to the extensive groundwater extraction in and
around the city and consequent rapid fall of the groundwater table (Bhuiyan and
Hossain 2006; Hoque et al. 2007; Akhter et al. 2010), but cannot represent the natural
subsidence rates over the Holocene timescale (Reitz et al. 2015). Nevertheless, if
176
Fig. 8 Schematic cross-section across the Bengal Basin and its surroundings (drawn approximately W-E along the line XY as shown in Fig. 1)
(modified after Murphy and Staff of BOGMC 1988; BOGMC 1997; Maurin and Rangin 2009a and Wang et al. 2014)
M. S. Hossain et al.
Synthesis of the Tectonic and Structural Elements … 177
such high rate of subsidence continues, channel steering behavior (Reitz et al. 2015)
is most likely going to occur in near future around the Dhaka city. However, based on
the study of the Synthetic Aperture Radar (SAR) images, Higgins et al. (2014) points
out that the ground subsidence in and around Dhaka varies significantly, and appears
to be related to the sediment properties with the lowest rates observed in the
Pleistocene Madhupur Clay and the highest rates in Holocene organic-rich muds.
Within the city, the subsidence rate ranges 0 to >10 mm/year, and outside the city the
subsidence rate ranges 0 to >18 mm/year.
The Hatia Trough of the Bengal Basin is located in the southeastern part of the
basin and forms the outermost part of the west-propagating CTFB, situated in
between CTFB to the east and the Barisal-Chandpur High to the west (Fig. 6). The
base of this tectonic element is formed by oceanic crust of the Indian Plate (Khan
and Chouhan 1996; Curray 2014; Singh et al. 2016). This tectonic element rep-
resents the deepest trough of the Bengal Foredeep Basin and has received the
highest accretion of clastic sediments. Sediment thickness in the Hatia Trough
exceeds 18 km, and beneath this sediment, the Indian Plate is not more than 16 km
thick (Singh et al. 2016). As pointed by Najman et al. (2016), the Himalaya forms
the dominant source for the Late Neogene-Recent sediments in the Hatia Trough.
Small amount of arc-derived sediments from the Trans-Himalaya or Paleogene
Indo-Burman Ranges are also reported. Data from the Shahbazpur 1 well in the
Hatia Trough indicate the presence of more than 2000 m of Plio-Pliestocene sed-
iments overlain by 480 m thick Holocene sediments (Alam et al. 2003). As in
Faridpur Trough, the more or less flat and tectonically less disturbed Moho has also
been observed at 31–32 km depth beneath this tectonic element (Khan and Agarwal
1993; Singh et al. 2016). Guha et al. (2010) claimed that the axis of the Bengal
Foredeep runs through the apex of this tectonic element. Char Kajal, Shahbazpur,
Kutubdia/Magnama, Sangu and a number offshore anticlinal structures are located
in this tectonic element, of which Sangu and Shahbazpur yield hydrocarbons.
Recently obtained GPS derived geodetic data suggest variable subsidence rate for
the Hatia Trough (Reitz et al. 2015). The northern onshore part of this tectonic
element (in Raipur, Noakhali) shows lower subsidence rate (*3 mm/year) than the
southern offshore part far from the Magnama structure (*5 mm/year).
The Bay of Bengal is the northeastern lobe of the Indian Ocean (Curray 2014)
bordered to the west and the north by the passive continental margin of the Indian
east coast and southern edge of the Bengal Basin, respectively (Fig. 2). It is bor-
dered by the oblique subduction zone between the Indian Plate to the west and the
Burmese/Sunda Plate to the east, and extends into the northernmost part of the
178 M. S. Hossain et al.
Indian Ocean to the south. The entire Bay of Bengal is floored by the Bengal Fan
(Curray and Moore 1971; Curray 2014). Along 16°N, the Bay of Bengal is about
1400 km in width. According to Desa et al. (2013), the Bay of Bengal is the drifted
half part of the Early Cretaceous Enderby Basin formed as a result of the Kerguelen
hotspot (Banerjee et al. 1995) as Australia drifted away from India during Early
Cretaceous (*132 Ma) with a fan-shaped expanse of ocean floor, separating the
two plates by gradually widening to the east (Ali and Aitchison 2008). Therefore,
the crust of the Bay of Bengal can be considered as a drifted part (Rangin and
Sibuet 2017). Gravity data revealed the presence of NE-SW trending highs and
lows in the western part of the Bay of Bengal along the coast of the Mahanadi-
Krishna-Godavari (KG)-Cauvery basins. NNW-SSE transverse oceanic fracture
zones truncated and offset the NE-SW trending gravity lows, which probably
correspond to the graben infilling of the detrital sediments (Rangin and Sibuet
2017). Along the western part of the Bay of Bengal oldest sediments deposited are
of Early Cretaceous in age (Desa et al. 2013).
Two aseismic ridges namely 90°E Ridge and 85°E Ridge are observed on the
floor of the Bay of Bengal. The 90°E Ridge is largely present in the eastern part of
the Bay of Bengal. Quaternary to Eocene sedimentary sequence was locally
deposited on top of the volcanic rocks that form 90°E Ridge. The volcanic rocks in
the 90°E Ridge are attributed to the drifted Kerguelen Hotspot Igneous Province
(Coffin et al. 2002), which show a general decreasing age trend (90–38 Ma) from N
to S (Royer et al. 1988). Presence of exhumed gabbro along the 90°E Ridge
suggests that the Ridge was formed during the Early Cenozoic along the previously
existing transform boundary due to injection of underplated spreading centre vol-
canism (Rangin and Sibuet 2017). The northern end of this ridge is traceable up to
19°N (Maurin and Rangin 2009b), but further northward of the Bay of Bengal, the
gravity as well as topographic signature of the ridge disappear completely (Rangin
and Sibuet 2017).
The 85°E Ridge (Sar et al. 2009) is an approximately linear and aseismic ridge
west of the 90°E Ridge, and named for its near-parallel strike along the 85°E
meridian (Fig. 2) (Bastia et al. 2010). The ridge extends from the Mahanadi Basin
in the north, shifts westwards by *250 km around 5°N, southeast of Sri Lanka and
then continues to the south in the Indian Ocean. In the northern part, it is blind
beneath the thick sediments of the Bay of Bengal and shows a negative gravity
anomaly and in the southern part, it occasionally rises above the sea floor and is
associated with a positive gravity anomaly (Michael and Krishna 2011). The vol-
canic rocks in the 85°E Ridge may be attributed to the drifted short lived hotspot,
which show a general decreasing age trend ranging from *80 Ma near the
Mahanadi Basin to *55 Ma near the Afanasy Nikitin Seamount (Michael and
Krishna 2011). Curray and Munasinghe (1991) proposed a model to explain the
origin of the 85°E Ridge. They proposed that the origin of Afanasy Nikitin
Seamount, 85°E Ridge, and Rajmahal and Sylhet Traps as the trace of the
Kerguelen hotspot that now lies beneath the Crozet Islands in the southern Indian
Ocean. Moreover, they placed the Kerguelen hotspot, which was also responsible
for the origin of the 90°E Ridge, at the triple junction between Greater India,
Synthesis of the Tectonic and Structural Elements … 179
Australia, and Antarctica before the breakup of eastern Gondwana at *120 Ma.
Based on marine magnetic data combining with petrology, seismic reflection, deep
seismic sounding, and magnetic information of the Bengal Basin, Talwani et al.
(2016) also suggest that Afanasy Nikitin Seamount, 85°E Ridge, and Rajmahal and
Sylhet Traps are related to the Kerguelen hotspot activity. Two sets of oceanic
fracture zones have been found in the Bay of Bengal. The oldest fractures are
transform faults with orientation S50–S60°E, are considered to be of Early
Cretaceous, and occured during an early spreading episode. The youngest fractures
are submeridian and resulted from the northward drift of the Indian Plate since the
end of the Mesozoic (Maurin and Rangin 2009a). Moreover, between 20° and
18°N, the tilted clastic sequences dip southwest and are truncated by the dominant
NE-SW trending listric normal faults (Rangin and Sibuet 2017). It is more or less
well established that the basement in the northern part of the Bay of Bengal is of
oceanic crust. However, Sibuet et al. (2016) and Rangin and Sibuet (2017) argued
that the crust in the northern Bay of Bengal is thinned continental in nature intruded
by volcanic products linked to the passage of the Kerguelen hotspot after the
continental rifting of the northern Bay of Bengal. The Moho is *15 km deep at the
northernmost part of the Bay of Bengal, and it gradually goes deeper to *37 km
below the northern edge of the Bengal Basin (Singh et al. 2016).
In the Bay of Bengal, another major geomorphic unit is observed known as the
Swatch of No Ground (SoNG) (Fig. 2), an active submarine canyon represents the
present submarine drainage system of the Bay of Bengal (Curray et al. 2003). It
serves as a point source for one of the major active fan valley deposits into the Bay
of Bengal/Bengal Fan. In the northern end, the head of the canyon lies in shallow
water depth (*38 m) at 21° 27′N, 89° 41′E, and then continues south for 160 km
as a long, straight trough with an average gradient of 8.2 m/km. At *20°12′N and
89° 12′E, it reaches *1400 m depth. At midshelf, the broad U-shaped trough has a
width of *20 km from rim to rim, and below 862 m from the rims, the width is
8 km (Curray et al. 2003). The canyon continues to the south as an active channel
and reaches as far as the 0° latitude (Curray 2014). The location of the mouth of this
canyon is presumably related to the confluence of the Ganges and the Brahmaputra
rivers during the last glacial low stand sea level, because an active submarine
canyon is related to river mouth migrations, sedimentation, and the Quaternary sea
level fluctuations. Due to sea level rise, the confluence of the Ganges and the
Brahmaputra rivers shifted *90 km further east, therefore, the canyon receives
very little sediments at present (Curray et al. 2003). Currently, the turbidity currents
generated within the SoNG are relatively small, and reach only to the middle
Bengal Fan, due to present highstand sea level (Curray et al. 2003; Curray 2014).
Besides, large influx of sediments mainly from Himalaya and to some extent from
the Indian peninsula is transported by the Ganges and the Brahmaputra rivers,
especially during the northeast and the southwest monsoons (Chauhan and
Vogelsang 2006). The origin of these sediments is dominantly detrital rather than
biogenic (Achyuthan et al. 2014).
180 M. S. Hossain et al.
The Folded Flank of the foredeep (Fig. 9) generally known as CTFB developed
within the upper parts of the thick deltaic sequence, and exhibits large-amplitude,
linear or elongated gravity anomalies (Bakhtine 1966). Structural complexity of the
Folded Flank increases towards east (Khan 1991a, b) and to the farther east, it
merges into the Indo-Burman Ranges (IBR)/Paleogene Inner Burman Ranges. This
folded belt is bounded by the central Foredeep Basin to the west and the IBR to the
east. Although the CTFB gradually merges into the IBR/Mizo Fold Belt, they are
structurally separated by approximately NNW-SSE running Kaladan Fault (Ram
and Venkataraman 1984; Reimann 1993). The strata exposed in this belt are
broadly lithologically similar to those exposed along the southern edge of the
Sub-Himalayan belt of the Himalayas. Although the timing of initiation of the IBR,
in the east, is widely debated, the estimates range from the Late Eocene–Early
Oligocene (Mitchell 1993) to the Late Miocene (Ni et al. 1989), thus costraining the
timing of development of CTFB (Najman et al. 2016). According to Maurin and
Rangin (2009a), a very rapid development and westward propagation of the CTFB
since 2 Ma is facilitated from uplift of the Shillong Plateau, and consequent
deepening and filling of the Surma basin with a large amount of unconsolidated
sediments. The progressive westward overprinting of thin-skinned tectonics by
thick-skinned tectonics within the IBR was due to rapid propagation of the
accretionary wedge westward since Pliocene (Maurin and Rangin 2009a; Najman
et al. 2016). Thus, the age of the structures gradually decreases from east (IBR) to
west (CTFB) and the youngest structures are observed at the western margin of the
CTFB, i.e. near the deformation front. This looks, therefore, like an ‘in-sequence’
deformation. According Wang et al. (2014), many of the folds in the western part of
the CTFB activated only from the Late Pliocene onward, but some of the folds in
the eastern CTFB appear inactive. In this regard, GPS derived geodetic data suggest
variable subsidence and uplift rates for the CTFB (Reitz et al. 2015). In general, the
eastern margin of the CTFB shows very small uplift signature, whereas the western
margin demonstrates small subsidence signature. Although Reitz et al. (2015)
suggest that these variable rates depend on whether the GPS sites are located in
synclines or anticlines, this is unlikely to get subsidence in the CTFB as the area
subjected to continuous convergence collision. It is most likely that the GPS sta-
tions used in their study are positioned in the underthrust flanks (culminate to
synclines) of the anticlines as seems to be the case at least for the Sitakund
Anticline. Based on balanced cross-section across the CTFB, Maurin and Rangin
(2009a) estimated about 11 km total east-west shortening in the past 2 Ma, which
in turn indicates *5 mm/year long-term shortening rate.
The CTFB consists of sediments that were deposited in the Bengal Basin, and
were subsequently uplifted and incorporated into the Neogene accretionary prism
due to the oblique subduction of the oceanic crust of the Indian Plate beneath the
Burmese Plate to the east during Pliocene (Gani and Alam 1999; Najman et al.
2012). It is one of the active orogenic belts in the world susceptible to produce
Synthesis of the Tectonic and Structural Elements … 181
anticlines with pop-up and en echelon character (Maurin and Rangin 2009a). The
widths of the synclines are generally broad in the north compared to those of the
south. The geometrical analysis of the bedding planes of the anticlinal structures
shows that the axial plane of the anticlines strikes 330°–345°, dips 75°–85° either
towards NE or SW, and having interlimb angle of 75°–95°. Plunge varies 5°–7°,
Synthesis of the Tectonic and Structural Elements … 183
JFig. 9 Simplified tectonic map of the Geotectonic Province-3 of the Bengal Basin and its
surroundings (modified after Bakhtine 1966). Faults and lineaments are mostly taken from Biswas
and Grasemann (2005a), Kayal (2008), and Wang et al. (2014). A: Anticline; Athar: Atharamura;
Band: Bandarban; Banu: Banuachari; Barj: Barjala; Bat: Batchia; Bela: Belasari; CCF: Chittagong
Coastal Fault; Chango: Changotaung; Char: Chargola; Dakhin: Dakhin Nhila; Dho: Dholai; F:
Fault; Fench: Fenchuganj; Gajal: Gajalia; Gila: Gilasari; Gobam: Gobamura; Ha: Habiganj; Hach:
Hach Har; Harar: Harargaj; Jumpa: Jumpai/Botiang; Kasa: Kasalong/Kasarang; Kho: Khowai;
Lamba: Lambaghona/Sarta; Langt: Langtari; Machh: Machhlithum; Magn: Magnama; Mahe:
Maheskhali; Mata: Matamuhuri; Moul: Moulvibazar; Mow: Mowdak; Mual: Mualuawm; Olat:
Olathang/Waylataung; Path: Patharia; Pin: Pincha; Ra: Rashidpur; Rajn: Rajnagar; Ram: Ramjtia;
S: Syncline; Sakh: Sakhan; Saku: Sakudaung; Salda: Saldanadi; Sarden: Sardeng/Sabtaung; Semu:
Semutang; Shish: Shishuk/Sisbak; Sita: Sitapahar; St: St. Martin’s; Tich: Tichna; Tulam: Tulamura
and Utan: Uttan Chatra
either to the north or to the south. Some of the prominent anticlines in this zone are
Mowdak, Sakudaung, Uttan Chatra, Barkal, Shishuk/Sisbak, Banuachari, Jumpai/
Botiang.
Middle Asymmetric Thrust Faulted Zone: This zone is situated west of the
highly compressed zone, where overall structural trend is *NNW-SSE. Generally, in
this zone axial plane is steeply inclined to the west, and structures plunge very gentle
towards N. As in the eastern highly compressed zone, the widths of the synclines are
commonly broad in the north than those to the south. Steep limbs of anticlines are
generally faulted. The geometrical analysis of the bedding planes of the anticlinal
structures shows that the axial plane of the anticlines strikes 338°–350°, dips 70°–78°
towards NE, and have interlimb angle of 75°–95°. Plunge varies 5°–10°, either to the
north or to the south (Chowdhury 1970; Hossain et al. 2018). Some of the prominent
anticlines in this zone are Matamuhuri, Bandarban, Gilasari, Belasari, Sitapahar
(Fig. 10a), Gobamura, Changotaung, Kasalong/Kasarang, Sardeng/Sabtaung.
Fig. 10 Lower hemisphere equal area projection of the bedding attitude data shows the
orientation of the fold geometric elements of the a Sitapahar and b Sitakund Anticlines of the
Chittagong Tripura Fold Belt (CTFB), Bengal Basin. Both the anticlines are west verging and
asymmetrical, and show similar structural trend, but the Sitakund anticline is overturned in nature
(Hossain et al. 2018)
184 M. S. Hossain et al.
Western Quiet Zone: This is the westernmost zone of the CTFB with similar
structural trends as in the other two zones. Most of the structures are doubly
plunging with axial planes steeply inclined to the east, i.e. those are asymmetric
overturned structures. The structures have very gentle dips in the crest and steep
dips in the flank. In general, the western flanks of the structures are underthrust,
e.g., Sitakund Structure (Abdullah et al. 2015; Hossain and Akhter 1983). The
geometrical analysis of the bedding planes of the anticlinal structures shows that the
axial plane of the anticlines strikes 335°–350°, dips 55°–65° in the NE direction,
having interlimb angle of 60°–70°. Plunge amount varies 3°–10°, either to the north
or to the south (Hossain et al. 2018). Some of the prominent anticlines in this zone
are St Martin’s Island, Dakhin Nhila, Inani, Olathang/Waylataung, Maheskhali,
Jaldi, Patia, Sarta/Lambaghona, Sitakund (Fig. 10b), and Semutaung. A major fault
is present in this zone running NNW-SSE along the western edge of the Dakhin
Nhila-Inani-Maheskhali-Sitakund structures. Among the three zones of the CTFB,
the western quiet zone is the most important and prospective tectonic sub-element
for oil and gas in Bangladesh, where more than 15 gas fields have been discovered.
Towards north some of these are Feni, Bakhrabad, Titas, Habiganj, Bibiyana,
Rashidpur, Moulvibazar, Fenchuganj, and Beanibazar.
However, recent studies (Wang et al. 2014; Steckler et al. 2016) suggest that
another fold belt exist in between western quiet zone and deformation front, and
stretches from offshore to the south and onshore to the north. This very low
intensity fold belt is characterized by long-wavelength de’collement folds, locally
west verging, and subparallel to the western quiet zone. The amplitude of these
folds decreases westward in accordance to the westward propagation of the
accretionary wedge and die out near deformation front to the west (Maurin and
Rangin 2009a).
Neotectonics and active tectonics of the Bengal Basin are directly related to the
oblique collision of the Indian Plate with the Eurasian Plate to the north and the
Burmese Plate to the east. The collision developed the Himalayan arc to the north,
and the Burmese–Andaman arc to the east. Based on recent seismic studies, it is
presumed that subduction has clogged below the Himalayan arc and is only actively
continuing below the Burmese–Andaman arc (Biswas and Das Gupta 1989; Biswas
et al. 1992; Biswas and Majumdar 1997). Due to the resistance to subduction below
the Eurasian Plate and plate convergence, tectonic deformation and seismic activity
occur in the Indian intra-plate region, including the stable shelf in the western part
of the Bengal Basin, and result in folding and brittle deformation (Mukherjee
2013c, 2014b) in conjunction with high-angle basement faults of variable trends
(NE-SW, E-W, NW-SE) (Khan and Rahman 1992; Khan and Agarwal 1993;
Biswas and Majumdar 1997; Mukherjee et al. 2009). To the north, tectonic loading
is accommodated along the Dauki Fault zone (Bilham and England 2001; Biswas
Synthesis of the Tectonic and Structural Elements … 185
et al. 2007), which consists of a set of high angle, deepseated reverse faults. To the
east, the Cenozoic tectonic evolution of the eastern parts of the Bengal Basin is
related mostly to the NE oblique subduction of the Indian Plate beneath the
Burmese Plate. The collision also results westward migration of accretionary prism
complexes as well as the deformation front (Gani and Alam 1999; Alam et al. 2003;
Wang et al. 2014; Steckler et al. 2016). The CDMP II (2013) reported that the rate
of convergence gradually decreases towards west. These multifaceted neotectonic
evolutions of the Bengal Basin result in several active faults in and around the
Bengal Basin, especially along its northern and eastern margins. Some of these
faults are regional, and capable of generating moderate to great earthquakes. The
rate of motion and the time since the last rupture indicate that the probability of
earthquakes from the existing active faults is high (CDMP II 2013). Overall, the
geotectonic setting of the Bengal Basin at the juncture of three converging litho-
spheric plates with the presence of regional scale seismogenic faults is conducive
for the frequent and recurrent earthquakes.
A general trend of increasing seismic/earthquake activity has been observed in
the Bengal Basin (Khan and Chouhan 1996; CDMP II 2013). Although the
apparent increase of tectonic activity may be attributed to modern and increased
recording facilities, the historical records and the present-day census suggest a
relative increase in tectonic activity (Khan and Chouhan 1996). This seeming
increase of seismicity in the Bengal Basin indicates new fractures/faults propaga-
tion from the preceding seismically undisturbed zones or the reactivation of some
earlier fractures/faults in pre-existing seismically active zones. Catalogues of seis-
micity of high intensity earthquakes (Table 7) in and around the Bengal Basin
(Bilham 2004; Akhter 2010; Martin and Szeliga 2010; Szeliga et al. 2010; Kundu
and Gahalaut 2012; Olympa and Abhishek 2015; Berthet et al. 2014; Kayal 2014;
Wang et al. 2014) show a number of earthquakes that caused not only catastrophic
damages of the property and life but also shifting of the large river courses as well
as dynamic change in landscape. Among such incidents, the Chittagong earthquake
of 1762 (R 8.5+), Sirajganj earthquake of 1787 (MM X), Cachar earthquake of
1869 (R 7.5), Bengal earthquake of 1885 (Mw 6.8), Great Indian/Assam earthquake
of 1897 (Mw 8.1), Srimangal earthquake of 1918 (Mw 7.6), Meghalaya earthquake
of 1923 (Ms 7.1), Dubri earthquake of 1930 (Mw 7.1), Bihar-Nepal earthquake of
1934 (Mw 8.1), Assam earthquake of 1950 (R 8.5), and Nepal earthquake of 2015
(Mw 7.8) are well known. The maximum felt intensity of these earthquakes is
localised in tectonically uplifted region, especially in the Bengal Basin (Martin and
Szeliga 2010). Majority of the earthquakes in the Bengal Basin generally occur at
shallow depth (<30 km). To the east, below the IBR, focal depths gradually
increase in the depth range of 30–70 km and define an eastward gently dipping
subducting plate (Kundu and Gahalaut 2012). Focal depths to the north also show
an increasing depth trend. In general, most of the seismic activities in the Bengal
Basin commonly occur along the regional active faults. The following discussion
on neotectonics and active tectonic of the Bengal Basin are in terms of the neo-
tectonic activities along these faults.
186 M. S. Hossain et al.
Table 7 Seismicity catalogues of high intensity (Mw > 5) earthquakes in the Bengal Basin and
its surroundings
Name of the Earthquake Year of Magnitude (Mw)
occurrence
Seismicity to the East (CTFB)
Compiled from Fergusson (1863, Ansary et al. (2000), Cummins (2007), Akhter (2010), Martin
and Szeliga (2010), Szeliga et al. (2010), Kundu and Gahalaut (2012) and Wang et al. (2014)
Chauk Earthquake (Myanmar) 2016 *6.8
Thabeikkyin Earthquake (Myanmar) 2012 *6.8
Thabeikkyin Earthquake (Myanmar) 2012 *6.8
Bandarban Earthquake (Bangladesh) 2003 *6.1
Moheskhali Earthquake (Bangladesh) 1999 *5.2
Kaladan Earthquake (Bangladesh-Myanmar border) 1955 *6.2
Chittagong/Arakan Earthquake (Bangladesh-Myanmar 1762 R 8.5+
border)
Seismicity to the North (Surma basin and adjacent region)
Compiled from Wesnousky et al. (1999), Bilham and England (2001), Morino et al. (2011),
Chopra et al. (2013), Olympa and Abhishek (2015), Berthet et al. (2014), Morino et al. (2014),
Baruah et al. (2016), Bilham et al. (2017) and Rajendran et al. (2017)
Sikkim Earthquake (India) 2011 *6.9
Assam Earthquake (India) 1950 *8.4
Assam Earthquake (India) 1943 *7.2
Bihar–Nepal Earthquake (India-Nepal border) 1934 *8.1
Dhubri Earthquake (India) 1930 *7.1
Kangra Earthquake (India) 1905 *7.8
Assam Earthquake (India) 1897 *8.1
Cachar Earthquake (India) 1869 *7.5
Dauki Earthquake (Bangladesh-India border) 1548 ?
Nepal Earthquake (Nepal) 1505 *8.7
Seismicity to the West (Foreland Shelf)
Compiled from Bilham (2004), Akhter (2010), Martin and Szeliga (2010), Nath et al. (2014),
Olympa and Abhishek (2015)
Pabna Earthquake (Bangladesh) 1935 *6.2
Hinge Zone Earthquake (Bangladesh) 1935 *6.2
Rajshahi Earthquake (Bangladesh) 1842 *7.3
Hinge Zone Earthquake (Bangladesh) 1842 *7.3
Sirajganj Earthquake (Bangladesh) 1787 MM X
Kolkata Earthquake (India) 1737 ?
Seismicity at the Centre of the Bengal Basin
Compiled from Akhter (2010), Martin and Szeliga (2010) and Nath et al. (2014)
Mymensingh Earthquake (Bangladesh) 2008 *5.1
Bengal (Manikganj) Earthquake (Bangladesh) 1885 *7
Mymensingh Earthquake (Bangladesh) 1846 *6.2
Sirajganj Earthquake (Bangladesh) 1845 *7.1
Dhaka Earthquake (Bangladesh) 1812 MM VIII
Synthesis of the Tectonic and Structural Elements … 187
The CTFB and the IBR fold belt regions are poorly explored due to the complex
geology, dense vegetation and mountainous terrains thus making it challenging for
field observation as well as geological and geophysical data acquisition. The CTFB
or Geotectonic Province 3 is an active tectonic element in the eastern part of the
Bengal Basin (Cummins 2007; Steckler et al. 2016). In the CTFB and the IBR fold
belt regions, five regional faults with *N-S to NNE-SSW orientation are recorded
(Nandy 2001; Acharyya 2007; Maurin and Rangin 2009a; Kundu and Gahalaut
2012; Hossain et al. 2014; Wang et al. 2014), which mostly accommodate this
convergence. Towards west, these are Sagaing Fault—west of Shan Plateau and
east of Central Burma Basin (CBM) as well as Chindwin Basin, Kabaw Fault—
west of CBM and east of IBR, Churachandpur Mao Fault (CMF)—at the core/
centre of the IBR, Kaladan Fault—west of the IBR and east of the CTFB, and the
Chittagong Coastal Fault (CCF)—at the western margin of the CTFB and east of
the Hatiya Trough (Figs. 1 and 8). CMF and CCF are also known as the Lelon Fault
(Maurin and Rangin 2009a), and the Chittagong-Cox’s Bazar Fault, respectively.
Geotectonic evolution and neotectonic activities in the eastern Bengal Basin,
Central Burmese Basin, Chindwin Basin and Rakhine Basin of Myanmar are
directly related to these four major faults which are parallel to the CTFB-IBR fold
belts. A shallow dipping megathrust structurally connects with the CCF, Kaladan
and CMF faults (Wang et al. 2014; Steckler et al. 2016). Although variable con-
vergence rate is reported (CDMP II 2013; Barman et al. 2016; Steckler et al. 2016),
the rate gradually decreases from east (IBR) to west (CTFB). Among the total
convergence of *46 mm/year, 13–17 mm/year is taken by the CTFB-IBR fold belt
and the shallow dipping megathrust, and rest being accommodated by the
dextral-slip motion of the Sagaing fault (*20 mm/year) and CMF (*10 mm/year).
The rate of convergence to the east of the Kaladan Fault is *13 mm/year, in
between the Kaladan Fault and CCF is *5 mm/year, and between CCF and
deformation front (Figs. 1 and 9) is *2 mm/year (CDMP II 2013). This rate of
convergence results *200 km advancement of the outer part of the accretionary
prism (represented by the CTFB) over the last 2 Ma (Maurin and Rangin 2009a).
The orientation of maximum compressive stress rotates from NE–SW across the
inner and northern IBR to E–W near the western CTFB of the Bengal Basin
(Angelier and Baruah 2009). The rotation of maximum stress orientation is con-
sistent with the deformation partitioning related to oblique collision and is reflected
in the rotation of relative displacement vectors across the region (CTFB and IBR).
This deformation partitioning results a variety of movements across the fold belt.
Towards west, right-lateral slip is observed in the east of the IBR (Sagaing Fault) to
oblique reverse-dextral slip in the middle and western edge of the IBR (Kabaw
Fault) and pure thrusting in the CTFB (CCF) (Fig. 8). Regardless of the highly
oblique plate motion and thick Tertiary sediments, the GPS measurements suggest
that subduction in this region to be active (Steckler et al. 2016). GPS studies also
suggest a locked megathrust signifies an underappreciated large seismic hazard in
the CTFB and IBR regions.
188 M. S. Hossain et al.
The *N-S oriented and *1400 km long Sagaing Fault is the major locus of
dextral-slip motion related to the northward translation of the Indian Plate against
the Burmese Plate (Wang et al. 2014). The northern end of this transform fault
develops horse-tail appearance and terminates into the eastern Himalayan syntaxis
(Fig. 1). This fault accommodates about 20 mm/year of the north component of the
relative motion of the oblique convergence between the Indian and the Burmese
Plates (Barman et al. 2016). The fault is well expressed geomorphically along most
of its length due to such high slip rate. Earthquake catalogues (Table 7) show that
about half of the Sagaing Fault has ruptured during many large earthquakes over the
decades. Based on bends, splays, terminations of historical ruptures, and distinct
secondary features, Wang et al. (2014) divide the Sagaing Fault into six distinct
segments. From south to north, these segments are Bago, Pyu, Nay Pyi Taw,
Meiktila, Sagaing, and Northern Sagaing. The Northern Sagaing segment is again
divided into three sub-segments, and these are Tawma-Ban Mauk, In Daw-Mawlu
and Shaduzup-Kamaing-Mogang. Most of these segments of the Sagaing Fault are
seismically active and produce major earthquakes except the Meiktila segments
(Wang et al. 2014). About 13 major earthquakes (Mw 6.8–7.7) have been recorded
since 1929 along this mega fault and most of them occurred at shallow depth
(<25 km). The most recent Thabeikkyin earthquake with 6.8 Mw occurred in
November 2012 in the Sagaing segment, which caused damage and panic as far as
in the Chittagong City of Bangladesh, approximately 400 km west of the epicentre.
A notable seismic gap along the Sagaing Fault is observed in the Meiktila segment,
between the new capital of Myanmar (Nay Pyi Taw) and the large city of
Mandalay. If this 220 km long Meiktila segment ruptures at once, it is capable of
producing an earthquake as large as Mw 7.8–7.9 (Blaser et al. 2010).
The oblique reverse-dextral slip and the east-dipping Kabaw Fault (Fig. 1) tra-
verses the eastern side of the Kabaw valley for *800 km. The shortening rate to
the locked Kabaw Fault is *13 mm/year (Steckler et al. 2016). Based on the
youthful appearance (fresh/unweathered exposure) on SRTM and ASTER images, a
segment of this fault between 22 and 24.8°N is considered as (seismically) active
(Wang et al. 2014). If this *280 km segment of the Kabaw Fault ruptures at once,
it could generate an Mw 8.4 earthquake. The most recent earthquake with 6.8 Mw
occurred in August 2016 along this fault near Chauk City of Myanmar. The tremors
were felt as far as in Bangladesh, Bihar (India), and Thailand and damaged
infrastructures and life.
The N-S to NNE-SSW oriented and *300 km long right-lateral Churachandpur-
Mao Fault (Fig. 1) is the most prominent at the core/centre of the IBR (Fig. 1). This
steeply east dipping fault accommodates dextral velocities of *10–18 mm/year
between the Indian and Burmese Plates through dextral strike-slip motion, predom-
inantly through aseismic slip (Gahalaut et al. 2013; Steckler et al. 2016). However,
both aseismic and seismic slips have been observed along this fault. The aseismic slip
part of this fault could be related to minor to moderate earthquakes. Wang et al. (2014)
suggest that this fault creeps only at sub-surface, and that its northern part is seismic. If
this fault segment ruptures, it could produce an Mw 7.6 earthquake. Although
earthquake catalogue does not show any historical seismic events that could be
Synthesis of the Tectonic and Structural Elements … 189
attributed to rupture of the Churachandpur-Mao Fault (Szeliga et al. 2010), the Imphal
earthquake with 6.8 Mw occurred in January 2016 in the northern segment of this
fault is a proof of seismic slip. This earthquake was strongly felt in Bangladesh and
caused significant damage.
The *450 km long west verging Kaladan Fault (Figs. 1 and 9), which mostly
follows the course of the Kaladan River is basically the contact between the Barail
and Surma range and is named after the Kaladan River (Nandy 2001; Maurin and
Rangin 2009a; Barman et al. 2016). Kaladan Fault marks the eastern boundary of
the CTFB and extends from the Arakan coast in the south to the northern-most part
of the CTFB and the IBR contact (Sikder and Alam 2003; Maurin and Rangin
2009a; Kundu and Gahalaut 2012; Wang et al. 2014). Along 20° 30′N, the dip of
the fault plane is 70°E with an 8°–10°S pitch, and shows a clear right-lateral offset.
In this area, the orientation of the maximum compression is *N80–N60°E (Maurin
and Rangin 2009a). Although, the Kaladan Fault shows both dextral strike-slip and
west verging thrust components, it is generally considered as thrust fault, and
seismically the fault is sparsely active. The right-lateral slip dominates its northern
termination whereas in the middle portion pure E-W or WSW-ENE thrusting focal
mechanisms are observed, suggesting that active motion is partitioned within a very
short distances along this fault. To the south, Ramree thrust is most likely the
southward continuation of the Kaladan Fault (Maurin and Rangin 2009a). While
Mw 6.2 earthquake in December 1955 along the Kaladan Fault near
Bangladesh-Myanmar border region at a depth of 35 km is a proof of seismic slip,
mainly shallow focus and low intensity earthquakes are observed along this fault
(Kundu and Gahalaut 2012).
Based on the geomorphic signatures on 30 m resolution SRTM image and
geological field reconnaissance, a *300 km long major fault can be inferred along
the western boundary of the CTFB in between *19° 40′N and 23° 30′N is known
as the Chittagong Coastal Fault (CCF). Multichannel seismic survey in the offshore
area also suggests the existence of the CCF as a crustal dextral strike-slip fault with
an appreciable thrust component, and is resultant product of thin-skinned and
thick-skinned tectonics (Maurin and Rangin 2009a). The fault is seismically less
active. To the north, the fault splays before disappearing near the Tichna anticline in
Tripura, just south of the Surma basin. The fault is probably younger than the
Kaladan Fault as the accretionary wedge is progressing westward during the last
2 Ma, and it is not affected by sufficient erosion yet. Maurin and Rangin (2009a)
suggest that both the CCF and the Kaladan Faults can be interpreted as deep-seated
basement reverse fault with dextral strike-slip component. The restoration of the
compressive fold belt and megathrust yields *11 km of shortening of the CTFB
including the CCF and the Kaladan Faults. Although published earthquake cata-
logues do not show any historical seismic events that can be attributed to rupture of
the CCF, Moheskhali earthquake with 5.2 Mw occurred along the CCF in
Moheskhali Island in July 1999 is a proof of seismic slip. This earthquake was
strongly felt around Bangladesh and caused significant damage locally (Ansary
et al. 2000). Moreover, occurrence of any large earthquake in the offshore area
along the east coast of the northern Bay of Bengal is likely to produce great tsunami
190 M. S. Hossain et al.
(Cummins 2007). One of such large earthquakes (R 8.5+) occurred in 1762 along
the Arakan coast (i.e. Arakan subduction zone) known as Chittagong/Arakan
earthquake (Fergusson 1863; Cummins 2007; Martin and Szeliga 2010; Kundu and
Gahalaut 2012; Wang et al. 2014). It is believed that rupture of the megathrust of
the Ramree domain through the southern portion of the Dhaka domain caused this
earthquake. The fault model to reproduce the observed subsidence and uplift
associated with this earthquake along the eastern coast of the northern Bay of
Bengal suggests that the fault’s upper edge coincides with the deformation front.
Cummins (2007) concluded that the evidence of oblique convergence and associ-
ated active subduction in conjunction with thrust earthquake activity along the
coastal region suggest the risk of a major tsunami in the northern Bay of Bengal.
The CTFB and IBR are also segmented by two major conjugate oblique/
transverse faults. The NW–SE oblique Mat Fault is sinistral, whereas the NE–SW
oblique Gumti Fault is dextral (Fig. 9) (Nandy 2001; Barman et al. 2016).
However, seismicity along these faults is not reported. In addition, medium to small
scale dip-slip, strike-slip and thrusts are also present in the CTFB area that have
played important role not only in shaping the present topography of the folded flank
of the Bengal Basin, but also controlling the stream courses. The dip-slip faults are
mainly oblique to the strike of the anticlines, whereas strike-slip and thrust faults
are mostly parallel to the strike. The latter faults mostly affected the folds of the
CTFB regions (Sikder and Alam 2003; Mandal et al. 2004; Kabir and Hossain
2009; Maurin and Rangin 2009a; Najman et al. 2012; Abdullah et al. 2015; Hossain
et al. 2018). A single large and few moderate magnitude earthquakes are recorded
in the CTFB region, which may be connected with these faults. Among such events,
the Srimangal earthquake in 1918 (Mw 7.5), the Bandarban earthquake in 1997
(Mw 6.1), and Barkal earthquake in 2003 (Mw 5.7) are well known (Akhter 2010).
On the other hand, based on the presence of fault scarp on the young fluvial/
alluvial surface at the flanks of the anticlines, some of the anticlinal structures in the
CTFB region are considered as tectonically active (CDMP II 2013). Among such
structures NNW trending Harargaj, Sitakund, Jaldi, Sandwip, Maheshkhali, and
Dakhin Nhila anticlines are well recognized. The seismic reflection profiles across
few of these anticlinal structures reveal growth strata that constrain the initiation of
folding (Mandal et al. 2004; Steckler et al. 2008; Najman et al. 2016). The age of
the growth strata of the Habiganj, Jaldi, and Kutubdia structures constrains the
initiation of anticlinal growth during Pliocene or younger (Johnson and Alam 1991;
Steckler et al. 2008; Maurin and Rangin 2009a; Wang et al. 2014). According to
Khan et al. (2015), the Dakhin Nhila structure is tectonically active and show
different uplift rates spatially and temporally due to faulting for the last
55,000 years. More recent activity recorded from young- aged soils from the crestal
portion of the Maheshkhali and Jaldi anticlines suggest uplifts of 18,000 and
35,000 years respectively (Khan et al. 2005). Farther north, the upper surface of the
Sandwip structure is estimated to be about 7000 years old as indicated by radio-
carbon date (Goodbred and Kuehl 1999). Similar young neotectonic activities are
also reported by Ansary et al. (2000), Steckler et al. (2008), Khan et al. (2017,
2018) in this region.
Synthesis of the Tectonic and Structural Elements … 191
The geodynamics and the neotectonics of the northern part of the Bengal Basin
(northern part of the Geotectonic Province 1 and 2) are directly related to the
tectonic evolution of the 2500 km long Himalayan frontal belt as well as to the
Main Frontal Thrust. The Himalayan frontal belt is the youngest portion of the
Himalayan Orogeny (Thakur 2004). The northward convergence of the Indian Plate
resulted in crustal shortening of the northern margin of the Indian continent,
accommodated by south-verging Main Central Thrust (MCT), Main Boundary
Thrust (MBT) and Main Frontal Thrust (MFT) (Fig. 7) (Thakur 2004). From MCT
to MFT, these thrusts are progressively younger in age and shallowing at depth in a
southward direction, suggesting southward migration of the Himalayan deformation
front. Exceptions to this are the Himalayan active/out-of-sequence structures, most
notably faults (Mukherjee 2015), some of which are even backthrust-faults (Bose
and Mukherjee, submitted-a, b). The Himalayan frontal belt is limited to the south
by the Main Frontal Thrust (MFT). Most of the collisional force in the Himalayan
frontal belt is accommodated by the uplift and convergence along the MFT
(Wesnousky et al. 1999). Although variable convergence rates (11–21 mm/year)
are observed across the Himalayan frontal belt, a general increasing trend is
observed towards the centre which is known as Nepal Himalaya (Wesnousky et al.
1999; Berthet et al. 2014; Bilham et al. 2017). Towards east, the Himalaya, except
in Pakistan, can be divided into Dehra Dun Himalaya (western Himalaya), Nepal
Himalaya (central Himalaya) and Bhutan-Arunachal Himalaya (Eastern Himalaya)
(Thakur 2004). Here after Sikkim, Bhutan and western Arunachal are collectively
referred as the Bhutan Himalaya. The rate of convergence along the Nepal
Himalaya is *20 mm/year and along the Bhutan Himalaya (directly north of the
Bengal Basin) is *16–18 mm/year. The Himalayan frontal belt is one of the few
collisional orogenic belts on-land, where earthquakes with great magnitudes anal-
ogous to those of oceanic subduction zones generally occur (Berthet et al. 2014).
The northern part of the Bengal Basin mainly comprises of the Surma basin to
the east, and northern edge of the Stable Shelf to the northwest. Assam Basin and
Shillong Plateau are situated in between Himalayan frontal belt to the north and
Surma basin to the south. Broadly, E-W oriented Brahmaputra fault is situated at
the southern margin of the Assam Basin. Oldham, Brahmaputra and few other faults
separate the Assam Basin from the Shillong Plateau, whereas the Dauki Fault
separates the Shillong Plateau from the Surma basin. Southern portion of the Surma
basin is transacted by two major faults, the Sylhet Fault to the southeast and the Old
Brahmaputra Fault to the southwest (Fig. 6). In the west, the Shillong Plateau is
separated from the Bengal Basin by the Jamuna Fault (Dhubri Fault). In the east, the
Shillong Plateau and the Assam Basin separate from the Naga domain/Naga Hills
by the Naga and Disang thrusts. In addition, a few other faults and lineaments also
play important role in neotectonic activities as well as geomorphic evolution of this
part of the Bengal Basin, and these are Dapsi (NW-SE), Haflong (*E-W), Dhudnoi
192 M. S. Hossain et al.
(N-S) and Kopili Faults (NW-SE), and Tista and Gangtok Lineaments (NW-SE)
(Fig. 1).
Approximately E-W oriented MFT separates the Assam Basin from the
Himalayan frontal belt. The MFT is associated with active low-angle thrusting and
associated uplift (Berthet et al. 2014). It represents a zone of active deformation
between the Sub-Himalaya and the Indo-Gangetic Alluvial Plain (IGAP) of the
Peninsular India, thereby thrusting sediments of the Siwalik Group southward over
the IGAP (Wesnousky et al. 1999). The MFT comprises of mainly homoclinal
strata with northerly dips generally ranging 20°–30°. In the Bhutan Himalaya, the
dip of the MFT is *25°–30° (Wesnousky et al. 1999; Long et al. 2011; Berthet
et al. 2014) and the Holocene slip rate along the fault is *20–28 mm/year (Burgess
et al. 2012; Berthet et al. 2014). The observed uplift rate across the MFT is *10–
13 mm/year (Wesnousky et al. 1999; Berthet et al. 2014). The MFT is tectonically
active in the Bhutan-Arunachal, Nepal, and Dehra Dun Himalaya (Wesnousky et al.
1999; Berthet et al. 2014). Although some of the N-S plate convergence motion is
accommodated by the upliftment of the Shillong Plateau, however they do not have
any significant effect on the high activity of the MFT in the Bhutan Himalaya. The
rupture zones of 1905 Kangra earthquake (Mw *7.8), 1505 western Nepal
earthquake (Mw *8.7), 1934 Bihar–Nepal earthquake (Mw *8.1), and 1950
Assam earthquake (Mw *8.4) extend N–S right across the MFT zone to the MBT
zone (Fig. 11) (Wesnousky et al. 1999; Berthet et al. 2014; Bilham et al. 2017;
Rajendran et al. 2017). In this context, the *350 km long Bhutan Himalaya zone
Fig. 11 Active tectonic map of the Central and Eastern Himalaya (to the north of the Bengal
Basin). Pink and orange rectangles show the historical events and the ages of major
surface-rupturing events, respectively. Blue and green rectangles give the Holocene shortening
rates and the geodetic slip rates, respectively. Distributions of the earthquake epicentres from 1973
to 2013 recorded by the National Earthquake Information Center (NEIC) are shown in yellow
circles (modified after Berthet et al. 2014)
Synthesis of the Tectonic and Structural Elements … 193
of the MFT, which is located between the epicentres of the 1934 Bihar–Nepal
earthquake and 1950 Assam earthquake, appears to be an area where no similar
great earthquakes has ever been documented. These observations and the lack of
primary surface rupture suggest a potential for great earthquakes along the
Himalayan Frontal Thrust in the Bhutan Himalaya zone (Wesnousky et al. 1999;
Berthet et al. 2014). An active zone of deformation, upliftment and slip observed
along the MFT, especially along the Bhutan Himalaya suggests increased seismic
hazard to the adjoining northern part of the Bengal Basin.
The WNW and ESE oriented Oldham Fault separates the Assam Basin from the
Shillong Plateau. The Oldham Fault is a *120 km long north-verging and south
dipping reverse fault, and generates co-seismic displacement. Although the 1897
devastating Assam earthquake (Mw *8.1) occurred due to co-seismic displace-
ment along this fault (Bilham and England 2001), the source is also attributed to
multiple faults around the Shillong Plateau including the Dauki, Dapsi and Oldham
Faults (Bilham and England 2001; Morino et al. 2014; Barman et al. 2016; Singh
et al. 2016). The Oldham Fault together with the Dauki Fault to the south of the
Shillong Plateau constitutes a high-angle conjugate crustal-scale shear zone,
uplifting the plateau as a pop-up structure (Biswas and Grasemann 2005a). Both the
hanging wall (i.e. Shillong Plateau) and the foot wall (i.e. southern margin of the
Assam Basin/Himalayan Foreland Basin) of the Oldham Fault are associated with
uplift of the Shillong Plateau. However the rate of uplift of the footwall block is
found to be less than that of the hanging wall block. The rate of vertical dis-
placement along the Oldham Fault is 0–0.68 mm/year (Biswas et al. 2007). By
considering *45°–50° dipping for the fault plane, average fault slip rate is esti-
mated to be *2–4 mm/year and the horizontal shortening rate accommodated by
the plateau is *2–5 mm/year (Bilham and England 2001). However, Biswas et al.
(2007) suggest much lower shortening rate (0.65–2.3 mm/year) that accounts for
10–15% contraction rate of the Himalaya. The GPS measurements indicate the
southward movement of the central Shillong Plateau at a rate of *6.5–7 mm/year
with respect to the reference points in southern and central India (Barman et al.
2016). From these rates, Bilham and England (2001) conclude that the recurrence
interval for earthquakes resembling the 1897 Assam earthquake to be 3000–
8000 years. However, examination of the past four high-intensity seismic events
along the northern edge of the plateau indicates a 500 years recurrence interval but
these may record shaking from moderate local earthquakes and large seismic events
in the adjoining Bhutan Himalaya. Geological and geophysical data suggest the
presence of another large fault at the north of the Oldham Fault with E-W orien-
tation known as Brahmaputra Thrust (Ghosh et al. 2015) that marks the northern
most edge of the Shillong Plateau, and may have played an important role in the
plateau uplifting. This fault is an active thrust and important for the crustal reset-
tlement in this part of the Indian Plate.
The *320 km-long E-W trending south-verging and north dipping Dauki Fault,
a major plate boundary-influenced fault marks the boundary between the Surma
basin to the south and the Shillong Plateau to the north (Hossain et al. 2016). The
strike of this fault is parallel to the MFT, which forms a plate boundary fault
194 M. S. Hossain et al.
whereas the Dauki Fault is an intra-plate active reverse fault. The continuous
subsidence of the southern footwall of the Dauki Fault has resulted in the Surma
basin, where structural relief between highest point of the plateau and lowest point
of the basin is estimated to be around 20 km (Mirkhamidov and Mannan 1981;
Hiller and Elahi 1984; Shamsuddin and Abdullah 1997; Biswas et al. 2007). Across
the Dauki Fault, a striking negative Bouguer gravity anomaly exceeding −70 mGal
has been recorded for the Surma basin, whereas the Shillong Plateau shows a
positive Bouguer gravity anomaly >+20 mGal (Rahman et al. 1990b; Rajendran
et al. 2004). Hence, the Dauki Fault might have been formed at the onset of the
exhumation of the Shillong Plateau, and the total vertical displacement of the
basement top along this fault is *10 km (Biswas et al. 2007). During the Late
Tertiary, the rate of upliftment of the hanging wall rock (i.e. the Shillong Plateau)
ranged 0.32–0.80 mm/year, whereas the rate of subsidence of the footwall rock (i.e.
the Surma basin) is *0.45 mm/year. Consequently, the resultant mean vertical
displacement rate along the Dauki Fault ranges 0.77–1.25 mm/year (Biswas et al.
2007). Higher vertical displacement rate along the Dauki Fault than the Oldham
fault is partially responsible for the *2°–3° northward tilt of the Shillong Plateau
surface. The Dauki Fault is segmented to at least 4 parts along its entire length
(Biswas and Grasemann 2005a). Geomorphic indices like mountain front sinuosity,
stream length gradient index, valley floor width to height, drainage basin asym-
metry, and transverse topographic symmetry factor suggest that the tectonic activity
of the fault is much higher in the eastern part than the western portion (Biswas and
Grasemann 2005a; Sharma and Sarma 2017).
Rapid movement along the Dauki Fault started in the Pliocene, resulting in
significant topographic rise and basin loading in the north and south, respectively
(Najman et al. 2016). The flexure related to seismic event has been identified along
the Dauki Fault near Jaflong (Sylhet) and is inferred to be AD 840–920 (Morino
et al. 2014). Another rupturing in the Dauki Fault estimated to be AD 1548 has
been identified in a trench at Gabrakhari village, Haluaghat, Mymensingh (Morino
et al. 2011, 2014). It is also believed that the 1897 Assam earthquake that occurred
along the Oldham Fault also caused slip along the Dauki Fault. Although it was
previously thought that the interval between these giant plateau-building earth-
quakes exceeds 3000 years (Bilham and England 2001), the signature of the above
seismic events indicates recurrence interval of 350–700 years (Morino et al. 2014).
Moreover, the active fan development along the southern foothills of the Shillong
Plateau is controlled by the neotectonic activities along the Dauki Fault (Alam and
Islam 2017).
In the southeastern margin of the Shillong Plateau, a branched extension of the
Dauki Fault with approximately E-W strike is known as the Haflong thrust fault.
The Haflong Fault marks the edge of the Shillong Plateau before it links up with the
Naga thrust to the east (Biswas and Grasemann 2005a). The WNW–ESE trending
*90–100 km long Dapsi thrust fault (Fig. 1) is located within the southwestern
edge of the Shillong Plateau. This fault is the north-western extension of the Dauki
Fault that separates the Cretaceous–Tertiary sediments to the south and the
Precambrian Gneissic complex to the north (Barman et al. 2016). The N-S trending
Synthesis of the Tectonic and Structural Elements … 195
Dudhnoi Fault lying towards the northwest of the Shillong Plateau was activated
during the 1897 Assam earthquake (Barman et al. 2016). A significant difference in
the rate of motions has been observed on the opposite side of the Dudhnoi Fault.
The central Shillong Plateau to the east shows 7 mm/year southward motion,
whereas the western Shillong Plateau shows 2.1–2.7 mm/year southward motion,
suggesting different tectonic settings of the plateau and dextral motion along the
Dudhnoi Fault (Barman et al. 2016). In addition, the development of the Shillong
Plateau as a pop-up structure is also facilitated to the west by the N-S trending
Dhubri and Jamuna Faults. The epicenter of the 1930 Dhubri earthquake (Mw
*7.1) was along the Dhubri Fault near Dhubri, Assam. To the south, the earth-
quake also caused slip in the Jamuna Fault, and west of the Garo Hills, fissures and
sand vents were reported (Olympa and Abhishek 2015).
The NW-SE trending strike-slip Kopili Fault located between the Shillong
Plateau and the Mikir Hills records a dextral slip of 4.7 ± 1.3 mm/year indicating
the fragmentation of Assam Basin/Brahmaputra valley across this fault (Barman
et al. 2016). The eastern edge of the Mikir Hills marks by the Bomdila Lineament
(Baruah and Hazarika 2008; Angelier and Baruah 2009). The transverse Kopili
fault may be a surface expression of the deep-seated seismogenic structure
(Bhattacharya et al. 2008). The western edge of the Mikir Hills appears to be locked
currently at 10.2 ± 1.4 km depth with the Assam Basin indicating strain accu-
mulation. This strain accumulation may cause seismic event in near future along
this seismically highly active 300 km long fault that reaches into the MFT system to
the north and Naga thrust to the south (Bhattacharya et al. 2008; Barman et al.
2016). The rupture area of the 1869 Cachar earthquake (Mw *7.5) and 1943
Assam earthquake (Mw *7.2) are located within the Kopili fault. The earthquakes
damaged over hundreds of kilometres across the region (Olympa and Abhishek
2015).
At eastern Arunachal Himalaya over the Assam Basin and the Naga Hills, three
major plates interact along two convergent boundaries: the Himalayan frontal belt
and the IBR, which meet at the 430 km-long Assam Syntaxis (Angelier and Baruah
2009; Wang et al. 2014). The northern end of the IBR (i.e. Naga Hills) and its
underlying lithosphere experience nearly arc perpendicular extension with ESE–
WNW trends, whereas easternmost Himalayan frontal belt (i.e. Arunachal
Himalaya) experiences WNW-ESE compression related to convergence between
the outer IBR and the Indian Plate. Together, the convergence and the extension
structures forms a complicated geotectonic system (Nandy 2001) that causes the
northeastern portion of the Indian Plate to be deformed and also results in the
development of the Naga and Disang Thrusts (Evans 1964; Bilham and England
2001; Angelier and Baruah 2009). The NE-SW oriented Naga Thrust with several
arcuate lobes separates the Assam Basin from the northern outermost edge of the
Naga Hills (Fig. 1). The fault with clear linear scrap extends from the Haflong
Thrust to the southwest and terminates at *96°E where it appears to cut by a
*NW-striking Mishmi Thrust and associated anticlines (Vaccari et al. 2011;
Mohanty et al. 2013, 2014). The fault shows variable dips (moderate to steep) and
breaks the surface at some positions (Wang et al. 2014). The lower Miocene rocks
196 M. S. Hossain et al.
constituting the Naga Hills are thrusted and folded over undeformed upper Pliocene
to Quaternary beds and alluvium of the Assam Basin at the mountain front along
the Naga Thrust (Aier et al. 2011). The clear geomorphological expression of some
anticlines and younger terrace with 20–50 m upliftment imply that it is still active
and the dip of the thrust fault decreases at depth. Additionally the slip on the Naga
Thrust dies out over *100 km from the crest of the Naga Hills (Wang et al. 2014).
In the immediate south, another linear thrust fault running parallel to the Naga
Thrust is known as the Disang Thrust. The narrow linear belt of imbricate thrust
slices of 8–25 km wide in between the Naga and Disang Thrusts is known as the
Schuppen Belt. Morphotectonic analysis reveals that the overall tectonic activity of
the Schuppen Belt (i.e. along the Naga and Disang Thrusts) is higher than the Dauki
Fault area (Sharma and Sarma 2017).
The NNW–SSE trending Tista and Gangtok Lineaments are the two prominent
tectonic features, extending from the Sikkim Himalaya to the north and Rangpur
Saddle to the south. The southern ends of these lineaments terminate against the
Jamuna Fault. The focal mechanisms in this area indicate right-lateral strike-slip
along the Tista and Gangtok Lineaments. The 2011 Sikkim earthquake (Mw *6.9)
may have been caused by the seismic slip along the Tista Lineament (Chopra et al.
2013; Baruah et al. 2016). Several smaller earthquakes are recorded along these
lineaments, especially along the Tista Lineament suggesting that the both linea-
ments are tectonically active (Baruah et al. 2016).
The Surma basin is an active tectonic element forming in response to the
Shillong Plateau upliftment to the north, and subduction of the Indian Plate beneath
the Burmese Plate to the east. Tectonic evolution and neotectonic activities of this
basin are controlled by the Dauki Fault to the north, gradual westward encroach-
ment of the CTFB and IBR through Haflong, Naga-Disang thrust to the east, the
Sylhet Fault and the CTFB to the south, and the Old Brahmaputra and Jamuna
Faults to the west. Except to the west, the rest of the structural features framing the
Surma basin are the results of mostly compressional tectonic movements in
Pliocene to Recent times (Reimann 1993). The 150-km long Sylhet Fault with
NE-SW orientation, passes through the southern edge of the basin, is a major fault
(Kayal 1998; Bhattacharya et al. 2008; Kayal 2008; Angelier and Baruah 2009;
Vaccari et al. 2011; Mohanty et al. 2013, 2014). This fault also known as Hail
Hayalua Lineament is probably the southwest extension of the Naga-Disang Thrust
(Ghosh et al. 2015) which is offset by the Dauki Fault through its dextral slip
component. In November 2017, a shallow focus (30.3 km depth) earthquake with
Mw 4.9 occurred at distance of 27 km SSW of Habiganj, Bangladesh is probably
related to the Sylhet Fault. Moreover, based on the presence of fault scarp on the
young fluvial/alluvial surface at the flank of the anticline, the Fenchunganj, Sylhet,
Habiganj and Rashidpur anticlines are considered as active structures within the
Surma basin region. In addition, neotectonic activities are also observed in the south
of the Dauki Fault in the Jaintiapur area. In the eastern part, to the immediate south
of the Dauki Fault, some small hillocks present in the Jaintiapur area (Chowdhury
et al. 1996), Sylhet which are capped by Dihing Formation. The Optically
Stimulated Luminescence (OSL) dating of this formation indicates young
Synthesis of the Tectonic and Structural Elements … 197
depositional age (73–24 ka), and variable uplift rate (Khan et al. 2006). This
suggests that the area is tectonically active and gone through a differential uplift-
ments in the recent past (Khan et al. 2018).
The western part of the Bengal Basin (Geotectonic Province 1) mainly comprises
the Stable Shelf, and contains two major tectonic elements: the Foreland Shelf to
the west, and the Eocene Hinge Zone or shelf break or shelf slope to the east
(Reimann 1993). Geographically, the area comprises the western part of
Bangladesh and the West Bengal state of India. Along its longitudinal direction, the
Stable Shelf can be divided structurally into four tectonic elements from northwest
to southeast: (i) the North Bengal Foreland, (ii) the Basin Margin Fault Zone/
western scarp zone, (iii) the Central Stable Shelf, and (iv) the Eocene Hinge Zone
(Nath et al. 2014). Except the Eocene Hinge Zone, alternating gravity highs and
lows indicate the presence of numerous graben and half-graben structures in most
of the Stable Shelf (Khan and Rahman 1992; Khan and Chouhan 1996). This part
of the Bengal Basin is intersected by several faults and lineaments. These are
Katihar-Nilphamari Fault, Jangipur–Gaibandha Fault, Debagram–Bogra Fault,
Rajmahal Fault, Ganga-Padma Fault, Malda–Kishanganj Fault, Medinipur-Farakka
Fault, Chhotanagpur Foot Hill Fault, Damodar Fault, Saithia–Bahmani Fault,
Purulia Shear Zone, Kardaha lineament, and Devikut lineament (Fig. 4) (Nandy
2001; Vaccari et al. 2011; Mohanty et al. 2013; Nath et al. 2014; Roy and
Chatterjee 2015; Prasad and Pundir 2017). The Medinipur-Farakka Fault is also
known as the Basin Margin Fault Zone (Raman et al. 1986; Roy and Chatterjee
2015). This fault zone broadly includes the Medinipur-Farakka Fault to the east,
and Chhotanagpur Foothill Fault to the west. The Rajmahal and Saithia-Brahmani
Faults are situated to the east and west of the Rajmahal Hills, respectively (Ghose
et al. 2017). The approximately N-S running Malda-Kishanganj Fault (Nandy 2001;
Vaccari et al. 2011; Mohanty et al. 2013, 2014; Prasad and Pundir 2017) reaches up
to the east and northeastern margin of the Rajmahal Hills. Rajmahal Fault to the east
of the Rajmahal Hills (Fig. 4) is considered as the northwestern margin of the
Bengal Basin (Mohanty et al. 2013; Roy and Chatterjee 2015).
Nath et al. (2014) have identified several tectonic features including faults and
lineaments that have the potential to generate earthquakes of Mw 3.5 and above in
this part of the Bengal Basin. Major active structures are the Jangipur–Gaibanda
Fault, the Katihar-Nilphamari Fault, and the Devikut Lineament. The Eocene Hinge
Zone is also identified as a seismically active tectonic element. This hinge is
reportedly triggered two earthquakes of magnitude Mw 7.3 and Mw 6.2 in 1842,
and 1935, respectively (Akhter 2010; Martin and Szeliga 2010; Nath et al. 2014).
Although this part of the Bengal Basin is not subjected to any direct seismic event
with high magnitude, the younger unconsolidated fluvial sediments are prone to
liquefaction under favourable ground shaking from distal seismic sources. The 1737
198 M. S. Hossain et al.
Kolkata earthquake, the 1787 Sirajganj earthquake (MM X), the 1842 Rajshahi
earthquake (Mw 7.3), the 1897 Assam earthquake (Mw 8.1), the 1934 Bihar-Nepal
earthquake (Mw 8.1), and the 1935 Pabna earthquake (Mw 6.2) had widely affected
the western part of the Bengal Basin (Bilham 2004; Akhter 2010; Martin and
Szeliga 2010; Olympa and Abhishek 2015; Nath et al. 2014). However, none of
these destructive earthquakes are reported to have caused any co-seismic surface
ruptures in this part of the Bengal Basin (Nath et al. 2014). Comparatively, this part
of the Bengal Basin is seismically less active, except in the southernmost part where
few low to moderate intensity seismic events are recorded.
The centre of the Bengal Basin comprises the deeper Foredeep Basin or
Geotectonic Province 2 excluding the northern Surma basin. This central part of the
basin is narrow at the northern end near the Madhupur-Tripura High, and then
gradually widens to *500 km to the south. A NE-SW Barisal-Chandpur High
separates the southern part of the basin into two sub-basins, the Faridpur Trough in
the west and the Hatia Trough in the east (Alam 1972; Das Gupta 1977; Guha 1978;
Reimann 1993). Topographically, this part of the basin is mainly composed of
low-lying flat land with the exception of slightly elevated Pleistocene uplands in the
northern and southeastern parts. The northern Pleistocene upland is known as the
Madhupur High/Madhupur Tract, whereas the southeastern Pleistocene upland is
known as the Lalmai Terrace (Bakr 1977; Khan et al. 2018). The Madhupur High is
forming the surface expression of a fault controlled buried anticline surrounded by
the Jamuna Fault to the west and the Old Brahmaputra Fault to the east. The High is
tilted toward the east and is faulted in the west (Morgan and McIntire 1959;
Steckler et al. 2008). The capital city Dhaka is situated at the southern edge of the
Madhupur High. GPS derived geodetic data suggest an overall very low subsidence
rate (<1 mm/year) for the Madhupur High (Reitz et al. 2015) compared to the other
part of the deeper Foredeep Basin. In and around the Dhaka City, variable but rapid
rates of subsidence are observed which probably related to usual sedimentary
compaction and extensive groundwater extraction. The coastal belt of this
Geotectonic Province 2 also shows moderate but variable rates of subsidence within
3–8 mm/year.
Except a few faults, tectonic deformation in this part of the basin is mainly
accommodated by the young and gently folded buried anticlines. Some of the
buried anticlines are Begumganj, Shahbazpur, Kachua, Daudkandi, Kamta, and
Narshingdi. All these folds are actively developing within the area bordered by the
CTFB to the east and the deformation front to the west (Fig. 9). The major active
faults in this part of the Bengal Basin are Madhupur, Dhaleswari, Padma, and
Meghna. Some other active faults and lineaments are also present in and around the
Dhaka City (Khan et al. 2011). These tectonic structures include the Bangshi and
Turag Faults to the west, Tongi Khal Fault to the north, Balu, Sitalaykha, Banar,
Synthesis of the Tectonic and Structural Elements … 199
and Arial Khan Faults to the east, and Buriganga Fault to the south. The 1812
Dhaka earthquake (MM VIII), the 1845 Sirajganj earthquake (Mw 7.1), the 1846
Mymensingh earthquake (Mw 6.2), the 1885 Bengal (Manikganj) earthquake (Mw
7), the 2001 Dhaka earthquake (Mw 4.5), the 2006 Narail earthquake (Mw 4.2), the
2008 Manikganj earthquake (Mw 3.8), the 2008 Mymensingh earthquake (Mw
5.1), and the 2008 Chandpur earthquake (Mw 4.5) caused damage, especially in
Dhaka City (Akhter 2010; Martin and Szeliga 2010; Nath et al. 2014). Among
these, the 1846 and 2008 Mymensingh earthquakes probably occurred along the
Old Brahmaputra Fault, the 1885 and 2008 Manikganj earthquakes probably
occurred along the Madhupur Fault, and the 2001 Dhaka earthquake probably
occurred along the Buriganga Fault.
The blind Madhupur Fault is constituted of a series of en-echelon faults, and
flanks the western side of the uplifted Madhupur High. It is believed that this fault
resulted either from torsion of the region or from the effect of shear along an
assumed buried fault or possibly a combination of both (Morgan and McIntire
1959). This thrust fault is east- dipping and is considered as an important structure
for the seismic hazard assessment of the central part of the Bengal Basin (CDMP II
2013). The uplift of the Madhupur High may have exerted a significant control on
the avulsion history of the Brahmaputra/Jamuna River. The cyclic avulsion history
of the Brahmaputra River with a periodicity of about 1800 years (Pickering et al.
2014) is most likely related to the uplift of the Madhupur High due to slip along the
Madhupur Fault.
In order to have a better understanding of seismicity and seismotectonics of the
Bengal Basin, especially in Bangladesh region it is necessary to focus on multi-
disciplinary and integrated studies. For an improved probabilistic and deterministic
seismic hazard mapping, it is indispensable to install and run short-period seismic
stations under a single seismological network covering the whole of Bengal Basin
and to elucidate sub-surface structures across the region.
The deformation front may exhibit distinctive oceanic subduction features such as
accretionary prism, forearc wedge, forearc basins and volcanic arc. In general,
deformation front line indicates a hypothetical approximate vertical surficial pro-
jection of the subsurface de’collement level or basal detachment where slip
diminishes to zero. In accordance, the Indo-Burmese deformation front also rep-
resents a near-vertical surficial projection of the western front of the subsurface
de’collement level where slip gradually diminishes to zero towards west. The
overall curvature of the Indo-Burmese deformation front might be related to the
differences in sediment thickness as well as the fabric of the subducting oceanic
crust of the eastern margin of the Indian Plate (Fig. 1) (Maurin and Rangin 2009a).
Although, Steckler et al. (2016) consider the deformation front (updip limit of the
large downdip extent locked detachment zone) potential for a very large earthquake,
200 M. S. Hossain et al.
it is more likely that the rupture deformation would be terminated updip more to
the east as an imbricate fault in the wedge rather than propagating up to the
Indo-Burmese deformation front. Although, Reitz et al. (2015) suggest that
the Indo-Burmese deformation front is blind and buried by active delta sedimen-
tation, however, deformation front is exposed on land as huge thickness of sedi-
ments entering the Indo-Burmese subduction zone (Steckler et al. 2008).
Wang et al. (2014) suggest the southern limit of the deformation front from
*14°N. However, the present authors would like to propose it’s southern limit at
*18°N in the offshore where the Sunda Arc trench appears to die out. From
offshore, deformation front coming ashore near the Mehgna estuary at 23°N
(Cummins 2007). In the north, the deformation front is terminated by the Dauki
Fault at the northern edge (middle) of the Surma basin (Steckler et al. 2016).
Nevertheless, some authors also suggest the northern end of the active deformation
front reaches as far as to the western end of the Dauki Fault (Steckler et al. 2008;
Maurin and Rangin 2009a). In the context of the east-west placement of the
deformation front, Acharyya (1998) placed the present subduction zone on the
western side of the CTFB, in the middle of the deeper foredeep basin (Geotectonic
Province 2). However, this is certainly the deformation front of the subduction zone
(Mukhopadhyay and Dasgupta 1988; Khan 1991a, b), but not the actual subduction
zone which lies further east beneath the CTFB-IBR. Wang et al. (2014) suggest that
the active deformation front has propagated as far west as the active GBM delta.
Deep seismic reflection survey is indicative of the blind deformation front more
west around the mouth of Padma River near Dhaka (Maurin and Rangin 2009a;
CDMP II 2013; Steckler et al. 2016). It is necessary to clarify here that the
deformation front does not represent the plate boundary, rather active faults on the
eastern margin of the CTFB represent the approximate plate boundary. The whole
length of the deformation front is characterized by the presence of Wadati-Benioff
zone extending about 300–450 km down dip to the east beneath the Burmese Plate.
High resolution SRTM images and seismic data display clear evidence of tectonic
shortening at and adjacent to the deformation front. Faults and folds along these
deformation front show right-stepping en-echelon character, i.e. dextral-reverse slip
(Wang et al. 2014). Accretion of sediment is particularly clear at the southern end in
the offshore part of the deformation front. To the east of this region, along the coast,
abundant evidence for youthful folding, and uplifted coastal terraces indicate the
tectonic activity along this front. In the Dhaka domain area (Wang et al. 2014), i.e.
in the middle part, most of the active faults and anticlines lie within 120 km east of
the deformation front. In this domain, structural and stratigraphic studies show
progressive westward growth of the fold belt, toward the deformation front
(Johnson and Alam 1991; Steckler et al. 2008; Maurin and Rangin 2009a; Wang
et al. 2014). It is worthwhile to mention here that the deformation front line both in
the offshore and onshore, marks the boundary where evidence of only youthful and
buried very gentle folds are present, but no surficial expression of the folding is
visible. This is an indication that the slip required for the structural deformation
above the de’collemen surface diminished to zero at the deformation front (Maurin
and Rangin 2009a).
Synthesis of the Tectonic and Structural Elements … 201
The Bengal Basin is located at the eastern part of the Indogangetic Plain and
extends up to the Bengal Deep Sea Fan in the Bay of Bengal. The development of
the Bengal Basin is controlled and limited by the interplay between the Eurasian,
the Indian and the Burmese Plates. The E-W oriented Himalayan Mobile Belt and
202 M. S. Hossain et al.
the N-S oriented Indo-Burman Ranges developed as the Indian Plate subducted
below the Eurasian Plate in the north and below the Burmese Plate in the east,
respectively with a syntexial bend conjugating the two belts in between. The
geometry, structural features and sediments of the Bengal Basin furnished an
archive of the past history of the area.
At the initial stage (Precambrian) only the Geotectonic Province 1 (Stable Shelf
of the Bengal Basin) was part of Indian Plate when the Indian Plate was a part of the
Gondwana Supercontinent. Subsequently, the Indian Plate occupied a central
position in the Gondwana Supercontinent until the beginning of Middle Jurassic
(*170–175 Ma) when the break-up of the Gondwana Supercontinent started. From
the Middle Jurassic onward up to the end of the Palaeocene (55.9 Ma) the Indian
Plate have drifted and isolated from its neighbouring cratonic blocks. During the
Late Paleozoic–Mid Mesozoic time the basin developed as an intra-cratonic rift
basin as evidenced by the continental Gondwana sediments deposition in the gra-
ben, half-graben structures in the Geotectonic Province 1. This episode of basin
development was followed by the Kerguelen igneous activity spreaded the SE
Indian Ocean. The Geotectonic Province 1 experienced widespread volcanism as
continental flood basalts known as the Rajmahal Trap covered the Gondwana
sediments. At that time, the peri-cratonic part on the eastern margin (the present
position of the Geotectonic Province 2) might have been influenced by marine
environment and also affected by this volcanic activity. Within the Geotectonic
Province 2 the Barisal–Chandpur Gravity High may have been formed as a rift
valley during the initial break-up of the Gondwana and formation of the Indian
Plate along which transitional zone between continent-ocean crust of the Indian
Plate has been marked by several phases of basalt flow. Subsequently, the
Geotectonic Province 2 continuously subsided and received a massive volume of
sediments from the Late Mesozoic through Tertiary to Recent times.
During the Late Cretaceous and Paleocene (between 120 and 157 Ma) the Indian
Plate migrated rapidly northwards until the oceanic part of the Indian Plate collided
with a Neotethyan intra-oceanic arc. With further migration, the continental crust of
the Indian Plate collided with the Tibetan part of the Eurasian Plate around the
Eocene-Oligocene boundary (*35 Ma) causing the northern part of Indian plate to
subduct beneath the southern Tibet (initiation of continent–continent collision).
This continental convergence resulted in folding and thrusting in the Himalayan
region.
The Bengal Basin was an integral part of the major basin that extended from the
Burmese Plate in the east to the Indian Shield in the west during the Paleocene.
During this Period, the Indian Plate moved northwards and the Tethys and the
young Indian Ocean were interconnected across the Upper Assam Shelf.
Subsequently, northern migration of the Indian plate was replaced by
north-eastward movement and resulted in the collision of north-eastern end of the
Indian Plate with the Burmese Plate uplifting the Indo–Burman Ranges (IBR) by
the Late Oligocene. The uplift of the IBR as an accretionary wedge produced a twin
gulf that separated the Burmese Tertiary Basins on the eastern and the Bengal–
Assam Basins on the western and northern flanks, respectively. These basins were
Synthesis of the Tectonic and Structural Elements … 203
fully separated with the rise of the Himalayas and IBR in the Miocene and the
Geotectonic Province 2 or the central Foredeep Basin came into existence by
*23 Ma considered as the ‘remnant ocean basin’.
The structural development of the Geotectonic Province 3 (CTFB) is related with
the development of IBR whereas the stratigraphy exposed in the region are broadly
similar to those exposed along the southern sub-Himalayan Belt. As collision and
further subduction of the Indian Plate beneath the Burmese Plate in an arc-trench
setting continued, it resulted in the development of accretionary prism and major
east dipping thrusts (Kaladan Fault to the east, and Chittagong Coastal Fault to the
west) that largely control the structural evolution of the Geotectonic Province 3.
The age and intensity of folding and elevation of the structures increase in the
easterly direction towards the IBR and the thrust component increases westward.
The CTFB comprises NNW-SSE trending low amplitude anticlinal and synclinal
folds. The western boundary of this belt is the Chittagong Coastal Fault (CCF) and
eastern boundary is Kaladan Thrust Fault. The Kaladan Thrust zone, sub-parallel to
anticlines mark the tectonic boundary between CTFB of the Bengal Basin and the
Indo-Burman Range. The CTFB comprises of Miocene-Recent sedimentary rock
successions. During the Miocene, the depositional environments changed from
marine to transitional marine. Deltaic and fluviatile conditions prevailed in the
Plio-Pleistocene. Structural (folding, faulting) development in the CTFB may have
started during Pliocene and is still going on.
A detail tectonic update of the Indian western part (Dasgupta and Mukherjee
2017), Himalaya (Mukherjee and Koyi 2010a, b; Mukherjee et al. 2012; Mukherjee
and Mulchrone 2012; Banerjee et al. 2019; Mahato et al. 2019) coupled with such
works/reviews from the eastern sector can provide a more detail tectonic evolution
of India.
Acknowledgements Thanks are due to Prof. Dr. S. M. Mahbubul Ameen for discussion and
suggestion during the writing of this manuscript. Dr. Soumyajit Mukherjee and an anonymous
reviewer are greatly appreciated for their insightful comments and thoughtful suggestions on the
initial version of this manuscript. The Springer proof-reading team and Annett Buettner are
thanked for their cooperation. Mukherjee (2019) summarizes this work.
References
Abdullah R, Yeasmin R, Ameen SMM, Khanam F, Bari Z (2015) 2D structural modelling and
hydrocarbon potentiality of the Sitakund structure, Chittagong Tripura Fold Belt, Bengal
Basin, Bangladesh. Journal Geological Society of India 85, 697–705
Abrajevitch AV, Ali JR, Aitchison JC, Badengzhu, Davis AM, Liu J, Ziabrev SV (2005)
Neotethys and the India-Eurasia collision: insights from a palaeomagnetic study of the
Dazhuqu ophiolite, southern Tibet. Earth and Planetary Science Letters 233, 87–102
Acharyya SK (1994) The Cenozoic foreland basin and tectonics of the Eastern Himalaya:
problems and perspectives. Himalayan Geology 15, 3–21
Acharyya SK (1998) Break-up of the greater Indo-Australian continent and accretion of blocks
framing South and East Asia. Journal of Geodynamics 26, 149–170
204 M. S. Hossain et al.
Acharyya SK (2003) A plate tectonic model for Proterozoic crustal evolution of central Indian
tectonic zone. Gondwana Geological Magazine 7, 9–31
Acharyya SK (2007) Collisional emplacement history of the Naga-Andaman ophiolites and the
position of the eastern Indian suture. Journal of Asian Earth Sciences 29, 229–242
Acharyya SK, Mitra ND, Nandy DR (1986) Regional geology and tectonic setting of Northeast
India and adjoining region. Memoir of the Geological Survey of India 119, 6–12
Achyuthan H, Nagasundaram M, Gourlan AT, Eastoe C, Ahmad SM, Padmakumari VM (2014)
Mid-Holocene Indian summer monsoon variability off the Andaman Islands, Bay of Bengal.
Quaternary International 349, 232–244
Acton GD (1999) Apparent polar wander of India since the Cretaceous with implications for
regional tectonics and true polar wander. In: Radhakrishna T, Piper JDA (eds) The Indian
subcontinent and Gondwana: a palaeomagnetic and rock magnetic perspective. Memoir of
Geological Society of India, vol 44, pp 129–175
Ahmed ST (1968) Cenozoic Fauna of the Cox’s Bazar coastal cliff. Unpublished M.Sc. thesis,
University of Dhaka, Dhaka, 68 p
Aier I, Luirei K, Bhakuni S, Thong GT, Kothyari GC (2011) Geomorphic evolution of
Medziphema intermontane basin and quaternary deformation in the Schuppen belt, Nagaland,
NE India. Geomorphology 55, 247–265
Aitchison JC, Ali JR, Davis AM (2007) When and where did India and Asia collide? Journal of
Geophysical Research 112, B05423
Akhter SH (2010) Earthquakes of Dhaka. In: Islam MA, Ahmed SU, Rabbani AKMG
(eds) Environment of capital Dhaka: Plants wildlife gardens park open spaces air water
earthquake. Asiatic Society of Bangladesh, pp 401–426
Akhter SH, Steckler MS, Seeber L, Agostinetti NP, Kogan MG (2010) GPS velocities and
structure across the Burma accretionary prism and Shillong Plateau in Bangladesh. AGU Fall
Meeting Supplement. Abstract T22A-08
Alam M (1972) Tectonic classification of the Bengal Basin. Geological Society of America
Bulletin 83, 519–522
Alam M (1989) Geology and depositional history of Cenozoic sediments of the Bengal Basin of
Bangladesh. Palaeogeography, Palaeoclimetology, Palaeoecology 69, 125–139
Alam MM (1991) Paleoenvironmental study of the Barail succession exposed in northeastern
Sylhet, Bangladesh. Bangladesh Journal of Scientific Research 9, 25–32
Alam M (1997) Bangladesh. In: Moores EM, Fairbridge RW (eds) Encyclopedia of European and
Asian regional geology. Chapman & Hall, London, pp 64–72
Alam AKMK, Islam MB (2017) Recent changes in Jadukata fan (Bangladesh) in response to
Holocene tectonics. Quaternary International 462, 226–235
Alam M, Alam MM, Curray JR, Chowdhury MLR, Gani MR (2003) An overview of the
sedimentary geology of the Bengal Basin in relation to the regional tectonic framework and
basin-fill history. Sedimentary Geology 155, 179–208
Ali JR, Aitchison JC (2004) Problem of positioning Palaeogene Eurasia: a review, efforts to
resolve the issue, implications for the India–Asia collision. In: Clift PD, Wang P, Khunt W,
Hayes DE (eds) Continent–ocean interactions within the East Asia Marginal Seas. AGU
Monograph 149, 23–35
Ali JR, Aitchison JC (2005) Greater India. Earth-Science Reviews 72, 169–188
Ali JR, Aitchison JC (2006) Positioning Paleogene Eurasia problem: solution for 60–50 Ma and
broader tectonic implications. Earth Planetary Science and Letters 251, 148–155
Ali JR, Aitchison JC (2008) Gondwana to Asia: plate tectonics, paleogeography and the biological
connectivity of the Indian sub-continent from the middle jurassic through latest Eocene (166–
35 Ma). Earth Science Reviews 88, 145–166
Ali SMM, Raghava MSV (1985) The Bouguer gravity map of Bangladesh and its tectono-geologic
implications. Bangladesh Journal of Geology 4, 43–56
Synthesis of the Tectonic and Structural Elements … 205
Allégre CJ, Courtillot V, Tapponnier P, Hirn A, Mattauer M, Coulon C, Jaeger JJ, Achache J,
Schärer U, Marcoux J, Burg JP, Girardeau J, Armijo R, Garie´ py C, Göpel C, Li TD, Xiao XC,
Chang CF, Li GQ, Lin BY, Teng JW, Wang NW, Chen GM, Han TL, Wang XB, Den WM,
Sheng HB, Cao YG, Zhou J, Qiu HR, Bao PS, Wang SC, Wang BX, Zhou YX, and Xu RH
(1984) Structure and evolution of the Himalayan-Tibet orogenic belt. Nature, 307, 17–22
Ameen SMM, Khan MSH, Akon E, Kazi AI (1998) Petrography and major oxide chemistry of
some Precambrian crystalline rocks from Maddhapara, Dinajpur. Bangladesh Geoscience
Journal 4, 1–19
Ameen SMM, Wilde SA, Kabir MZ, Akon E, Chowdhury KR, Khan MSH (2007)
Paleoproterozoic granitoids in the basement of Bangladesh: a piece of the Indian Shield or
an exotic fragment of the Gondwana jigsaw? Gondwana Research 12, 280–387
Ameen SMM, Hossain MS, Hossain MS, Abdullah R, Bari Z, Uddin MN, Das SC, Tapu AT,
Jahan H, Shahriar MS (2016) Deciphering the Precambrian crust between Shillong Massif and
northeast India. In: International Association of Gondwana Research (IAGR) 2016
Convention, Trivandum, India. Abstract volume, pp 75–79
Angelier J, Baruah S (2009) Seismotectonics in Northeast India: a stress analysis of focal
mechanism solutions of earthquakes and its kinematic implications. Geophysical Journal
International 178, 303–326
Ansary MA, Al-Hussaini TM, Sharfuddin M (2000) Damage assessment of July 22, 1999
Moheshkhali earthquake, Bangladesh. Paper presented at 8th ASCE specialty conference on
probabilistic mechanics and structural reliability. Indiana, USA, 6 p
Bage AK, Maurya VP, Shalivahan VP, Singh S, Tripathi A (2014) Preliminary magnetotellurics
results over Rajmahal traps. Extended Abstract, 22nd EM induction workshop, Weimar,
Germany, pp 1–4
Bakhtine MI (1966) Major tectonic features of Pakistan: part II. The Eastern Province. Science &
Industry 4, 89–100
Bakr MA (1977) Quaternary geomorphic evolution of the Brahmanbaria–Noakhali area, Comilla
and Noakhali districts, vol 1. Records of the Geological Survey of Bangladesh, Bangladesh, p 44
Baksi AK (1995) Petrogenesis and timing of volcanism in the Rajmahal flood basalt province,
northeastern India. Chemical Geology 121, 73–90
Baksi AK, Barman T, Paul D, Farrar E (1987) Widespread Early Cretaceous flood basalt
volcanism in eastern India: geochemical data from the Rajmahal–Bengal–Sylhet Traps.
Chemical Geology 63, 133–141
Ball V (1877) Geology of the Rajmahal Hills. Memoir of the Geological Survey of India 13,
155–248
Banerjee B, Sengupta BJ, Banerjee PK (1995) Signals of Barremian (116 Ma) or younger oceanic
crust beneath the Bay of Bengal along 14°N latitude between 81°E and 93°E. Marine Geology
128, 17–23
Banerjee P, Burgmann R, Nagarajan B, Apel E (2008) Intraplate deformation of the Indian
subcontinent. Geophysical Research Letters 35, L18301
Banerjee S, Bose N, Mukherjee S (2019) Field structural geological studies around Kurseong,
Darjeeling-Sikkim Himalaya, India. In: Mukherjee S (ed) Tectonics and structural geology: Indian
context. Springer International Publishing AG, Cham, pp 425–440. ISBN: 978-3-319-99340-9
Banerji RK (1979) Disang Shale, its stratigraphy, sedimentation history and basin configuration in
northeastern India and Burma. Eighth Colloquium Indian Micropalaeontology and
Stratigraphy. Baroda, India, 12 p
Banerji RK (1981) Cretaceous-Eocene sedimentation, tectonism and biofacies in the Bengal Basin,
India. Palaeogeography, Palaeoclimetology, Palaeoecology 34, 57–85
Barman P, Jade S, Shrungeshwara TS, Kumar A, Bhattacharyya S, Ray JD, Jagannathan S,
Jamir WM (2016) Crustal deformation rates in Assam Valley, Shillong Plateau, Eastern
Himalaya, and Indo-Burmese region from 11 years (2002–2013) of GPS measurements.
International Journal of Earth Sciences 106, 2025–2038
Baruah S, Hazarika D (2008) A GIS based tectonic map of north eastern India. Current Science 95,
176–177
206 M. S. Hossain et al.
Baruah S, Saikia S, Baruah S, Bora PK, Tatevossian R, Kayal JR (2016) The september 2011
Sikkim Himalaya earthquake Mw 6.9: is it a plane of detachment earthquake? Geomatics,
Natural Hazards and Risk 7, 248–263
Bastia R, Radhakrishna M, Das S, Kale AS, Catuneanu C (2010) Delineation of the 85°E ridge and
its structure in the Mahanadi Offshore Basin, Eastern Continental Margin of India (ECMI),
from seismic reflection imaging. Marine and Petroleum Geology 27, 1841–1848
Baxter AT, Aitchison JC, Ali JR, Chan JS-L, Chan GHN (2016) Detrital chrome spinel evidence
for a Neotethyan intra-oceanic island arc collision with India in the Paleocene. Journal of Asian
Earth Sciences 128, 90–104
Bender F (1983) Geology of Burma. Borntraeger, Berlin, 260 p
Berthet T, Ritz J-F, Ferry M, Pelgay P, Cattin R, Drukpa D, Braucher R, Hetényi G (2014) Active
tectonics of the eastern Himalaya: new constraints from the first tectonic geomorphology study
in southern Bhutan. Geology 42, 427–430
Bhattacharya PM, Mukhopadhyay S, Majumdar RK, Kayal JR (2008) 3-D seismic structure of the
northeast India region and its implications for local and regional tectonics. Journal of Asian
Earth Sciences 33, 25–41
Bhuiyan AH, Hossain MS (2006) Groundwater constrains and opportunity: a key for future urban
planning of Dhaka City, Bangladesh. Journal of Environmental Science (Dhaka) 4, 63–75
Bilham R (2004) Earthquakes in India and the Himalaya: tectonics, geodesy and history. Annales
Geophysicae 47, 839–858
Bilham R, England PC (2001) Plateau “pop-up” in the great 1897 Assam earth-quake. Nature 410,
806–809
Bilham R, Mencin D, Bendick R, Bürgmann R (2017) Implications for elastic energy storage in
the Himalaya from the Gorkha 2015 earthquake and other incomplete ruptures of the Main
Himalayan Thrust. Quaternary International 462, 3–21
Biswas B (1961) Geology of the Bengal Basin with special reference to stratigraphy and
micropaleontology. Dissretation, University of Calcutta, 138 p
Biswas B (1963) Results of exploration of petroleum in western part of the Bengal basin, India. In:
Proceedings of the 2nd symposium on development of petroleum resources ECAFE, mineral
resource development series no. 18, 241–250. Flood basalt province, northeast India. Chemical
Geology 121, 133–141
Biswas S, Das Gupta A (1989) Distribution of stresses in the Himalayan and the Burmese arcs.
Gerlands Beiträge zur Geophysik 98, 223–239
Biswas S, Grasemann B (2005a) Quantitative morphotectonics of the southern Shillong Plateau
(Bangladesh/India). Australian Journal of Earth Science 97, 82–93
Biswas S, Grasemann B (2005b) Structural Modelling of the Subsurface Geology of the Sylhet
Trough, Bengal Basin. Bangladesh Geoscience Journal 11, 19–33
Biswas S, Majumdar RK (1997) Seismicity and tectonics of the Bay of Bengal: evidence for
intraplate deformation of the northern Indian plate. Tectonophysics 269, 323–336
Biswas S, Coutand I, Grujic D, Hager C, Stockli D, Grasemann B (2007) Exhumation and uplift of
the Shillong Plateau and its influence on the eastern Himalayas: new constraints from apatite
and zircon (U–Th–[Sm])/He and apatite fission track analysis. Tectonics 26, TC6013
Biswas S, Majumdar RK, Das Gupta A (1992) Distribution of stress axes orientation in the
Andaman-Nicobar island region: a possible stress model and its significance for extensional
tectonics of the Andaman Sea. Physics of Earth and Planetary Interior 70, 57–63
Blaser L, Krüger F, Ohrnberger M, Scherbaum F (2010) Scaling relations of earthquake source
parameter estimates with special focus on subduction environment. Bulletin of Seismological
Society of America 100, 2914–2926
BOGMC (1997) Petroleum exploration opportunities in Bangladesh. Bangladesh Oil, Gas and
Mineral Corporation (Petrobangla), Dhaka, March, 1997
Bosch D, Garrido CJ, Bruguier O, Dhuime B, Bodinier J-L, Padròn-Navarta JA, Galland B (2011)
Building an island-arc crustal section: time constraints from a LA-ICP-MS zircon study. Earth
and Planetary Science Letters 309, 268–279
Synthesis of the Tectonic and Structural Elements … 207
Bose N, Mukherjee S (submitted-a) Field documentation and genesis of the back-structures from a
part of the Garhwal Lesser Himalaya, Uttarakhand, India: Tectonic implications. In: Sharma,
Villa IM, Kumar S (eds) Crustal architecture and evolution of the Himalaya-Karakoram-Tibet
Orogen. Geological Society of London Special Publications
Bose N, Mukherjee S (submitted-b) Field documentation and genesis of back-structures from the
foreland part of a collisional orogen: Examples from the Darjeeling-Sikkim Lesser Himalaya,
Sikkim, India. Gondwana Research
Bracciali L, Najman Y, Parrish RR, Akhter SH, Millar I (2015) The Brahmaputra tale of tectonics
and erosion: early Miocene river capture in the Eastern Himalaya. Earth and Planetary Science
Letters 415, 25–37
Brammer H (2012) The physical geography of Bangladesh. University Press Ltd., Dhaka, p 547
Brammer H (2014) Bangladesh’s dynamic coastal regions and sea-level rise. Climate Risk
Management 1, 51–62
Bulbul MAU (2015) Neotectonics of the Surma Basin, Bangladesh from GPS analysis. AGU Fall
Meeting Supplement. Abstract T41B-2884
Burgess PW, Yin A, Dubey CS, Shen Z-K, Kelty TK (2012) Holocene shortening across the main
frontal thrust zone in the eastern Himalaya. Earth and Planetary Science Letters 357–358,
152–167
CDMP II (2013) Report of active fault mapping in Bangladesh: paleo-seismological study of the
Dauki Fault and the Indian-Burman plate boundary fault. Comprehensive Disaster
Management Programme (CDMP II), Ministry of Disaster Management and Relief,
Bangladesh, 67 p
Chatterjee N (2017) Constraints from monazite and xenotime growth modelling in the
MnCKFMASH-PYCe system on the P–T path of a metapelite from Shillong-Meghalaya
Plateau: implications for the Indian shield assembly. Journal of Metamorphic Geology 35,
393–412
Chauhan OS, Vogelsang E (2006) Climate induced changes in the circulation and dispersal
patterns of the fluvial sources during late quaternary in the middle Bengal Fan. Journal Earth
System and Science 115, 379–386
Chopra S, Sharma J, Sutar A, Bansal BK (2013) Estimation of source parameters of Mw 6.9
Sikkim Earthquake and modeling of ground motions to determine causative fault. Pure and
Applied Geophysics 171, 1311–1328
Chowdhury KR (1970) Strcture and petrology of the Tertiary sedimentary sequence of the
northern part of the Changotaung Hill Range, Chittagong Hill Tracts. M.Sc. thesis, University
of Dhaka, 192 p
Chowdhury KR, Biswas S, Ahmed AMM (1996) The structural and tectonic set-up of Jaintiapur
and adjacent areas, Sylhet District, Bangladesh. Bangladesh Geoscience Journal 2, 1–14
Chowdhury KR, Huq NE, Ibna-Hamid ML, Ameen SMM (2003) Evidence of Mud intrusion into
the deposits of the Surma Group, Bengal Basin, Bangladesh. Bangladesh Geoscience Journal 9,
173–178
Chowdhury SQ (1982) Palynostratigraphy of the Neogene sediments of the Sitapahar anticline
(western flank), Chittagong Hill Tracts, Bangladesh. Bangladesh Journal of Geology 1, 35–49
Coffin MF, Pringle MS, Duncan RA, Gladczenko TP, Storey M, Müller RD, Gahagan LA (2002)
Kerguelen hotspot magma output since 130 Ma. Journal of Petrology 43, 1121–1137
Cummins PR (2007) The potential for giant tsunamigenic earthquakes in the northern Bay of
Bengal. Nature 449, 75–78
Curiale JA, Covington GH, Shamsuddin AHM, Morelos JA, Shamsuddin AKM (2002) Origin of
petroleum in Bangladesh. AAPG Bulletin 86, 625–652
Curray JR (1991) Possible green schist metamorphism at the base of a 22 km sedimentary section,
Bay of Bengal. Geology 19, 1097–1100
Curray JR (1994) Sediment volume and mass beneath the Bay of Bengal. Earth and Planetary
Science Letters 125, 371–383
Curray JR (2014) The Bengal depositional system: from rift to orogeny. Marine Geology 352,
59–69
208 M. S. Hossain et al.
Curray JR, Moore DG (1971) Growth of the Bengal deep-sea fan and denudation in the
Himalayas. Geological Society of America Bulletin 82, 563–572
Curray JR, Moore DG (1974) Sedimentary and Tectonic processes in the Bengal deep-sea fan and
geosyncline. In: Burk CA, Drake CL (eds) The geology of continental margins. Springer,
Berlin, Heidelberg
Curray JR, Munasinghe T (1991) Origin of the Rajmahal Traps and the 85°E ridge: preliminary
reconstructions of the trace of the Crozet hotspot. Geology 19, 1237–1240
Curray JR, Emmel FJ, Moore DG (2003) The Bengal Fan: morphology, geometry, stratigraphy,
history and processes. Marine and Petroleum Geology 19, 1191–1223
Das Gupta AB (1977) Geology of Assam-Arakan region. Oil Commentary, India 15, 4–34
Das JD, Saraf AK, Shujat Yazdana (2010) A remote sensing technique for identifying geometry
and geomorphological features of the Indo-Burman frontal fold belt. International Journal of
Remote Sensing 31, 4481–4503
Das JD, Shujat Y, Saraf AK (2011a) Spatial technologies in deriving the morphotectonic
characteristics of tectonically active western Tripura region, Northeast India. Journal of Indian
Society of Remote Sensing 39, 249–258
Das JD, Shujat Y, Saraf AK, Rawat V, Sharma K (2011b) Morphotectonic features and fault
propagation folding of Bhuban Hills, NE India using satellite image and DEM. Journal of
Indian Society of Remote Sensing 39, 73–81
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). Journal of Geology 125, 561–591
Dasgupta T, Mukherjee S (submitted) Sediment compaction and applications in petroleum
geoscience. In: Swenner R (ed) Springer series: advances in oil and gas exploration &
production. ISSN: 2509-372X
DeCelles PG (2012) Foreland basin systems revisited: variations in response to tectonic settings.
In: Busby C, Pérez A (eds) Tectonics of sedimentary basins: recent advances. Wiley,
Chichester, UK, pp 405–426
DeCelles PG, Kapp P, Gehrels GE, Ding L (2014) Paleocene-Eocene foreland basin evolution in
the Himalaya of southern Tibet and Nepal: implications for the age of initial India-Asia
collision. Tectonics 33, 824–849
DeCelles PG, Robinson DM, Zandt G (2002) Implications of shortening in the Himalayan
fold-thrust belt for uplift of the Tibetan Plateau. Tectonics 21, 1062
Desa MA, Ramana MV, Ramprasad T, Anuradha M, Lall MV, Kumar BJP (2013) Geophysical
signatures over and around the northern segment of the 85°E Ridge, Mahanadi offshore,
Eastern Continental Margin of India: tectonic implications. Journal of Asian Earth Sciences 73,
460–472
Desikachar SV (1974) A review of the tectonic and geologic history of eastern India in terms of
plate tectonics theory. Geological Society of India Journal 15, 137–149
Dietz RS (1953) Possible deep-sea turbidity-current channels in the Indian Ocean. Geological
Society of America Bulletin 64, 375–378
Dina NT, Rahman MJJ, Hossain MS, Sayem ASM (2016) Provenance of the Neogene succession
in the Bandarban structure, South-East Bengal Basin, Bangladesh: insights from petrography
and petrofacies. Himalayan Geology 37, 141–152
Emmel FJ, Curray JR (1985) Bengal Fan, Indian Ocean. In: Bouma AH, Barnes NE, Normark WR
(eds) Submarine fans and related turbidite sequences. Springer, New York, pp 107–112
Evans P (1932) Tertiary succession in Assam. Transactions of the Mining and Geological Institute
of India 27, 155–260
Evans P (1964) The tectonic framework of Assam. Journal Geological Society of India 5, 80–96
Farhaduzzaman M, Abdullah WH, Islam MA (2014) Hydrocarbon source potential and
depositional environment of the Surma Group shales of Bengal basin, Bangladesh. Journal
of Geological Society of India 83, 433–446
Fergusson J (1863) On recent changes in the delta of the Ganges. Quarterly Journal of the
Geological Society 19, 321–354
Synthesis of the Tectonic and Structural Elements … 209
Hiller K, Elahi M (1984) Structural development and hydrocarbon entrapment in the Surma Basin.
In: Bangladesh (northwest Indo-Burman fold belt): Fifth Offshore Southwest Conference,
Singapore, pp 656–663
Hiller K, Elahi M (1988) Structural growth and hydrocarbon entrapment in the Surma basin,
Bangladesh. In: Wagner HC, Wagner LC, Wang FFH, Wong FL (eds) Petroleum resources of
China and related subjects, Houston, Texas. Circum-Pacific council for energy and mineral
resources earth science series, vol 10, pp 657–669
Hodges KV (2000) Tectonics of the Himalaya and southern Tibet from two perspectives. GSA
Bulletin 112, 324–350
Holtrop JF, Keizer J (1970) Some aspects of the stratigraphy and correlation of the Surma Basin
wells, East Pakistan. In: ECAFE mineral resources development series, vol 36. United Nations,
New York, pp 143–154
Hoque MA, Hoque MM, Ahmed KM (2007) Declining groundwater level and aquifer dewatering
in Dhaka metropolitan area, Bangladesh: causes and quantification. Hydrogeology Journal 15,
1523–1534
Hossain A, Steckler MS, Akhter SH (2015) Resistivity imaging of strata and faults in Bangladesh.
AGU Fall Meeting Supplement. Abstract T41B-2886
Hossain I, Tsunogae T, Rajesh HM, Chen B, Arakawa Y (2007) Palaeoproterozoic U-Pb SHRIMP
zircon age from basement rocks in Bangladesh: a possible remnant of the Columbia
supercontinent. Comptes Rendus Geosciences 339, 979–986
Hossain I, Tsunogae T, Tsutsumi Y, Takahashi K (2017) Petrology, geochemistry and
LA-ICP-MS U-Pb geochronology of Paleoproterozoic basement rocks in Bangladesh: an
evaluation of calc-alkaline magmatism and implication for Columbia supercontinent amalga-
mation. Journal of Asian Earth Sciences 157, 22–39
Hossain KM (1985) Plate tectonic theory and the development of the folds of Chittagong and
Chittagong Hill Tracts. Bangladesh Journal of Geology 4, 57–67
Hossain KM, Akhter SH (1983) Structural behaviour of the Sitakund Hill range. Bangladesh
Journal of Geology 2, 17–27
Hossain MS, Chowdhury KR, Khan MSH, Abdullah R (2016) Geotectonic settings of the Dauki
fault—a highly potential source for a significant seismic threat. In: Kruhl JH (ed) International
conference Humboldt Kolleg on living under threat of earthquake—Kathmandu, Nepal.
Abstract volume, p 25
Hossain MS, Islam MM, Islam O, Chowdhury KR, Khan MSH (2018) Deformation characteristics
of the Chittagong Tripura fold belt—an insight from the geometrical analysis of folded
structures (in preparation)
Hossain MS, Khan MSH, Chowdhury KR, Afrooz M (2014) Morpho-structural classification of
the Indo-Burman ranges and the adjacent regions. In: National conference on rock deformation
& structures (RDS-III), Assam, India. Abstract volume, p 31
Ingersoll RN, Graham SA, Dickinson WR (1995) Remnant ocean basins. In: Busby CJ,
Ingersoll RV (eds) Tectonics of sedimentary basins. Blackwell, Oxford, pp 363–391
Ismail M (1978) Stratigraphical position of Bogra limestone of the platform area of Bangladesh.
In: Proceedings of the 4th annual conference of the Bangladesh geological society, pp 19–26
Johnson SY, Alam AMN (1991) Sedimentation and tectonics of the Sylhet trough, Bangladesh.
Geological Society of America Bulletin 103, 1513–1527
Kabir ASS, Hossain D (2009) Geophysical interpretation of the Rashidpur structure Surma Basin,
Bangladesh. Journal of Geological Society of India 74, 39–48
Kabir MZ, Chowdhury KR, Akon E, Kazi AI, Ameen SMM (2001) Petrogenetic study of
Precambrian basement rocks from Maddhapara, Dinajpur, Bangladesh. Bangladesh Geoscience
Journal 7, 1–18
Kayal JR (1998) Seismicity of northeast India and surroundings—development over the past 100
years. Journal of Geophysics 19, 9–34
Kayal JR (2008) Microearthquake seismology and seismotectonics of South Asia. Springer,
Dordrecht, p 449
Synthesis of the Tectonic and Structural Elements … 211
Kayal JR (2014) Seismotectonics of the great and large earthquakes in Himalaya. Current Science
106, 188–197
Kayal JR, Arefiev SS, Barua S, Hazarika D, Gogoi N, Kumar A, Chowdhury SN, Kalita S (2006)
Shillong plateau earthquakes in northeast India region: complex tectonic model. Current
Science 91, 109–114
Kent RW, Saunders AD, Kempton PD, Ghose NC (1997) Rajmahal Basalts, Eastern India: Mantle
sources and melt distribution at a volcanic rifted margin. In: Mahoney JJ, Coffin MF
(eds) Large igneous provinces: continental, oceanic, and planetary flood volcanism. American
Geophysical Union, Washington, DC
Kent RW, Pringle MS, Muller RD, Saunders AD, Ghose NC (2002) 40Ar/39Ar geochronology of
the Rajmahal Basalts, India, and their relationship to the Kerguelen Plateau. Journal of
Petrology 43, 1141–1153
Khan AA (1991a) Tectonics of the Bengal Basin. Journal of Himalayan Geology 2, 91–101
Khan AA, Agarwal BNP (1993) The crustal structure of western Bangladesh from gravity data.
Tectonophysics 219, 341–353
Khan AA, Chouhan RKS (1996) The crustal dynamics and the tectonic trends in the Bengal Basin.
Journal of Geodynamics 22, 267–286
Khan AA, Rahman T (1992) An analysis of gravity field and tectonic evaluation of the
northwestern part of Bangladesh. Tectonophysics 206, 35l–364
Khan AA, Sattar GS, Rahman T (1994) Tectogenesis of the Gondwana rifted basins of Bangladesh
in the so-called Garo-Rajmahal gap and their pre-drift regional tectonic correlation. In:
Proceedings 9th international Gondwana symposium. Geological Survey of India, Calcutta,
India
Khan FH (1991b) Geology of Bangladesh. The University Press Limited, Dhaka, 203 p
Khan MAM (1978) Geology of the eastern and northeastern parts of Sadar subdivision, Sylhet
district. Records of the Geological Survey of Bangladesh, II(IV), Bangladesh
Khan, MAM, Ismail M, Ahmad M (1988) Geology and hydrocarbon prospects of the Surma
Basin. In: Seventh offshore southeast Asia conference, Singapore, Bangladesh, pp 364–387
Khan MSH, Biswas S, Singh S, Pati P (2006) OSL choronology of Dihing formation and recent
upliftment rate along the Dauki Fault, NE Bangladesh. Bangladesh Geoscience Journal 12,
1–11
Khan MSH, Haque MM, Pati P, Chowdhury KR, Biswas S (2015) OSL derived uplift rate of
Dakhin Nhila anticline along the southeastern coast of the Bay of Bengal, Bangladesh.
Himalayan Geology 36, 143–152
Khan MSH, Hossain MS, Chowdhury KR (2017) Geomorphic Implications and active tectonics of
the Sitapahar Anticline – CTFB, Bangladesh. Bangladesh Geoscience Journal 23, 1–20
Khan MSH, Hossain MS, Uddin MA (2018) Geology and active tectonics of the Lalmai Hill area
—overview from Chittagong Tripura fold belt perspective. Journal of Geological Society India
(accepted)
Khan MSH, Parkash B, Kumar S (2005) Soil–landform development of a part of the fold belt
along the eastern coast of Bangladesh. Geomorphology 71, 310–327
Khan MSH, Saha RK, Akhter S (2011) Soil geomorphic evolution of Shitalakhya-Dhaleswari
interfluve. Bangladesh Geoscience Journal 17, 53–72
Khanam F, Rahman MJJ, Alam MM, Abdullah R (2017) Facies characterization of the Surma
Group (Miocene) sediments from Jalalabad gas field, Sylhet Trough, Bangladesh: study from
cores and wireline log. Journal of Geological Society of India 89, 155–164
Khandoker RA (1989) Development of major tectonic elements of the Bengal Basin: a plate
tectonic appraisal. Bangladesh Journal of Scientific Research 7, 221–232
Klootwijk CT, Gee JS, Peirce JW, Smith GM, Mcfadden PL (1992) A early India-Asia contact:
palaeomagnetic constraints from Ninetyeast Ridge, ODP Leg 121. Geology 20, 395–398
Kuehl SA, Hairu TM, Moore WS (1989) Shelf sedimentation off the Ganges–Brahmaputra river
system—evidence for sediment bypassing to the Bengal Fan. Geology 17, 1132–1135
Kumar S, Rino V, Hayasaka Y, Kimura K, Raju S, Terada K, Pathak M (2017) Contribution of
Columbia and Gondwana supercontinent assembly- and growth-related magmatism in the
212 M. S. Hossain et al.
evolution of the Meghalaya Plateau and the Mikir Hills, Northeast India: constraints from U-Pb
SHRIMP zircon geochronology and geochemistry. Lithos 277, 356–375
Kundu B, Gahalaut VK (2012) Earthquake occurrence processes in the Indo-Burmese wedge and
Sagaing fault region. Tectonophysics 524–525, 135–146
Lietz JK, Kabir J (1982) Prospects and constraints of oil exploration in Bangladesh. In:
Proceedings of the 4th offshore Southeast Asia conference, Singapore, pp 1–6
Lindsay JF, Holliday DW, Hulbert AG (1991) Sequence stratigraphy and the evolution of the
Ganges–Brahmaputra Delta complex. AAPG Bulletin 75, 1233–1254
Lohmann HH (1995) On the tectonics of Bangladesh. Swiss Association of Petroleum Geologists
and Engineers Bulletin 62, 29–48
Long S, McQuarrie N, Tobgay T, Grujic D, Hollister L (2011) Geological map of Bhutan. Journal
of Maps 7, 184–192
Mahato S, Mukherjee S, Bose N (2019) Documentation of brittle structures (back shear and
arc-parallel shear) from Sategal and Dhanaulti regions of the Garhwal Lesser Himalaya
(Uttarakhand, India). In: Mukherjee S (ed) Tectonics and structural geology: Indian context.
Springer International Publishing AG, Cham, pp 411–423. ISBN: 978-3-319-99340-9
Mandal BC, Woobidullah ASM, Guha DK (2004) Structural style analysis of the Semutang
anticline, Chittagong Hill tracts, Eastern Fold Belt of the Bengal Basin, Bangladesh. Journal of
Geological Society India 64, 211–222
Mannan A (2002) Stratigraphic evolution and geochemistry of the Neogene Surma Group, Surma
Basin, Sylhet, Bangladesh. Published doctoral dissertation. Department of Geology, University
of Oulu, p 190. ISBN: 951-42-6711-7
Martin S, Szeliga W (2010) A catalog of felt intensity data for 589 earthquakes in India,
1636–2008. Bulletin of the Seismological Society of America 100, 562–569
Mathur LP, Evans P (1964) Oil in India. 22nd session international geological congress
proceedings. New Delhi, India
Matin A, Misra S (2009) Repeated cataclasis in a reactivated fault zone—an example from Bengal
Basin Margin Fault, Jharkhand, India. Journal of Virtual Explorer 32
Matin MA, Fariduddin M, Hussain MMT, Khan MAM, Boul MA, Kononov AI (1986) New
concepts on the tectonic zonation of the Bengal Foredeep. In: 6th offshore South East Asia
conference and exhibition, 28–31 Jan 1986, pp 51–54, Singapore
Matin MA, Khan MAM, Fariduddin M, Boul MA, Hossain MMT, Kononov AI (1983) The
tectonic map of Bangladesh—past & present. Bangladesh Journal of Geology 2, 29–36
Maurin T, Rangin C (2009a) Structure and kinematics of the Indo-Burmese Wedge: recent and fast
growth of the outer wedge. Tectonics 28, TC2010
Maurin T, Rangin C (2009b) Impact of the 90°E ridge at the Indo-Burmese subduction zone
imaged from deep seismic reflection data. Marine Geology 266, 143–155
McDougall I, McElhinny MW (1970) The Rajmahal traps of India—KAr ages and palaeomag-
netism. Earth and Planetary Science Letters 9, 371–378
Metcalfe I (2013) Gondwana dispersion and Asian accretion: tectonic and palaeogeographic
evolution of eastern Tethys. Journal of Asian Earth Sciences 66, 1–33
Michael L, Krishna KS (2011) Dating of the 85°E ridge (northeastern Indian Ocean) using marine
magnetic anomalies. Current Science 100, 1314–1322
Milliman JD, Rutkowski C, Meybeck M (1995) River discharge to sea: a global river index
(GLORI). LOICZ reports and studies. LOICZ Core Project Office, Institute for Sea Research
(NIOZ), 125 p
Mirkhamidov FM, Mannan MM (1981) The nature of gravitional field and it’s relation with
geotectonics of Bangladesh. Unpublished report. Petrobangla, Dhaka
Misra AA, Mukherjee S (2015) Tectonic inheritance in continental rifts and passive margins.
Springer Briefs in Earth Sciences. ISBN 978–3-319-20576-2
Misra AA, Bhattacharya G, Mukherjee S, Bose N (2014) Near N-S paleo-extension in the western
Deccan region in India: does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences 103, 1645–1680
Synthesis of the Tectonic and Structural Elements … 213
Misra AA, Sinha N, Mukherjee S (2015) Repeat ridge jumps and microcontinent separation:
insights from NE Arabian Sea. Marine and Petroleum Geology 59, 406–428
Misra S (2006) Precambrian chronostratigraphic growth of Singhbhum-Orissa Craton, Eastern
Indian Shield: an alternative model. Journal of Geological Society of India 67, 356–378
Mitchell AHG (1993) Cretaceous-Cenozoic tectonic events in the western Myanmar (Burma)–
Assam region. Journal of Geological Society 150, 1089–1102
Mitra S, Priestley K, Bhattacharyya AK, Gaur VK (2005) Crustal structure and earthquake focal
depths beneath northeastern India and southern Tibet. Geophysical Journal International 160,
227–248
Mohanty WK, Mohapatra AK, Verma AK, Tiampo KF, Kislay K (2014) Earthquake forecasting
and its verification in northeast India. Geomatics, Natural Hazards and Risk 7, 194–214
Mohanty WK, Verma AK, Vaccari F, Panza GF (2013) Influence of epicentral distance on local
seismic response in Kolkata city, India. Journal of Earth System Science 122, 321–338
Molnar P, Tapponnier P (1975) Cenozoic tectonics of Asia: effects of a continental collision.
Science 189, 419–426
Monsur MH (1995) An introduction to the Quaternary geology of Bangladesh. City Press and
Publications, Dhaka, Bangladesh, 70 p
Morgan JP, McIntire WG (1959) Quaternary geology of the Bengal Basin, East Pakistan and India.
Bulletin of the Geological Society of America 70, 319–341
Morino M, Kamal ASMM, Akhter SH, Rahman MZ, Ali RME, Talukder A, Khan MMH,
Matsuo J, Kaneko F (2014) A paleo-seismological study of the Dauki fault at Jaflong, Sylhet,
Bangladesh: historical seismic events and an attempted rupture segmentation model. Journal of
Asian Earth Sciences 91, 218–226
Morino M, Kamal ASMM, Muslim D, Ali RME, Kamal MA, Rahman MZ, Kaneko F (2011)
Seismic event of the Dauki Fault in 16th century confirmed by Trench investigation at
Gabrakhari Village, Haluaghat, Mymensingh, Bangladesh. Journal of Asian Earth Sciences 42,
492–498
Mukherjee S (2013a) Channel flow extrusion model to constrain dynamic viscosity and Prandtl
number of the higher Himalayan Shear Zone. International Journal of Earth Sciences 102,
1811–1835
Mukherjee S (2013b) Higher Himalaya in the Bhagirathi section (NW Himalaya, India): its
structures, backthrusts and extrusion mechanism by both channel flow and critical taper
mechanisms. International Journal of Earth Sciences 102, 1851–1870
Mukherjee S (2013c) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy,
Berlin. 111 p
Mukherjee S (2014a) Review of flanking structures in meso- and micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2014b) Atlas of Shear Zone structures in meso-scale. Springer Geology, Cham.
124 p
Mukherjee S (2015) A review on out-of-sequence deformation in the Himalaya. In: Mukherjee S,
Carosi R, van der Beek P, Mukherjee BK, Robinson D (eds) Tectonics of the Himalaya.
Geological Society Special publications, London, vol 412, pp 67–109
Mukherjee S (2017) Airy’s isostatic model: a proposal for a realistic case. Arabian Journal of
Geosciences 10, 268
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Khonsari MM (2017) Brittle rotational faults and the associated shear heating.
Marine and Petroleum Geology 88, 551–554
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Zanskar section-microstructural
studies & extrusion mechanism by a combination of simple shear & channel flow. International
Journal of Earth Sciences 99, 1083–1110
214 M. S. Hossain et al.
Mukherjee S, Koyi HA (2010b) Higher Himalayan Shear Zone, Sutlej section-structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Mukherjee S, Kumar N (in press) A first-order model for temperature rise for uniform and
differential compression of sediments in basins. Int J Earth Sci https://doi.org/10.1007/s00531-
018-1634-6
Mukherjee A, Fryar AE, Thomas WA (2009) Geologic, geomorphic and hydrologic framework
and evolution of the Bengal basin, India and Bangladesh. Journal of Asian Earth Sciences 34,
227–244
Mukherjee S, Mulchrone K (2012) Estimating the viscosity and Prandtl number of the Tso Morari
Gneiss Dome, western Indian Himalaya. International Journal of Earth Sciences 101,
1929–1947
Mukherjee S, Carosi R, van der Beek PA, Mukherjee BK, Robinson DM (2015) In: Mukherjee S,
Carosi R, van der Beek P, Mukherjee BK, Robinson D (eds) Tectonics of the Himalaya: an
introduction. Geological Society Special publications, London, vol 412, pp 1–3
Mukherjee S, Koyi HA, Talbot CJ (2012) Implications of channel flow analogue models in
extrusion of the higher Himalayan Shear Zone with special reference to the out-of-sequence
thrusting. International Journal of Earth Sciences 101, 253–272
Mukherjee S, Misra AA, Calvès G, Nemčok M (2017) Tectonics of the Deccan Large Igneous
Province: an introduction. In: Mukherjee S, Misra AA, Calvès G, Nemčok M (eds) Tectonics
of the Deccan Large Igneous Province. Geological Society Special publications, London,
vol 445, pp 1–9
Mukherjee S, Mukherjee B, Thiede R (2013) Geosciences of the Himalaya-Karakoram-Tibet
orogen. International Journal of Earth Sciences 102, 1757–1758
Mukhopadhyay M, Dasgupta S (1988) Deep structure and tectonics of the Burmese arc:
constraints from earthquake and gravity data. Tectonophysics 149, 299–322
Mukhopadhyay M, Verma RK, Ashraf MH (1986) Gravity field and structures of Rajmahal hills:
examples of the Palaeo-Mesozoic continental margin in eastern India. Tectonophysics 131,
353–367
Murphy RW, Staff of BOGMC (1988) Bangladesh enters the oil era. Tulsa. Oil & Gas Journal,
76–82
Nag S, Gaur RK, Paul T (2001) Late Cretaceous-Tertiary sediments and associated faults in
southern Meghalaya Plateau of India vis-a-vis South Tibet: their interrelationships and regional
implications. Journal of Geological Society of India 57, 327–338
Najman Y (2006) The detrital record of orogenesis: a review of approaches and techniques used in
the Himalayan sedimentary basins. Earth-Science Reviews 74, 1–72
Najman Y, Allen R, Willett EAF, Carter A, Barford D, Garzanti E, Wijbrans J, Bickle M,
Vezzoli G, Ando S, Oliver G, Uddin M (2012) The record of Himalayan erosion preserved in
the sedimentary rocks of the Hatia Trough of the Bengal Basin and the Chittagong Hill tracts,
Bangladesh. Basin Research 24, 499–519
Najman Y, Bickle M, BouDagher-Fadel M, Carter A, Garzanti E, Paul M, Wijbrans J, Willett E,
Oliver G, Parrish R, Akhter SH, Allen R, Ando S, Chisty E, Reisberg L, Vezzoli G (2008) The
Paleogene record of Himalayan erosion: Bengal Basin, Bangladesh. Earth and Planetary
Science Letters 273, 1–14
Najman Y, Bracciali L, Parrish RR, Chisty E, Copley A (2016) Evolving strain partitioning in the
Eastern Himalaya: the growth of the Shillong Plateau. Earth and Planetary Science Letters 433,
1–9
Najman Y, Jenks D, Godin L, Boudagher-Fadel M, Millar I, Garzanti E, Horstwood M, Bracciali L
(2017) The Tethyan Himalayan detrital record shows that India-Asia terminal collision
occurred by 54 Ma in the Western Himalaya. Earth and Planetary Science Letters 459,
301–310
Nandy DR (1986) Tectonics, seismicity and gravity of Northeastern India and adjoining region.
Memoir Geological Society of India 119, 13–16
Synthesis of the Tectonic and Structural Elements … 215
Nandy DR (2001) Geodynamics of Northeast India and the adjoining region. ABC Publications,
Calcutta, pp 1–209
Naqvi SM (2005) Geology and evolution of the Indian plate. Capital Publishing Company, New
Delhi, p 450
Nath SK, Adhikari MD, Maiti SK, Devaraj N, Srivastava N, Mohapatra LD (2014) Earthquake
scenario in West Bengal with emphasis on seismic hazard microzonation of the city of Kolkata,
India. Natural Hazards and Earth System Sciences 14, 2549–2575
National Data Repository (2015) Directorate general of hydrocarbons. Under Ministry of
Petroleum & Natural Gas, India
Ni JF, Bevis M, Holt WE, Wallace TC, Seager WR (1989) Accretionary tectonics of Burma and
the three-dimensional geometry of the Burma subduction zone. Geology 17, 68–71
Nielsen C, Chamot-Rooke N, Rangin C, Cruise Team ANDAMAN (2004) From partial to full
strain partitioning along the Indo-Burmese hyper-oblique subduction. Marine Geology 209,
303–327
Norton IO, Sclater JG (1979) A model for the evolution of the Indian Ocean and the break-up of
Gondwanaland. Journal of Geophysical Research 84, 6803–6830
Oldham RD (1893) A Mannual of the geology of India and Burma, 2nd edn, vol 1, 483 p
Olympa B, Abhishek K (2015) A review on the tectonic setting and seismic activity of the Shillong
Plateau in the light of past studies. Disaster Advances 8, 34–45
Ovi MMH, Khan MSH, Haque MM (2014) Geomorphic signature of active tectonics from Sylhet
City and adjoining areas, Surma Basin, Bangladesh. Bangladesh Geoscience Journal 20, 19–34
Pascoe EHA (1975) A manual of geology of India and Burma, vol 2, 3rd edn. Controller of
Publications, New Delhi
Paul DD, Lian HM (1975) Offshore basins of southwest Asia—Bay of Bengal to South Sea. In:
Proceedings of the 9th world petrol congress, vol 3. Tokyo, pp 1107–1121
Pickering JL, Goodbred SL, Beam JC, Ayers JC, Covey AK, Rajapara HM, Singhvi AK (2017)
Terrace formation in the upper Bengal basin since the Middle Pleistocene: Brahmaputra fan
delta construction during multiple highstands. Basin Research 30, 550–567
Pickering JL, Goodbred SL, Reitz MD, Hartzog TR, Mondal DR, Hossain MS (2014) Late
Quaternary sedimentary record and Holocene channel avulsions of the Jamuna and Old
Brahmaputra River valleys in the upper Bengal delta plain. Geomorphology 227, 123–136
Prasad B, Pundir BS (2017) Gondwana biostratigraphy of the Purnea Basin (Eastern Bihar, India),
and its correlation with Rajmahal and Bengal Gondwana Basins. Journal Geological Society of
India 90, 405–427
Rabbani MG, Chowdhury KR, Huq MM (2000) Stratigraphic analysis by interpretation of seismic
and drill hole data of Rangpur Dinajpur area, Bangladesh. Bangladesh Geoscience Journal 6,
1–16
Rabinowitz PD, Woods S (2006) The Africa-Madagascar connection and mammalian migrations.
Journal of African Earth Sciences 40, 270–276
Racey A, Ridd MF (2015) Petroleum geology of Myanmar. Geological Society, London, Memoirs
45, 93–108
Rahman A (1987) Geology of Maddhapara area, Dinajpur district, Bangladesh. Records of the
Geological Survey Bangladesh 5, 1–61
Rahman MA, Blank HR, Kleinkopf MD, Kucks RP (1990a) Aeromagmetic anomaly map of
Bangladesh, scale 1:1000000. Geological Survey, Bangladesh, Dhaka
Rahman MA, Mannan MA, Blank HR, Kleinkopf MD, Kucks RP (1990b) Bouguer gravity
anomaly map of Bangladesh, scale 1:1000000. Geol. Surv, Bangladesh, Dhaka
Rahman MJJ, Suzuki S (2007) Geochemistry of sandstones from the Miocene Surma Group,
Bengal Basin, Bangladesh: implications for provenance, tectonic setting and weathering.
Geochemical Journal 41, 415–428
Rahman MJJ, Worden RH (2016) Diagenesis and its impact on the reservoir quality of Miocene
sandstones (Surma Group) from the Bengal Basin, Bangladesh. Marine and Petroleum
Geology 77, 898–915
216 M. S. Hossain et al.
Rahman MJJ, Xiao W, McCann T, Songjian A (2017) Provenance of the Neogene Surma Group
from the Chittagong Tripura Fold Belt, southeast Bengal Basin, Bangladesh: constraints from
whole-rock geochemistry and detrital zircon U-Pb ages. Journal of Asian Earth Sciences 148,
277–293
Rajendran CP, Rajendran K, Duarah BP, Baruah S, Earnest A (2004) Interpreting the style of
faulting and paleoseismicity associated with the 1897 Shillong, northeast India, earthquake:
implications for regional tectonism. Tectonics, 23, TC4009
Rajendran K, Parameswaran RM, Rajendran CP (2017) Seismotectonic perspectives on the
Himalayan arc and contiguous areas: inferences from past and recent earthquakes.
Earth-Science Reviews 173, 1–30
Ram J, Venkataraman B (1984) Tectonic framework and hydrocarbon prospects of Mizoram.
Petroliferous basins of India—II. Petroleum Asia Journal VII, 60–65
Raman, KS, Kumar S, Neogi BB (1986) Exploration in Bengal Basin India—an overview. In:
Proceedings of the South East Asia petroleum exploration society, vol VII, pp 180–191
Ramana MV, Nair RR, Sarma KVLNS, Ramprassad T, Krishna KS, Subramanyam V, D’Cruz M,
Subramanyam C, Paul J, Subramanyam AS, Chandra Sekhar DV (1994) Mesozoic anomalies
in the Bay of Bengal. Earth and Planetary Science Letters 121, 469–475
Rangin C, Sibuet J-C (2017) Structure of the northern Bay of Bengal offshore Bangladesh:
evidences from new multi-channel seismic data. Marine and Petroleum Geology 84, 64–75
Ray JS, Pattanayak SK, Pande K (2005) Rapid emplacement of the Kerguelen plume related
Sylhet traps, eastern India: evidence from 40Ar–39Ar geochronology. Geophysical Research
Letters 32, L10303
Reimann K-U (1993) Geology of Bangladesh. Gebrueder Borntraeger, Berlin, 160 p
Reitz MD, Pickering JL, Goodbred SL, Paola C, Steckler MS, Seeber L, Akhter SH (2015) Effects
of tectonic deformation and sea level on river path selection: theory and application to the
Ganges-Brahmaputra-Meghna River Delta. Journal of Geophysical Research Earth Surface
120, 671–689
Reitz MD, Steckler MS, Paola C, Seeber L (2012) Modeling coupled avulsion and earthquake
timescale dynamics. AGU Fall Meeting Supplement. Abstract EP41A-3506
Richards S, Lister G, Kennett B (2007) A slab in depth: three-dimensional geometry and evolution
of the Indo-Australian plate. Geochemistry, Geophysics, Geosystems 8, Q12003
Rowley DB (1996) Age of initiation of collision between India and Asia; a review of stratigraphic
data. Earth and Planetary Science Letters 145, 1–13
Roy AB (2014) Indian subcontinent, reference module in earth systems and environmental
sciences. Elsevier
Roy AB, Chatterjee A (2015) Tectonic framework and evolutionary history of the Bengal Basin in
the Indian subcontinent. Current Science 109, 271–279
Royer JY, Patriat P, Bergh HW, Scotese CR (1988) Evolution of the southwest Indian Ridge from
the Late Cretaceous (anomaly 34) to the Middle Eocene (anomaly 20). Tectonophysics 155(1),
235–260
Sager WW, Zhang J, Korenaga J, Sano T, Koppers AAP, Mahony JJ (2013) An immense shield
volcano within the Shatsky rise oceanic plateau, northwest Pacific Ocean. Nature Geoscience 6,
976–981
Salt CA, Alam MM, Hossain MM (1986) Bengal Basin: current exploration of the hinge zone area
of south-western Bangladesh. In: 6th offshore SE Asia (SEAPEX) conference, Singapore,
pp 55–67
Sar D, Maheshwari MK, Rangarajan S, Bahuguna CS (2009) Eighty five degrees east ridge & its
hydrocarbon potential. Geohorizons Society of Petroleum Geophysicists, India 15, 15–23
Schettino A, Scotese CR (2005) Apparent polar wander paths for the major continents (200 Ma to
the present day): a palaeomagnetic reference frame for global plate tectonic reconstructions.
Geophysical Journal International 163, 727–759
Sclater JG, Fisher RL (1974) Evolution of the east central Indian Ocean, with emphasis on the
tectonic setting of the Ninetyeast Ridge. GSA Bulletin 85, 683–702
Synthesis of the Tectonic and Structural Elements … 217
Searle MP, Corfield RI, Stephenson B, Mccarron J (1997) Structure of the north Indian continental
margin in the Ladakh-Zanskar Himalayas: implications for the timing and obduction of the
Spontang Ophiolite, India-Asia collision and deformation events in the Himalaya. Geological
Magazine 134, 297–316
Sengupta S (1966) Geological and geophysical studies in the western part of the Bengal Basin,
India. AAPG Bulletin 50, 1001–1017
Shamsuddin AHM (1989) Organic geochemistry of oligocene-miocene deposits of the Bengal
Foredeep, Bangladesh. Journal of Geological Society of India 33, 332–352
Shamsuddin AHM, Abdullah SKM (1997) Geological evolution of the Bengal Basin and its
implication in hydrocarbon exploration in Bangladesh. Indian Journal of Geology 69, 93–121
Shamsuddin AHM, Brown TA, Lee S, Curiale J (2001) Petroleum systems of Bangladesh. Proc
13th Southeast Asia Petroleum Exploration Conference, 4 – 6th April, Singapore
Shanmugam G (2016) Submarine fans: a critical retrospective (1950–2015). Journal of
Palaeogeography 5, 110–184
Sharma RS (2010) Cratons and fold belts of India. In: Lecture notes in earth sciences, Springer,
Berlin, Heidelberg
Sharma S, Sarma JN (2017) Application of Drainage basin morphotectonic analysis for assessment
of tectonic activities over two regional structures of the northeast India. Journal of the
Geological Society of India 89, 271–280
Sibuet J-C, Klingelhoefer F, Huang Y-P, Yeh Y-C, Rangin C, Lee C-S, Hsu S-K (2016) Thinned
continental crust intruded by volcanics beneath the northern Bay of Bengal. Marine and
Petroleum Geology 77, 471–486
Sikder AM (1998) Tectonic evolution of eastern folded belt of Bengal Basin. Ph.D. thesis. Dhaka
University, Dhaka, 175 p
Sikder AM, Alam MM (2003) 2-D modelling of the anticlinal structures and structural
development of the eastern fold belt of the Bengal Basin, Bangladesh. Sedimentary Geology
155, 209–226
Singh A, Bhushan K, Singh C, Steckler MS, Akhter SH, Seeber L, Kim W-Y, Tiwari AK,
Biswas R (2016) Crustal structure and tectonics of Bangladesh: new constraints from inversion
of receiver functions. Tectonophysics 680, 99–112
Singh A, Singh C, Kennett BLN (2015) A review of crust and upper mantle structure beneath the
Indian subcontinent. Tectonophysics 644–645, 1–21
Singh AP, Kumar N, Singh B (2004) Magmatic underplating beneath the Rajmahal Traps: gravity
signature and derived 3-D configuration. Journal of Earth System Science 113, 759–769
Smith AG, Hallam A (1970) The fit of the southern continents. Nature 225, 139–144
Socquet A, Vigny C, Chamot-Rooke N, Simons W, Rangin C, Ambrosius B (2006) India and
Sunda plates motion and deformation along their boundary in Myanmar determined by GPS.
Journal of Geophysical Research 111, B05406
Steckler MS, Akhter SH, Seeber L (2008) Collision of the Ganges-Brahmaputra Delta with the
Burma Arc. Earth and Planetary Science Letters 273, 367–378
Steckler MS, Akhter SH, Seeber L, Bilham RG, Kogan MG, Masson F, Maurin T, Mondal D,
Piana Agostinetti N, Rangin C, Saha P (2012) GPS velocities and structure across the Burma
accretionary prism and Shillong anticline in Bangladesh. AGU Fall Meeting Supplement.
Abstract T51F-2667
Steckler MS, Mondal D, Akhter SH, Seeber L, Feng L, Gale J (2016) Locked and loading
megathrust linked to active subduction beneath the Indo-Burman ranges. Nature Geoscience 9,
615–618
Storey M, Mahoney JJ, Saunders AD, Duncan RA, Kelley SO, Coffin MF (1995) Timing of
hot-spot related volcanism and the breakup of Madagascar and India. Science 267, 852–855
Sultana DN, Alam MM (2001) Facies analysis of the Neogene Surma group succession in the
subsurface of the Sylhet Trough, Bengal Basin, Bangladesh. In: 10th geological confernce.
Bangladesh Geological Society, Dhaka, Abstract, p 70
Syvitski JP, Vörösmarty CJ, Kettner AJ, Green P (2005) Impact of humans on the flux of terrestrial
sediment to the global coastal ocean. Science 308, 376–380
218 M. S. Hossain et al.
Szeliga W, Hough S, Martin S, Bilham R (2010) Intensity, magnitude, location and attenuation in
India for felt earthquakes since 1762. Bulletin of Seismological Society of America 100,
570–584
Talwani M, Desa MA, Ismaiel M, Krishna KS (2016) The Tectonic origin of the Bay of Bengal
and Bangladesh. Journal of Geophysical Research Solid Earth 121, 4836–4851
Tapu AT, Ameen SMM, Abdullah R, Zaman MN (2016) Geochemical evaluation of the diorite
basement in Barapaharpur, Rangpur, northwest Bangladesh. Bangladesh Geoscience Journal
22, 17–36
Thakur VC (2004) Active tectonics of Himalayan Frontal Thrust and Seismic Hazard to Ganga
plain. Current Science 86, 1554–1560
Torsvik TH, Tucker RD, Ashwal LD, Carter LM, Jamtveit B, Vidyadharan KT, Venkataramana P
(2000) Late cretaceous India-Madagascar fit and timing of break-up related magmatism. Terra
Nova 12, 220–224
Uddin A, Lundberg N (1998) Cenozoic history of the Himalayan-Bengal system: sand
composition in the Bengal Basin, Bangladesh. Geological Society of America Bulletin 11,
497–511
Uddin A, Lundberg N (1999) A paleo-Brahmaputra? Subsurface lithofacies analysis of Miocene
deltaic sediments in the Himalayan-Bengal system, Bangladesh. Sedimentary Geology 123,
239–254
Uddin A, Lundberg N (2004) Miocene sedimentation and subsidence during continent–continent
collision, Bengal basin, Bangladesh. Sedimentary Geology 164, 131–146
Uddin MN, Ahmed Z (1989) Palynology of the Kopili formation at GDH-31, Gaibandha District,
Bangladesh. Bangladesh Journal of Geology 8, 31–42
Vaccari F, Walling MY, Mohanty WK, Nath SK, Verma AK, Sengupta A, Panza GF (2011)
Site-specific modeling of SH and P-SV waves for microzonation study of Kolkata metropolitan
city, India. Pure and Applied Geophysics 168, 479–493
Valdiya KS (2010) The making of India: geodynamic evolution. Macmillan Publishers India.
816 p
Valdiya KS (2016) The making of India: geodynamic evolution. In: Tripathi SC (ed) Society of
earth scientists series, 2nd edn. Springer, Cham, 945 p
Vernant P, Bilham R, Szeliga W, Drupka D, Kalita S, Bhattacharyya AK, Gaur VK, Pelgay P,
Cattin R, Berthet T (2014) Clockwise rotation of the Brahmaputra Valley relative to India:
tectonic convergence in the eastern Himalaya, Naga Hills, and Shillong Plateau. Journal of
Geophysical Research Solid Earth 119, 6558–6571
Wadia DN (1953) Geology of India. Rev Ed, Macmillan, 552 pp
Wang Y, Sieh K, Tun ST, Lai K-Y, Myint T (2014) Active tectonics and earthquake potential of
the Myanmar region. Journal of Geophysical Research 119, 3767–3822
Webb AAG, Yin A, Dubey CS (2013) U-Pb zircon geochronology of major lithologic units in the
eastern Himalaya: implications for the origin and assembly of Himalayan rocks. Geological
Society of America Bulletin 125, 499–522
Wesnousky S, Kumar GS, Mohindra R, Thakur VC (1999) Uplift and convergence along the
Himalayan frontal thrust of India. Tectonics 18, 967–976
Woodside PR (1983) Petroleum geology of Bangladesh. Oil & Gas Journal 81, 149–155
Yin A, Dubey CS, Webb AAG, Kelty TK, Grove M, Gehrels GE, Burgess WP (2010) Geologic
correlation of the Himalayan orogen and Indian craton: part 1. Structural geology, U-Pb zircon
geochronology, and tectonic evolution of the Shillong Plateau and its neighboring regions in
NE India. Geological Society of America Bulletin 122, 336–359
Zaher MA, Rahman A (1980) Prospects and investigations for minerals in the northwestern part of
Bangladesh. Petroleum and Mineral Resources of Bangladesh. Seminar and Exhibition, Dhaka,
pp 9–18
Zhu BD, Kidd WSF, Rowley DB, Currie BS, Shaffique N (2005) Age of initiation of the
India-Asia collision in the east-central Himalaya. Journal of Geology 113. 265–285
Fold-Thrust Belt Architecture
and Structural Evolution
of the Northern Part of the Nallamalai
Fold Belt, Cuddapah Basin, Andhra
Pradesh, India
1 Introduction
Fold and thrust belts play an important role in the understanding of fragmentation
and amalgamation of continental fragments such as those of Columbia and Rodinia
supercontinents. Temporal changes in structural patterns of fold belts may indicate
major tectonics, which is relatable to the global tectonic episodes. The 400 km long
poly-deformed Mesoproterozoic Nallamalai Fold Belt (NFB) is an arcuate orogeny,
also called as orocline, whose eastern and western boundaries are bounded by thrust
zones. NFB is thrust to the western part of the Cuddapah basin along Nallamalai
thrust/Rudravaram line and in the east by the Nellore Schist Belt (NSB) along
Vellikonda thrust (Fig. 1). The two thrusts restrict the NFB in the eastern part of the
Cuddapah basin displaying structural features, which are incomparable to the
structures recorded from rest of the basin. The structural history of NFB in con-
jugation with their relation to various regional tectonic events has widely been
discussed earlier (Nagaraja Rao et al. 1987; Meijerink et al. 1984; Venkatakrishna
and Dotiwala 1987; Matin and Guha 1996; Mukherjee 2001; Saha 1994, 2002;
Saha et al. 2010; Tripathy and Saha 2013, 2015; Matin 2014). However, the
importance of the thrust zones in the northern part of the NFB has not yet been fully
realized because of limited data. We present new detailed field data to highlight the
changes in structural fabric and their significance in the evolution of the northern
part of the NFB.
V. Tripathy (&)
Geological Survey of India, Training Institute, Hyderabad, India
e-mail: vikashtripathy@gmail.com
Satyapal
Geological Survey of India, Western Region, Jaipur, India
S. K. Mitra V. V. Sesha Sai
Geological Survey of India, Southern Region, Hyderabad, India
Fig. 1 Regional geology and abridged stratigraphic successions of the Cuddapah basin, modified
after Nagaraja Rao et al. (1987) and Geological Survey of India (1990). AF: Atmakuru fault;
EGMB: Eastern Ghats Mobile Belt; GK: Gani-Kalva fault; KF: Kona fault; NFB: Nallamalai Fold
Belt; NSB: Nellore Schist Belt; NK: Nekarikallu; PG: Pranhita-Godavari Basin; SGT: Southern
Granulite Terrain
Attention is also given towards the nature and style of deformation, partitioned
in space and time as observed in the rock units of the Nallamalai and the Kurnool
Groups in the northern extent of the NFB, and hitherto unreported klippe and
window structures. Overall, the area displays structures which are of pre-Kurnool
and post-Kurnool affinity. The pre-Kurnool deformations are restricted within the
Nallamalai Group while the post-Kurnool deformations are either pre-thrust or post
thrust, later of which are imprinted both on the Nallamalai and the Kurnool Groups.
This paper puts together the structural and stratigraphic data recorded from the
northern part of NFB i.e., around Vamikonda thrust (Venkatakrishnan and
Dotiwalla 1987) to establish thrusting of NFB over the Palnad sub-basin that
display thin-skinned tectonics. These observations are important in the context of
constraining the temporal evolution of the NFB and in regional correlation with
adjoining basins and mobile belt. In this contribution, we discuss structural fabrics
in the northern part of the NFB and correlate them with tectonics related to possible
amalgamation of SE India and East Antarctica fragments.
Fold-Thrust Belt Architecture and Structural Evolution … 221
2 Geological Setting
(Natarajan and Rajagopalan Nair 1977; Saha and Chakraborty 2003; Chakraborti
2006; Chakraborti and Saha 2009; Saha et al. 2010). The overall arcuate pattern of
the fold belt is an outcome of all the major deformation events, which are evaluated
in the present work.
An aerial view of the area shows NE trending lineaments (Fig. 2), which has
marred the folding pattern within the NFB. The folds at such locations shows
juxtaposition of simultaneous antiforms along thrust plane, so that the downthrown
synform lies below the antiform. Here the fold belt shows changes in the orientation
of the fabric from NE-SW to E-W and juxtaposition of the two thrust fronts i.e., the
Nallamalai and the Vellikonda thrust near Nekarikallu (Fig. 3). Matin (2014)
suggests the NFB to be a syncline with eastern limb being overturned and an eroded
anticline. In such a model the synclinal closure is envisaged in the northern end of
the NFB near Nekarikallu. Matin (2014) viewed the deformations within the NFB
related to contemporaneous faulting and folding in front of Vellikonda thrust and
modeled the Nallamalai thrust as its imbricate. The deformation is understood to
have resulted due to fault propagation folding. In the present discourse we bring
forth the structural geometry and architecture of NFB to provide a plausible
structural model which explains the oroclinal nature of NFB. The details of the
structural fabric in northern part of NFB is discussed in detail (cf. Matin 2014)
where it changes from NE-SW to *E-W trending near to the thrust boundary with
Palnad sub-basin. The data presented here from northern part of NFB is in detail as
compared to Matin 2014 and gives better understanding of tectonic set up of the
area.
The northern part of the NFB falling in the Survey of India toposheet no. 56P/11
and 56P/15 south of Durgi and Nekarikallu is mapped (Fig. 3). The paper provides
Fig. 2 Google Earth image of the northern part of the Nallamalai Fold Belt showing prominent
NE to E trending lineaments. Dotted line is the boundary of the NFB with Palnad sub-basin and
granitic gneiss on east. Imagery Date: 31/12/2005
224
Fig. 3 Geological map of the northern part of the Nallamalai Fold Belt, Cuddapah basin showing boundaries of various sectors
V. Tripathy et al.
Fold-Thrust Belt Architecture and Structural Evolution … 225
an account of various structural features from the area which is divided in six
sectors towards east (Fig. 3) as described below.
The studied area can be divided into two major domains i.e., the footwall and
hanging wall of the major thrust plane (Satyapal and Tripathy 2014). The
Nallamalai Group of NFB acts as hanging wall while the Kurnool Group of Palnad
sub-basin is footwall. Of these sectors, the Ayyanapalem and Chejarla sectors are
represented by NE-SW trending Cumbum phyllite and minor Cumbum quartzite
and dolomite bands of Nallamalai Group. This sector is part of the hanging wall of
the major thrust. viz., Vamikonda thrust. The Chejarla sector is intruded by younger
granitoids (Sesha Sai 2004, 2013).
In the Chejarla sector, Cumbum phyllite, quartzite and dolomite are exposed
(Fig. 5a). Here, Chejarla granite expose with intrusive contact with the Cumbum
quartzite and phyllite. The sedimentary rock units in this sector are overturned to
recumbent (Fn1 fold) with northwest vergence (Fig. 5b–f). The Sn0 and Sn1 trans-
pose locally (Fig. 5e, f) with fold axis plunging gently in variable directions
(Fig. 5c, d, and f). These Fn1 folds developed in earlier deformation phase and
reoriented to recumbent by later deformation. The recumbent Fn1 folds are over-
printed by inclined asymmetric Fn2 folds which are the last folding event recorded
from the area (Fig. 5g). The superposition of F1 and F2 folds form hook shaped
Type 3 interference pattern (Ramsay 1967).
226
Fig. 4 a Geological map of Ayyanapalem sector. b Field photograph of Sn2 fabric cutting through the transposed Sn0–Sn1 in Cumbum Phyllite, western flank
of Kuchuchula Bodu. c Field photograph of sheared dolomite intercalated with phyllite, Kuchuchula Bodu. S- and C-shear planes prominent (see Mukherjee
2011, 2012, 2013, 2014b, 2015 for kinematics). d Field photograph of SE dipping overturned isoclinal folds in phyllite intercalated with dolomite, N of
V. Tripathy et al.
Fig. 5 a Geological map of Chejerla sector. b Field photograph of inclined folds in Cumbum
Dolomite SE of Venkatareddipuram. Inset: Field photograph of recumbent folds in Cumbum
Dolomite SE of Venkatareddipuram. c–f Stereographic projection of Sn0, Sn1, axial plane and fold
axis of Fn1 respectively. g Field photograph of hook shaped pattern formed by superposition of Fn1
and Fn2 folds giving rise to Type-3 interference pattern (Ramsay 1967). Inset: sketch of the field
photograph
Fig. 5 (continued)
This sector in the western most part of the studied area (Fig. 6a) and hosts the Narji
Limestone, Owk Shale and Paniam Quartzite in the footwall part while the hanging
wall is represented by phyllite and quartzite of Cumbum Formation of Nallamalai
Group. The multiple deformed Nallamalai Group has thrust over the Kurnool
Group along the *E-W trending south dipping Vamikonda thrust. The thrusting of
NFB over the Palnad sub-basin is marked by deformations both in footwall and
hanging wall. Evidence of thrust related deformation is manifested by its
overprinting on earlier events of deformations of NFB and deformation of generally
undeformed Kurnool Group.
In the area around Kakirala, the major outcrop trends NE-SW due to the defor-
mation of the Cumbum Formation which differentiates itself from the structural
disposition recorded from the rock units of the Kurnool Group. The Dn1
Fold-Thrust Belt Architecture and Structural Evolution … 229
Faults such as the NNE trending Chakralamallayya fault (Ramalinga Swamy 1972)
in particular southeast of Gajapuram, deformed both the Nallamalai and Kurnool
Groups. This steep NNE trending fault dragged the NE trending doubly plunging
folds and axial planar cleavage with a left-lateral sense, within the NFB and
230 V. Tripathy et al.
Fold-Thrust Belt Architecture and Structural Evolution … 231
Fig. 6 (continued)
232 V. Tripathy et al.
JFig. 6 a Geological map of Kakirala sector along with bedding plane poles of the Cumbum
Formation (Sn0) and Kurnool Group (Sk0) plotted in equal area stereo-net. b Overturned folds in
Owk Shale with axial plane Sk1, underlying the folded Cumbum Quartzite (Sn1), 500 m SW of
Kakirala. c Overturned and thrust sheet with NW vergence in Cumbum Quartzite near the thrust
zone, part of klippe, SW of Kakirala. d Overturned limb in Cumbum Quartzite as evident by the
younging direction of the cross beds, S of Kakirala. e Geological cross section along A-B, west of
Kakirala. A-B marked in Fig. 6a. f Varigated laminations and near orthogonal SE dipping slaty to
disjunctive axial planar cleavage in Cumbum phyllite
Kurnool Group, in NNE trend (Fig. 6a). As the fault is sub-vertical with beds
dipping both in eastern and western side its hanging wall and footwall is difficult to
establish. However, the stratigraphic positioning suggests that the upthrown west-
ern block has brought the Kurnool Group in juxtaposition to the Cumbum
Formation. Along this zone the Owk Shale shows vertical to sub-vertical dips on
both east and west directions along which ferruginous enrichment has taken place.
Another later fault with NE trend cutting through the NE trending folded Paniam
Quartzite east of Kakirala overprint earlier deformations. This fault is recognized on
the basis of slickensides, brecciation, steepening of beds, tension gashes etc.
The Cumbum Formation rocks show doubly plunging folds that are NE to ENE
trend with beds dipping either towards NW or SE (Fig. 7a). Outcrop pattern at
many places indicate thrust between alternate antiforms (e.g. in Kakirala Reserved
forest Fig. 7a). Such structures are well documented in the west and northwest of
Karempudi Tanda where alternate antiforms are stacked due to the formation of
break thrust along the forelimb of the folds forming an imbricate thrust system.
However, the areas south of Karempudi Tanda and around Mantralamma has open
folds with one limb dipping vertical to sub-vertical towards NNW/N while other
limb dip gently to moderately towards south and southeast. Variations in pattern
and style of folds in strike extension i.e., west of Karempudi Tanda and around
Mantralamma is possibly because of basement configuration and topography at the
Fold-Thrust Belt Architecture and Structural Evolution … 233
Fig. 7 a Geological map of Papayapalem sector showing Sn0 and Sn1 foliations. b Geological
cross section along C-D, east of Gajapuram (Fig. 7a). c Sub-horizontal Narji Limestone as
window, surrounded by deformed Cumbum phyllite, east of Gajapuram
Satellite image study (Fig. 2) revealed NE-SW trending doubly plunging folds
(Fig. 2) in the southeast of Durgi. The SE limb of the fold has a thrust contact with
southern lithounits. Extensive erosion in this part exposes of older Cumbum
phyllite as an klippe and Narji Limestone and Owk Shale as window lying below
the Nallamalai Group (Fig. 7a–c). The thrust stacks represent imbricate thrust
system while the presence of klippe and window structures are understood to be sole
thrust along which the NFB is thrust over the Palnad sub-basin (Fig. 7b).
234 V. Tripathy et al.
JFig. 8 a Geological map of the Belum sector. b–e Stereographic projection of Sn0, Sn1, axial plane
and fold axis from Cumbum Formation exposed in the Belum sector. f Field photograph of the
overturned north vergent folds in Cumbum phyllite east of Sanigandla Tanda. g Field photograph
of the overturned isoclinal folds in transverse section of Cumbum Quartzite at the contact with
granite (not seen in this photograph). Inset shows ridge and groove on XY plane. h Panaromic
view of the E hill range east of Sanigandla Tanda. Notice the variations in the vegetation density
and the lithological changes as viewed from the cross section. Lower part of figure shows sectional
view of the hill range with earlier formed horst and graben structures in the Kurnool Group
overthrust by the Cumbum Formation of NFB. (i) Stereographic projection showing relation
between Sn0 and fold axis from the thrust Cumbum Formation. Data collected from east of
Sanigandla Tanda (h). j Stereographic projection of bedding plane, Sk0 from Narji Limestone
showing gentle dip and gently plunging stretching lineation. k Sectional view of the sub-horizontal
Narji Limestone showing bedding parallel shear, SE of Sanigandla Tanda. Note the plan view in
inset showing bedding parallel stretching lineation in Narji Limestone
Fig. 8 (continued)
In the northern part of this sector the granitic basement for the Kurnool Group in
the Palnad sub-basin is exposed west of Gopalapuram. The granite is phaneritic
with porphyry of K-feldspar, quartz and schlieren of biotite and minor hornblende.
About 3.5 km south of Mahbubnagar, within the Guttikonda reserved forest
phaneritic granite is exposed which has thrust contact with the Cumbum Formation
(Fig. 8g). The quartzite is isoclinally folded and overturned with west vergent
gently dipping axial plane (Fig. 8g). Slickenside on fold limb is represented with
the development of ridge and groove (similar to the ridge and groove of Power and
Tullis 1989; Means 1987 and Steps—Sheltering trails of Doblas 1998) suggesting
faulted contact between the Cumbum quartzite and the granite (Fig. 8g inset).
This sector is the easternmost part of the study area and is important because of the
presence of the junction of the Vamikonda (a.k.a Nallamalai) and Vellikonda
thrusts (Fig. 9a). The later imprints of deformations recorded in this sector is
dependent up on the orientation of the structural fabric.
The ESE trending ridge of Cumbum Formation S of Nekarikallu is represented
by NW trending folds (Fn1, Fig. 9b–d) and south to southeast dipping axial planar
cleavage (Sn1, Fig. 9e) with SE plunging fold axis (Fig. 9f). The ridge is cut
through by a number of imbricate thrust (D2, Fig. 9b, c) whose movement direction
parallels stretching lineation and fold axis trend (Fig. 9g). Due to this thrusting, the
isoclinal recumbent folds (Fn1) is overprinted by later NW vergent crenulations
(Fn2) and mylonite related to the Dn3 deformation. Near the thrust zones the ESE
trending Sn1 (||Sn0) cleavage is folded/crenulated to generate reclined folds (Fig. 9d,
f–h). These imbricate thrusts are linked to a sole thrust as envisaged from gently
southeast dipping thrust zone (Fig. 9a, i) and cross section (Fig. 9b). The sole thrust
is evident by the presence of thrust contact between the overturned folded phyllite
over the asymmetrical folds (Fig. 9i). The formation of recumbent folds is possibly
Fold-Thrust Belt Architecture and Structural Evolution … 237
due to the drag of Fn1 folds along the thrust zone. These late thrusts overprint earlier
structural fabrics and generate Fn2 folds and thrust slices.
However, the NE trending ridge NW of Adda road show transposition of Sn0 and
Sn1 (Fig. 9j, k). The folds are west vergent overturned to recumbent (Fig. 9m, n)
and vary in style as observed from the ESE-WNW ridge. These structures are
overprinted by shear band with top-to-west thrust movement (Fig. 9o). The late
thrust related structures include development of gently east plunging stretching
lineation (Fig. 9l).
238 V. Tripathy et al.
Fig. 9 (continued)
Fold-Thrust Belt Architecture and Structural Evolution … 239
Fig. 9 (continued)
240 V. Tripathy et al.
JFig. 9 a Geological Map of the Nekarikallu sector. b Cross section along A-B, along the ridge
axis of ESE ridge, south of Nekarikallu. c Cross section along C-D, across the ridge axis of ESE
ridge, south of Nekarikallu. d–f Stereographic projection of Sn0, Sn1 and Fn1 fold axis;
g Stereographic projection showing relation between stretching lineation, Fn1 fold axis and shear
plane. h Field photograph of Cumbum phyllite along the NE-SW imbricate thrust zone south of
Nekarikallu with development of reclined folds (Fn2) due to the folding of transposed Sn0/Sn1.
i Field photograph of Sole thrust SW of Nekarikallu marked by the thrust recumbent folds over the
asymmetrical folds of Cumbum phyllite. Note variations in orientation of Fn1 folds in hanging wall
and footwall. j–k Stereographic projection of Sn0 and Sn1 showing transposition of the both the
fabrics, along the NE trending ridge, NW of Adda Road. l Stereographic projection of the
stretching lineation along the shear zone (n) showing consistent ESE plunge. m Field photograph
of recumbent folds in Cumbum phyllite close to the NE-SW trending thrust zone SW of
Nekarikallu. n Field photograph of west vergent overturned tight folds (Fn1) in heterolithic
Cumbum quartz-phyllite, along the NE trending ridge NW of Adda Road. o Field photograph of
shear zone cutting through the overturned Fn1 folds (m) with a top-to-west tectonic movement in
Cumbum dolomite, developed along the NE trending ridge NW of Adda Road
4 Discussions
The Nallamalai Group is folded and cut through by a number of thrust sheets,
which are convex towards the Palnad sub-basin resembling the direction in which
the thrust verge. The events of folding and faulting are categorized in time and
space that generated changes in the fold pattern and geometry. The fold belt
(NFB) acted as footwall with respect to the Nellore schist belt (NSB) along the
Vellikonda thrust during Dn2 stage of deformation (Mesoproterozoic, Saha 2002;
Table 1) which is unrelated to the Nallamalai/Vamikonda thrust, against which the
fold belt is acting as hanging wall (Dn3 Dk2). As the Vamikonda thrust is jux-
taposing Mesoproterozoic NFB and Neoproterozoic Kurnool Group, it is envisaged
that the thrust must at least be of Late Neoproterozoic age. The geometry of fold
thus formed and modified during different deformations indicates multiple com-
pressions in the Eastern part of the East Dharwar craton. The prime cause of this
variation can be because of the vorticity within the shear zone related to thrusting,
rheological differences and geometry of early formed folds (Carreras et al. 2005).
To understand the structural evolution, the geometry of this belt is to be considered
which is characterized by a number of W and NW vergent thrusts, prominent
detached folds and oroclinal nature of the fold belt. An orthogonal change in the
amount of curvature of the belt from N to E (Figs. 1 and 2) represent an oroclinal
bend as theoretically suggested by Johnston et al. (2013). The presence of window,
klippe and imbricate thrust system along the Vamikonda thrust are evidences for
allochonous nature of NFB. Further, the presence of Palnad Nappe (Natarajan and
Rajagopalan Nair 1977; Saha and Chakraborty 2003; Chakraborti 2006;
Chakraborti and Saha 2009; Saha et al. 2010) suggests an exotic nature of NFB
Fold-Thrust Belt Architecture and Structural Evolution … 241
with respect to rest of Cuddapah basin (Saha 1994, 2002; Saha and Chakraborty
2003; Tripathy and Saha 2010).
Variations in fold styles in different lithounits are well documented in the
northern part of NFB. The NE-SW trending open folds in Cumbum Formation,
south of Minchalapadu in Kakirala sector, define earlier Fn1 folds related to Dn1
deformation (Fig. 6a). The forelimb of these folds are cut through by later thrusts
resulting in the juxtaposition of two antiforms with synform being downthrown
below the former during the Dn2 deformation (Fig. 10a). These thrusts are con-
nected to a sole thrust along which the NFB is thrust over the Kurnool Group of
Palnad sub-basin as evidenced by the klippe SW of Kakirala (Fig. 6e). The NW
vergent Fn1 tight to isoclinal and overturned folds in Cumbum phyllite suggests
overall top-to-NW shear component. The competency contrast between quartzite
and phyllite is the principle cause for such modifications in fold style, which
deformed predominantly by simple/general shear (Mukherjee 2012, 2013) during
242 V. Tripathy et al.
Fig. 10 a Cartoon showing variations in fold style related to rheological variations as observed
from the field. Note that the phyllite being mechanically incompetent are tighly folded to
transposed fabrics, glide along a detachment plane while quartzite has development of imbricate
thrust. b (i) Model showing first phase of folding in NFB (ii) which been modified and reoriented
by later phase of folding and thrusting. (iii) The first phase of folding and later reorientation is
show in plan map. (iv) During syn-thrust phase, the fold belt developed oroclinal bending with
development of right-lateral movement in northern part of NFB
Fold-Thrust Belt Architecture and Structural Evolution … 243
overlain by competent units (Mitra 2003; Nemčock et al. 2005; Mukherjee et al.
2010). The deformations within the NFB imbricated thrust stacks and sole thrust
along which rock units such as shale, phyllite and carbonate frequently associate
with such weak horizons. Where thrust planes passes through competent lithounits
e.g., quartzite, they steepens and form flat-ramp geometry. The Fn1 folds are
modified to overturned to recumbent in phyllite rich layers (e.g., Fig. 5b, 9 m)
while in resistant quartzite layers the strain is accommodated by development of
thrust surfaces.
The Dn1 deformation is predominantly associated with axial planar slaty
cleavage in phyllite and disjunctive cleavage in quartzite showing variations in
orientations in different sectors due to later deformations such as thrusting and
related folding. The thrust has reoriented the fold and cleavage and provided them
the asymmetric geometry that defines the overall NW/N vergence interpreted in
terms of a non-coaxial Dn2 deformation. There are marked variation in the modi-
fications and overprinting of Dn2 deformation over the first generation structures.
Depending upon the position of the fold relative to the thrust zone the axial plane of
the Fn1 fold are close to the plane of simple shear (e.g. in the dolomite bands east of
Venkatareddipuram and heterolithic quartz-phyllite south west of Nekarikallu,
Fig. 5b). Vergence of overturned to inclined nature of fold is common throughout
the area with similar slip sense. The overthrusting of NFB over the Palnad
sub-basin must had happened after the development of thrust restricted within the
fold belt as thrust surfaces has also changed their trend from NE to E similar to the
variations in the fold trends. Oroclines (Carey 1955; Johnston et al. 2013) are result
of the action of later stress over existing orogen either in a transverse (progressive)
or in a longitudinal (secondary) manner. The progressive orocline involves basal
décollement with compressive stress perpendicular to the strike of the orogen which
are responsible for bending of fold and thrust belt undergoing thin-skinned tectonics
in response to ongoing crustal compression. The thin-skinned tectonic nature of
NFB is envisaged by Chakraborti and Saha (2009) from western part of the NFB.
The presence of window and klippe structures (Figs. 6e and 7b, c) from present
study area supports the thin-skinned nature of the NFB. Thus, the geometry and
architecture of the NFB presented here corroborate the progressive orocline as
(i) the belt shows thin-skinned nature and (ii) the late compressional structures are
result of compressive stress perpendicular to the strike of the orogen. However,
such late compressional structures in central and southern part of NFB where it
trends *N-S, are difficult to establish as it approximate Type 0 interference pattern
(Ramsay 1967). The bending of the NFB fabric along Vamikonda thrust front is
thus in response to the compression directed from SE to ESE. The presence of
granitic basement and Kurnool Group in the Palnad sub-basin must had acted as a
tectonic obstacle (i.e., a backstop) due to which the northern part of the
fold-and-thrust belt dragged right laterally (Fig. 10b). The footwall uplifts as rep-
resented by horst and graben (Fig. 8h) south of Karempudi Tanda represented by
hills of Narji Limestone must also had acted as obstacle due to which the Cumbum
Quartzite in this part is stacked to develop overturned to inclined folds with one
Fold-Thrust Belt Architecture and Structural Evolution … 245
limb being vertical to sub-vertical as opposed to the open doubly plunging folds
west of Karempudi Tanda of Papayapalem sector.
The Kurnool Group in the Palnad sub-basin acted as footwall with respect to the
Vamikonda thrust and bears various stages of deformation which are pre-, syn- and
post-thrust (Table 1). The term thrust here is related to the Vamikonda thrust. The
pre-thrust deformations are mostly related to extensional structures such as N-S
trending normal faults. The horst and graben structures present in the Kunool Group
are overthrust by the Cumbum Formation in Belum sector. The former structures
represent such extensional events related to pre-thrust deformations (Fig. 8h).
The development of NE trending folds and SE dipping disjunctive to slaty
cleavage in Narji Limestone, Owk Shale and Paniam Quartzite are well documented
from the Kakirala sector as discussed earlier. Similar development of cleavage,
gently ESE plunging bedding parallel stretching lineation and development of
mylonite in Narji Limestone in Belum sector with north or northwest vergence are
evidences of syn-thrust deformation (Fig. 8i–k).
The N-S trending Chakralamallayya fault cutting through the Cumbum
Formation and Kurnool Group east of Kakirala reoriented the structural fabric of
both the Groups in N-S direction. The steepening of Owk Shale and Paniam
Quartzite along the fault and reorientation of folds and thrusts to N-S direction are
evidences of post-thrust deformation. The post thrust deformation is also reported in
the southern part of Cuddapah basin where NNE-SSW strike-slip faults are
arranged in right stepover pattern (Goswami et al. 2016). The development of dome
and basin structures of Type- 1 interference pattern (Ramsay 1967) in Paniam
Quartzite along the fault zone in the study area are evidences of post-thrust
deformation possibly related to compression that acted on early formed NE trending
folds.
In EGB, the Western Charnockite Zone (WCZ) (cf. Domain 1A of Dasgupta and
Sengupta 2003; Ongole domain, Krishna Province of Dobmeier and Raith 2003)
has suffered *1.6 Ga granulite metamorphism without signs of Grenvillian over-
print (Rickers et al. 2001; Dasgupta and Sengupta 2003; Dobmeier and Raith 2003;
Bhattacharya et al. 2010) as opposed to its central and eastern parts. Such meta-
morphic imprints of Columbia affinity in Eastern Dharwar craton possibly affected
the deformations further westward in NSB and NFB (Tripathy 2010; Tripathy and
Saha 2013, 2015). The Nallamalai Group has undergone two major folding events
as evident from different parts of the fold belt (Matin and Guha 1996; Mukherjee
2001; Saha 2002; Tripathy and Saha 2010; Saha et al. 2010). It is understood to
have developed as a result of compressive stresses active during the Columbia
supercontinent (Saha 2002, 2004; Tripathy and Saha 2013).
The thrust related emplacement of Kandra igneous complex and Kanigiri
ophiolitic melange (Sesha Sai 2009; Saha 2011; Sesha Sai 2013) in southern
Nellore schist belt (NSB) suggests a Paleoproterozoic (1847 + 14 Ma, Vijaya
Kumar et al. 2010) plate convergence along the SE margin of India (Bhadra et al.
2004; Saha 2011; Saha and Patranabis-Deb 2014). The earlier phase of deformation
in metapelites of younger Udaygiri Group of NSB attended presumably before
1933 ± 11 Ma based on retrograde metamorphic xenotime dating (Das et al.
2015). The 1.8–1.9 Ga deformation and metamorphism related to the amalgamation
of these East Gondwana fragments deformed and metamorphosed the NSB and its
thrusting to NFB along Vellikonda Thrust (Tripathy and Saha 2013). Various
radiometric ages of igneous intrusions in NFB (reviewed in geological setting of
this paper) cluster around Late Paleoproterozoic to Early Mesoproterozoic (1.5–
1.6 Ga). The two major folding events in NFB are also understood to be prior to
such igneous intrusions within them (Saha and Chakraborty 2003). It is thus
plausible that the first deformation phase generating the Fn1 folds and thrusts in
NFB and later structures of NSB were related to compression of Late
Paleoproterozoic of Columbia Supercontinental affinity. However, sediments
deposited during *1.9 Ga in Papaghni sub-basin in western part of Cuddapah
basin lack regional deformation and thus suggest its variance from deformed NFB
in terms of deformational chronology, tectonic set up and thus paleogeographic
position.
Steep basement cutting faults such as the Gani-Kalva and the Kona faults in
western part of the Cuddapah basin opened the Kurnool sub-basin (Tripathy and
Saha 2013) possibly coeval to the emplacement of *1090 Ma kimberlite pipes
west of the basin margin (Kumar et al. 1993). Recent discoveries of 1.4–1.3 Ga
lamproite dyke-sill complex in Tadpatri Shale are proposed as diamond provenance
for Banganapalle conglomerate (Joy et al. 2012). The deposition of Kurnool Group
thus has happened during Early to Mid Neoproterozoic. Such crustal-scale exten-
sion during the Late Mesoproterozoic to Early Neoproterozoic is evident
from *1300 Ma Prakasam alkaline province (PAkP), which has a number of
alkaline intrusive bodies in the schist belt and considered to represent rift related
magmatism (Leelanandam 1989; Upadhyay et al. 2006; Leelanandam et al. 2006;
Vijaya Kumar and Leelanandam 2008; Dharma Rao et al. 2010).
Fold-Thrust Belt Architecture and Structural Evolution … 247
The thrusting of NFB over the Palnad sub-basin and development of folds and
cleavage in the Kurnool Group are related to compressional activities possibly
related to Pan-African orogeny (Saha 2004; Tripathy and Saha 2013; Saha and
Patranabis-Deb 2014; Tripathy and Saha 2015). In the eastern part of the belt,
Nekarikallu sector, the N trending imbricate thrust are thus related to the com-
pression of the Late Neoproterozoic, impressions of which are also recorded
through syn-kinematically grown phengitic white mica in mylonite from
Vinukonda granite during 501–474 Ma (Dobmeier et al. 2006). The first report of
window and klippe along the Vamikonda Thrust from northern part of NFB rep-
resent such compression involving the sedimentary packages in a thin-skinned set
up. In addition, an earlier extensional episode is recorded only within the Kurnool
Group of rocks as envisaged by the presence of horst and graben structures in the
Belum sector. These horst and graben affecting the Kurnool Group are overthrust
by the folded Nallamalai Group of rocks (Fig. 8h), suggesting such extensional
event to be of pre-thrust (related to Vamikonda thrust) in nature.
5 Conclusion
The present work suggests thrust stacking in eastern and northern part of NFB
where interaction of the two temporally variant thrusts viz., Vellikonda and
Nallamalai thrusts happened during Neoproterozoic-Paleozoic under the compres-
sional stress regimes active during Gondwana assemblage. The chronological
sequence of deformation is summarised in Table 1 and concluded below.
• Structures within the northern part of NFB are categorised here on their
chronological occurrences in relation to the thrusting of NFB over Palnad
sub-basin. Accordingly, pre-, syn- and post-thrust episodes are clubbed together.
• The pre- and syn-thrust deformation is responsible for the major regional
structural trends presently hosted by the NFB.
• Following major structural changes are recorded in the northern NFB due to the
syn-thrust deformation:
i. Thrusting of NFB over Palnad sub-basin.
ii. Bringing two spatially and temporally distinct thrusts i.e., the Vellikonda
thrust and Vamikonda (Nallamalai) thrust in close vicinity at the NE part of
the NFB near Nekarikallu.
iii. Due to this syn-thrust compressional episode, during Late Neoproterozoic
contemporaneous with Pan-African orogeny, the structural fabric (cleavage,
fold axis, axial plane, thrust internal to NFB etc.) within NFB shows
reorientation of its trend from N-S to NE-SW to E-W. It is understood to
have formed due to the presence of granitic basement and Kurnool Group of
rocks in Palnad sub-basin which must had acted as a tectonic barrier due to
which the NFB in its northern part is dragged with a right lateral component.
248 V. Tripathy et al.
iv. Change in the orientation of the major structural fabric on regional scale is
evident all along the strike of the NFB. This change is explained with the
model of progressive orocline which is a result of (i) thin-skinned nature of
the fold belt and (ii) the late compressive stress perpendicular to the strike
of the orogen bend produce the curvature in the fold belt.
Acknowledgements The authors are grateful to the Director General, Geological Survey of India
and Additional Director General, GSI, SR, Hyderabad for kindly according permission to publish
the work. This work is part of the GSI project for FS 2012-14 at state unit: Andhra Pradesh. The
authors are grateful to Dr. Soumyajit Mukherjee (IIT, Bombay), Editor for his kind invitation for
contribution to this book. This manuscript benefited greatly from the constructive reviews and
criticisms by editor and an anonymous reviewer. The views expressed in this paper are of authors
and not necessarily of GSI. Vide Mukherjee et al. (in press) for recent tectonic updates from the
southern portion of the Cuddapah basin, and Goswami and Upadhyay (2019) in a similar context.
Mukherjee (2019) summarizes this work.
References
Acharyya SK, Roy A (2000) Tectonothermal history of the Central Indian Tectonic Zone and
reactivation of major faults/shear zones. Journal of the Geological Society of India 55:239–256
Anand M, Gibson SA, Subbarao KV, Kelley SP, Dickin AP (2003) Early proterozoic melt
generation processes beneath the intra-cratonic Cuddapah basin, Southern India. Journal of
Petrology 44, 2139–2171
Bhadra S, Gupta S, Banerjee M (2004) Structural evolution across the Eastern Ghats Mobile Belt
Bastar craton boundary: hot over cold thrusting in an ancient collision zone. Journal of
Structural Geology 26, 233–245
Bhaskar Rao YJ, Pantulu GVC, Damodar Reddy V, Gopalan K (1995) Time ofearly sedimentation
and volcanism in the proterozoic Cuddapah basin, South India: evidence from Rb-Sr age of
Pulivendla mafic sill. Geological Society of India Memoirs 33, 329–338
Bhattacharya S (1996) Eastern Ghats granulite terrain of India: an overview. Journal of Southeast
Asian Earth Sciences 14, 165–174
Bhattacharya S, Das P, Chaudhary AK, Saw AK (2010) Mafic granulite xenoliths in the eastern
ghats granulite belt: implications for lower crustal processes in the Southeastern Indian
Peninsula. Indian Journal of Geology 80, 55–69
Bhowmik SK, Roy A (2003) Garnetiferous metabasites from the Sausar MobileBelt: petrology,
P-T path and implications for the tectonothermal evolution of the Central Indian Tectonic
Zone. Journal of Petrology 44, 387–420
Bhowmik SK, Pal T, Roy A, Pant NC (1999) Evidence for Pre-Grenvillian high-pressure granulite
metamorphism from the northern margin of the Sausarmobile belt in Central India. Geological
Society of India 53, 385–399
Carey SW (1955) The orocline concept in geotectonics. Proceedings of the Royal Society of
Tasmania 89, 255–288
Carreras J, Druguet E, Griera A (2005) Shear zone-related folds. Journal of Structural Geology 27,
1229–1251
Chakraborti S (2006) Analysis of fold-and-thrust structures in parts of Neoproterozoic Palnad
Sub-basin, Cuddapah basin, south India. Ph.D. thesis, University of Calcutta
Chakraborti S, Saha D (2009) Tectonic stresses and thin-skinned tectonics in a Proterozoic
fold-and-thrust belt read from calcite mylonites in the Cuddapah basin, south India. Indian
Journal of Geology 78, 37–54
Fold-Thrust Belt Architecture and Structural Evolution … 249
Chalapathi Rao NV, Miller JA, Gibson SA, Pyle DM, Madhavan V (1999) Precise40Ar/39Ar age
determinations of the Kotakonda kimberlite and Chelima lamproite, India: implication for the
mafic dyke swarm emplacement in the Eastern Dharwar craton. Journal of Geological Society
of India 53, 425–432
Chalapathi Rao NV, Fu-Yuan Wu, Srinivas M (2012) Mesoproterozoic emplacement and enriched
mantle derivation of the Racherla alkali syenite, Palaeo-Mesoproterozoic Cuddapah Basin,
southern India: insights from in situ Sr–Nd isotopic analysis on apatite. Geological Society,
London, Special Publication 365, 185–195
Chetty TRK (2011) Tectonics of proterozoic Cuddapah Basin, Southern India: a conceptual model.
Journal Geological Society of India 78, 446–456
Collins A, Patranabis-Deb S, Alexander E, Bertram CN, Falster GM, Gore RJ, Mackintosh J,
Dhang PC, Saha D, Payne JL, Jourdan F, Backé G, Halverson GP, Wade BP (2015) Detrital
mineral age, radiogenic isotopic stratigraphy and tectonic significance of the Cuddapah Basin,
India. Gondwana Research 28, 1294–1309
Crawford AR, Compston W (1973) The age of the Cuddapah and Kurnool systems, Southern
India. Journal of Geological Society of Australia 19, 453–464
Das S, Shukla D, Bhattacharjee S, Mitra SK (2015) Age constraints of Udayagiri Domain of
Nellore schist belt by xenotime dating around Pamuru, Prakasam district, Andhra Pradesh.
Geological Society of India 85, 289–298
Dasgupta PK, Biswas A (2006) Rhythms in Proterozoic sedimentation: an example from
Peninsular India. Satish Serial Publishing House, Delhi, p 340
Dasgupta S, Sengupta P (2003) Indo-Antarctic correlation: a perspective from the Eastern Ghats
Belt. In: Yoshida M, Windley BF, Dasgupta S (eds) Proterozoic East Gondwana:
supercontinent assembly and breakup. Geological Society London, Special Publication, 206,
131–143
Dharma Rao CV, Santosh M, Wu Yuan-Bao (2010) Mesoproterozoic ophiolitic mélange from the
SE periphery of the Indian plate: U-Pb zircon ages and tectonic implications. Gondwana
Research. https://doi.org/10.1016/j.gr.2010.06.007
Doblas M (1998) Slickenside kinematic indicators. Tectonophysics 295, 187–197
Dobmeier CJ, Raith MM (2003) Crustal architecture and evolution of the Eastern Ghats Belt and
adjacent regions of India. In: Yoshida M, Windley BF, Dasgupta S (eds) Proterozoic East
Gondwana: supercontinent assembly and breakup, 206. Geological Society, Special
Publications, London, pp 145–168
Dobmeier C, Lütke S, Hammerschmidt H, Mezger K (2006) Emplacement and deformation of the
Vinukonda granite—implications for the geological evolution of peninsular India and for
Rodinia reconstructions. Precambrian Research 146, 165–178
French JE, Heaman LM, Chacko T, Srivastava RK (2008) 1891–1883 Ma SouthernBastar–
Cuddapah mafic igneous events, India: a newly recognized large igneous province.
Precambrian Research 160, 308–322
Geological Survey of India (1990) Geological quadrangle map of degree sheet No. 57I
Goswami S, Upadhyay PK (2019) Tectonic history of the granitoids and Kadiri schist belt in the
SW of Cuddapah basin, Andhra Pradesh, India. In: Mukherjee S (ed) Tectonics and structural
geology: Indian context. Springer International Publishing AG, Cham, pp 253–278. ISBN
978-3-319-99340-9
Goswami S, Mukherjee A, Zakaulla S, Rai AK (2016) Stress states, faulting and their effects on
the Papaghni Group, Cuddapah Basin, India: a study along Giddankivaripalle-Madyalabodu
tract. Indian Journal of Geosciences 70, 17–33
Gupta JN, Pandey BK, Chabria T, Banerjee DC, Jayaram KMV (1984) Rb–Sr geochronological
studies on the granites of Vinukonda and Kanigiri, Prakasam district, Andhra Pradesh, India.
Precambrian Research 26, 105–109
Henderson B, Collins AS, Payne J, Forbes C, Saha D (2014) Geologically constraining India in
Columbia: the age, isotopic provenance and geochemistry of the protoliths of the Ongole
Domain, Southern Eastern Ghats, India. Gondwana Research 26, 888–906
250 V. Tripathy et al.
Jain SC, Yedekar DB, Nair KKK (1991) Central India shear zone: a major precambrian crustal
boundary. Journal of Geological Society of India 37, 521–531
Johnston ST, Weil AB, Gutiérrez-Alonso G (2013) Oroclines: thick and thin. Geological Society
of America Bulletin 125, 643–663
Joy S, Jelsma HA, Preston RF, Kota S (2012) Geology and diamond provenance of the Proterozoic
Banganapalle conglomerates, Kurnool Group, India. From: In: Mazumder R, Saha D
(eds) Palaeoproterozoic of India. Geological Society, London, Special Publications, 365,
pp 197–218
Kumar Vijaya, Leelanandam C (2008) Evolution of the Eastern Ghats belt, India: a plate tectonic
perspective. Journal of Geological Society of India 72, 720–749
Kumar A, Kumari VMP, Dayal AM, Murthy DSN, Gopalan K (1993) Rb–Sr ages of proterozoic
kimberlites of india: evidence for contemporaneous emplacement. Precambrian Research 79,
363–369
Leelanandam C (1989) The Prakasam Alkaline Province in Andhra Pradesh, India. Journal
Geological Society of India 34, 25–45
Leelanandam C, Burke K, Ashwal LD, Webb SJ (2006) Proterozoic mountain building in
Peninsular India: an analysis based primarily on alkaline rock distribution. Geological
Magazine 143, 195–212
Matin A (2014) Tectonics in the Cuddapah fold-thrust belt in the Indian shield, Andhra Pradesh,
India and its implication on the crustal amalgamation of India and Rayner craton of Antarctica
during Neoproterozoic orogenesis. International Journal of Earth Science 103, 7–22
Matin A, Guha J (1996) Structural geometry of the rocks of the southern part of the Nallamalai
Fold Belt, Cuddapah Basin, Andhra Pradesh. Journal Geological Society of India 47, 535–545
Means WD (1987) A newly recognized type of slickenside striation. Journal of Sturctural Geology
9, 585–590
Meijerink AMJ, Rao DP, Rupke J (1984) Stratigraphic and Structural development of the
Precambrian Cuddapah Basin, SE India. Precambrian Research 26, 57–104
Mitra S (2003) A unified kinematic model for the evolution of detachment folds. Journal of
Structural Geology 25, 1659–1673
Mukherjee MK (2001) Structural pattern and kinematic framework of deformation in the southern
Nallamalai fold-fault belt, Cuddapah district, Andhra Pradesh, Southern India. Journal Asian
Earth Sciences 19, 1–15
Mukherjee S (2011) Mineral fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth Sciences 100, 1303–1314
Mukherjee S (2012) Simple shear is not so simple! Kinematics and shear senses in Newtonian
viscous simple shear zones. Geological Magazine 149, 819–826
Mukherjee S (2013) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy.
Berlin. pp 1–111. ISBN 978-3-642-25608-0
Mukherjee S (2014a) Review of flanking structures in meso- and micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2014b) Atlas of shear zone structures in meso-scale. Springer Geology, Cham,
pp 1–124. ISBN 978-3-319-0088-6
Mukherjee S (2015) Atlas of structural geology. Elsevier, Amsterdam. ISBN 978-0-12-420152-1
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Talbot CJ, Koyi HA (2010) Viscosity estimates of salt in the Hormuz and Namakdan
salt diapirs, Persian Gulf. Geological Magazine 147, 497–507
Mukherjee S, Punekar J, Mahadani T, Mukherjee R (2015) A review on intrafolial folds and their
morphologies from the detachments of the western Indian Higher Himalaya. In: Mukherjee S,
Mulchrone KF (eds) Ductile shear zones: from micro- to macro-scales. Wiley Blackwell
pp 182–205
Fold-Thrust Belt Architecture and Structural Evolution … 251
Mukherjee S, Goswami S, Mukherjee A (in press) Structures and their tectonic implications form
the southern part of the Cuddapah basin, Andhra Pradesh, India. Iranian Journal of Science and
Technology, Transaction A: Science. http://doi.org/10.1007/s40995-018-0566-0
Mukhopadhyay D, Basak K (2009) The Eastern Ghats belt- a polycyclic granulite terrain. Journal
Geological Society of India 73, 489–518
Nagaraja Rao BK, Rajurkar ST, Ramalingaswamy G, Ravindra Babu B (1987) Stratigraphy,
structure and evolution of the Cuddapah Basin. In: Radhakrishna BP (ed) Purana basins of
Peninsular India (Middle to Late Proterozoic). Geological Society of India, Memoir, 6, pp 33–
86
Nance RD, Murphy JB, Santosh M (2014) The supercontinent cycle: a retrospective essay.
Gondwana Research 25, 4–29
Natarajan V, Rajagopalan Nair S (1977) Post Kurnool thrust and structural features in the northeast
part of the Palnad basin, Krishna district. A.P. Journal Geological Society of India 18, 111–116
Nemčock M, Schamel S, Gayer R (2005) Thrust belts structural architecture, thermal regimes and
petroleum systems. Cambridge University press, 541 p
Power WL, Tullis TE (1989) The relationship between slickenside surfaces in fine-grained quartz
and the seismic cycle. Journal of Structural Geology 11, 879–893
Rajurkar ST (1972) Geology of Vami Konda ranges and adjoining Palnad area, Cuddapah basin,
Andhra Pradesh. Unpublished report, Geological Survey of India
Ramlinga Swamy G (1972) Geology of the Karempudi, Gummanampadu and adjoining areas in
Cuddapah basin, Guntur district, Andhra Pradesh. Geological Survey of India, unpublished
report
Ramsay JG (1967) Folding and fracturing of rocks. MacGraw Hill, New York, p 568p
Ramsay JG, Casey M, Kligfield R (1983) Role of shear in development of the helvetic fold-thrust
belt of Switzerland. Geology 11, 439–442
Rickers K, Mezger K, Raith MM (2001) Evolution of the continental crust in the proterozoic
Eastern Ghats Belt, India and new constraints for Rodinia reconstruction: implications from
Sm–Nd, Rb–Sr and Pb–Pb isotopes. Precambrain Research 112, 183–210
Rogers JJW, Santosh M (2002) Configuration of Columbia, a mesoproterozoic supercontinent.
Gondwana Research 5, 5–22
Saha D (1994) Fold-fault structures of the Nallamalai range, Diguvametta-Nandi Ka Nama Pass,
Prakasam district, Andhra Pradesh, South India. Indian Journal of Geology 66, 203–213
Saha D (2002) Multi-stage deformation in the Nallamalai Fold Belt, Cuddapah Basin, South India
—implications for mesoproterozoic tectonism along Southeastern margin of India. Gondwana
Research 5, 701–719
Saha D (2004) Structural asymmetry and Plate tectonic set-up for a proterozoic fold and thrust belt:
Nallamalai fold belt and adjoining terrane, South India. Geological Survey of India, Special
Publication 84, 101–119
Saha D (2011) Dismembered ophiolites in Paleoproterozoic nappe complexes of Kandra and
Gurramkonda, South India. Journal of Asian Earth Sciences 42, 158–175
Saha D, Chakraborty S (2003) Deformation pattern in the Kurnool and Nallamalai Groups in the
Northeastern Part (Palnad Area) of the Cuddapah Basin, South India and its Implication on
Rodinia/Gondwana. Tectonics, Gondwana Research 6, 573–583
Saha D, Patranabis-Deb S (2014) Proterozoic evolution of Eastern Dharwar and Bastar cratons,
India—an overview of the intracratonic basins, craton margins and mobile belts. Journal of
Asian Earth Sciences 91, 230–251
Saha D, Tripathy V (2012) Palaeoproterozoic sedimentation in the Cuddapah Basin south India
and regional tectonics—a review. In: Mazumder R, Saha D (eds) Paleoproterozoic of India, vol
365. Geological Society of London Special Publication, pp 159–182
Saha D, Chakraborti S, Tripathy V (2010) Intracontinental thrusts and inclined transpression along
Eastern margin of the East Dharwar craton, India. Journal Geological Society of India 75, 323–
337
252 V. Tripathy et al.
Satyapal, Tripathy V (2014) Specialized thematic mapping in northern part of Nallamalai Fold
Belt, Cuddapah basin, Guntur and Prakasam districts, Andhra Pradesh. Unpublished
Geological Report, Geological Survey of India, Hyderabad
Sesha Sai VV (2004) Petrographic and petrochemical charecterisation of Proterozoic granites in
Nellore schist belt and northeastern fringes of Cuddapah basin. Records Geological Survey of
India 137, 184–188
Sesha Sai VV (2009) Sheeted dykes in Kandra ophiolite complex, Nellore schist belt, Andhra
Pradesh—vestiges of oceanic crust. Journal of Geological Society of India 74, 509–514
Sesha Sai VV (2013) Proterozoic granite magmatism along the terrane boundary tectonic zone to
the East of Cuddapah Basin, Andhra Pradesh—petrotectonic implications for Precambrian
crustal growth in Nellore Schist Belt of Eastern Dharwar Craton. Journal of Geological Society
of India 81, 167–182
Sesha Sai VV, Khanna TC, Krishna Reddy NR (2016) Red beds in the Cuddapah Basin, eastern
Dharwar craton, India: implications for the initiation of sedimentation during the proterozoic
oxygenation event. Journal of Indian Geophysical Union 20, 342–350
Sesha Sai VV, Tripathy V, Bhattacharjee S, Khanna TC (2017) Paleoproterozoic magmatism in
the Cuddapah basin, India. Journal of Indian Geophysical Union 21, 516–525
Tripathy V (2010) Brittle deformation in the Western Cuddapah Basin, South India and
implications for intra-continental tectonics. Jadavpur University, pp 264 (Ph.D. thesis)
Tripathy (2011) Brittle deformation in the western Cuddapah basin, South India and implications
for intra-continental tectonics. Unpublished Ph.D. thesis, Jadvapur University, 264 p
Tripathy V, Saha D (2010) Structure and low grade metamorphism of the east central part of the
proterozoic Nallamalai fold belt, south India—thrust stacking and discontinuous metamorphic
gradients along eastern margin of East Dharwar craton. Indian Journal of Geology 80(1–4),
173–188
Tripathy V, Saha D (2013) Plate margin paleostress variations and intracontinental deformations in
the evolution of Cuddapah basin through the Proterozoic. Precambrian Research 235, 107–130
Tripathy V, Saha D (2015) Inversion of calcite twin data, paleostress reconstruction and
multiphase weak deformation in cratonic interior—evidence from the Proterozoic Cuddapah
basin, India. Journal of Structural Geology 77, 62–81
Upadhyay D, Raith MM, Mezger K, Hammerschmi K (2006) Mesoproterozoic rift-related alkaline
magmatism at Elchuru, Prakasam Alkaline Province, SE India. Lithos 89, 447–477
Venkatakrishnan R, Dotiwala FE (1987) The Cuddapah salient: a tectonic model for the Cuddapah
basin, India, based on Landsat image interpretation. Tectonophysics 136, 237–253
Vijaya Kumar K, Ernst WG, Leelanandam C, Wooden JL, Grove MJ (2010) First Paleoproterozoic
ophiolite from Gondwana: geochronologic—geochemical documentation of ancient oceanic
crust from Kandra, SE India. Tectonophysics 487, 22–32
Zhang S, Li Z-X, Evans DAD, Wu H, Li H, Dong J (2012) Pre-Rodinia supercontinent Nuna
shaping up: a global synthesis with new paleomagnetic results from North China. Earth and
Planetary Science Letters 353–354, 145–155
Tectonic History of the Granitoids
and Kadiri Schist Belt in the SW
of Cuddapah Basin, Andhra Pradesh,
India
1 Introduction
The Kadiri greenstone terrane in the eastern Dharwar craton has not yet been
studied extensively. The petrotectonic associations in the Kadiri greenstone granite
terrain reflect its tectonic setting and field observations give ideas on evolution of
the belt. Several physical and chemical factors are to be considered to reconstruct
the geologic history of Archean granite-greenstone terranes. The geological and
geochemical features of this terrane provide clues to describe their genesis and
evolution. Since, greenstone belt development cannot be considered in isolation
from models for the origin of Archean highgrade granitic terranes because of close
allience between Earth’s thermal history and crustal evolution, all the rock types
such as granitoids, greenstone belt rocks and dykes are to be considered. Based on
geochemical studies of meta-volcanics of this schist belt, Satyanarayana et al.
(2000) conclude island arc and active continental marginal environment. The
calc-alkaline association of basalt, andesite, dacite and rhyolite (BADR) is the
distinguishing volcanic rock suite of convergent margins. The low-grade green-
schist facies metamorphism and associated structural deformations of rocks is
noteworthy. The Kadiri schist locates below the unconformity plane defined by the
overlying sediments of Cuddapah basin near the village Dorigallu (Fig. 1).
According to Jayananda et al. (2013b), the crystallization age of the felsic
magmas of Kadiri schist belt rocks is 2556 ± 13 Ma, dated by U–Pb SIMS. In fact,
the 2.7 Ga mafic dominated episode of greenstone volcanism is coeval to
emplacement of TTG suites surrounding the greenstone belt. On the other hand, the
2.6–2.54 Ga volcanism was dominated by intermediate and particularly felsic lavas
in the Dharwar craton (Jayananda et al. 2013a, b). The 2.6–2.54 Ga episode is
Fig. 1 Geological map of the northern extreme of Kadiri greenstone belt near the non-conformity
contact with Cuddapah sediments
2 Geology
The narrow linear Kadiri Greenstone Belt is located SW of Cuddapah Basin in the
Eastern Dharwar Craton (EDC) covering parts of Anantapur and Chittoor districts
located between 13° 45′ to 14° 7′N Latitude and 78° 2′ to 78° 15′E Longitude.
Survey of India Toposheet Nos. 57J/3, 57J/4, 57K/1 and 57K/2 covers this region.
Tectonic History of the Granitoids and Kadiri Schist Belt in … 255
Fig. 2 Two different granitoids. a Peninsular gneiss equivalent granitoids with feeble N-S
gneissic foliation. b Younger grey to pink closepet equivalent granitoids
Fig. 4 The solid fragments and rounded bomb indicates violent eruption of molten lava, which
cooled into solid fragments before they reach the ground and acquired aerodynamic round shapes
>64 mm are also embedded on the lava (Fig. 4). The broken angular blocks indi-
cate solid fragments and rounded clasts are bomb formed during violent eruption of
molten lava, which cool into solid fragments before they reach the ground and
acquired round shapes. Flow foliation is another characteristic feature of the felsic
volcanics of the schist belt (Fig. 5) where flow beds occur with *N strike, 5–10°
dip towards E. Autoclastic fragmental rocks (Fig. 6) are identified to interpret
in situ fragmentation of coherent magmatic bodies (Fisher and Schmincke 1984;
Mc Phie et al. 1993). These autobreccias commonly form at the base and the top of
lava flows (Fisher and Schmincke 1984; Cas and Wright 1988; McPhie et al. 1993).
Feebly developed gneissosity is observed in the surrounding gneissic rocks.
Granite-gneiss complexes comprise the dominant portion in this granite-greenstone
terrane. The foliation in the gneissic complexes is N-S similar to the greenstone belt
near greenstone-gneiss contacts and becomes exceedingly variable in intervening
Fig. 5 Flow foliation developed in acid volcanics due to high viscosity. The foliation planes show
*N-S strike, 2°–8°=>E dip
Tectonic History of the Granitoids and Kadiri Schist Belt in … 257
Fig. 6 Section of flow layers showing auto breccia with in layer suggest in situ fragmentation of
coherent magmatic bodies
regions. Mechanical rotation and elongation of clast along preferred N-S trend and
the internal deformation and crushing of grains at places by cataclasis during
rotation due to E-W compression are noteworthy. The deformation signature cor-
responds to a shallow low temperature environment of regional metamorphism.
Isoclinal folds with steeply dipping axial planes and moderate to steep plunges are
preserved. The cross-section through rhyolite lava exhibits flow banding of visually
distinct layers of differing crystallinity or vesicularity. The youngest dykes and quartz
reefs cut across both the granite and the greenstone belt. The linear andesite and
dolerite dykes often run along top of granitic hillocks for km with loose angular blocks
arranged apparently like unsorted dump. These younger intrusives are identified as
blocky lava with very rough surface composed of often loose clinkers and rubble
(Fig. 7). The Cuddapah sediments overlie the basement complex along the Eparchaean
unconformity with basal conglomerates (Goswami et al. 2015). Above this noncon-
formity the polymictic conglomerate (Fig. 8) with abundant clasts of chert, jasper and
BIF indicates supracrustal provenance with Archaean volcano-sedimentary rocks.
Fig. 7 Basic emplacement along NE-SW cut across granitoids exhibiting stack of pieces of
andesitic blocks are characterised as blocky lava
258 S. Goswami and P. K. Upadhyay
3 Structures
E, NE, NW and N are the different fracture trends. The first two trends are more
dominant (Fig. 9). Schistosity and gneissosity show a regional N and the younger
granitic intrusions too are N-S. The granitoids are intruded by basic intrusions along
NW, NE, WNW and ENE trends. Quartz reefs are also remarkable intrusive fea-
tures along NW, NE and WNW. Hematitisation is noted locally near the contact
zones of quartz reef and basement granitoids. The rocks are deformed with a
specific texture and structure that records the deformation by developing the pre-
ferred clast orientation. The fabric appears to be a type of tectonites (Mukherjee
2015). Granitic apophyses in the deformed meta-volcanic matrix of the greenstone
Fig. 9 a b diagram showing all fracture plane attitudes of the study area. b p diagram showing
poles of all fracture planes. c p diagram superimposed with corresponding rose diagram to
visualize the paleostress condition. d Superimposed pie and beta diagram of fracture shows NE
trend to be most abundant followed by E
260 S. Goswami and P. K. Upadhyay
Fig. 10 a Outcrop view of L-S tectonites. b Zoomed in view of the three dimensional section of
tectonite outcrop. The N-S stretching is compensated by E-W flattening and undisturbed
intermediate stress direction. c Plan view of tectonite outcrop. d Later deformation affected
tectonite
4 Petrology
The Kadiri schist belt metavolcanics are very fine-grained and mostly are felsic.
Minerals cannot be identified even under a hand lens. Under optical microscope
quartz, plagioclase, orthoclase and chlorite are seen as the major minerals followed
by sphene, epidote and calcite. Very fine-grained, foliated acid volcanics composed
chiefly of quartz, sericite and chlorite and other flakey clay minerals and titanate
minerals with flow foliation. Epidote, calcite, plagioclase feldspar and zoisite are
the minor/accessory mineral phases. The systematic approach of studying tectonite
fabric elements under microscope (Mukherjee and Chakraborty 2007; Mukherjee
2010a, b, Mukherjee 2011a, b, 2012, 2013, 2014a, b; Mukherjee and Koyi 2009,
2010a, b) is to study both clast and matrix components. The clasts of granitic
apophyses and volcanic matrix are shown in Fig. 11.
The Peninsular gneiss equivalent and younger closepet granite equivalent rocks
are recognised and differentiated as per these field guidelines:
A. Geomorphologically it is clear that the Peninsular gneiss equivalents forms the
peneplane mostly soil covered area where as younger granitoids of Closepet
forms ridges and mounds with circular to elongate masses of high and rugged
relief.
Fig. 11 a Very fine grained foliated preferentially aligned quartz, clayey minerals and titanate.
Minor quartz veins are visible (2X, TL, 2N). b Alignment of chlorite along foliation (10X, TL,
1N). c Phenocryst of plagioclase and sanidine in fine grained matrix (2X, TL, 2N). d Granitic
texture with quartz, plagioclase, orthoclase and minor chlorite (2X, TL, 2N)
262 S. Goswami and P. K. Upadhyay
5 Geochemistry
Fig. 12 Andesite porphyry displaying a aphanitic porphyritic texture with subhedral plagioclase
phenocrysts (white crystals) and the fine—grained aphanitic grey groundmass in outcrop scale.
b Porphyritic texture under microscope (2X, TL, 2N). c Quartz, feldspar, biotite and minor chlorite
rich groundmass (10X, TL, 1N). d Zone plagioclase phenocryst (10X, TL, 2N). e Deformed
plagioclase phenocryst (10X, TL, 2N)
264
Table 1 Major element oxide (%) data of selective samples from Kadiri granite-greenstone terrain
S. No. Rock Symbol SiO2 Al2O3 TiO2 Fe2O3 MnO MgO CaO Na2O K2O P2O5
S-1.8/6 Altered Basic 33.03 16.19 3.86 25.85 0.2 16.29 2.01 <0.01 <0.01 0.31
S-1.8/9 Diorite 48 17.12 0.62 10.17 0.11 9.78 10.24 2.15 0.29 0.09
1.8SP1 Basalt porphyry 64.34 14.06 0.88 8.22 0.1 0.63 3.19 3.57 4.01 0.4
1.8SP2 Pink Granite 73.9 13.27 0.23 1.92 0.03 0.34 0.93 3.65 5.45 0.11
1.8SP3 Diorite 57.72 14.8 0.72 7.19 0.1 5.01 5.36 4.01 2.15 0.29
1.8SP4 Basalt porphyry 65.05 15.83 0.54 4.87 0.06 0.99 3.78 5.51 1.56 0.22
1.8SP5 Basalt porphyry 54.87 13.46 0.71 8.78 0.13 7.94 5.97 4.24 1.67 0.32
1.8SP6 Granite 73.14 13.25 0.14 1.41 0.03 0.55 0.5 4.74 5.15 0.07
1.8SP7 Gray granite 73.59 13.45 0.13 1.12 0.01 0.15 0.11 4.59 5.44 0.07
1.8SP8 Diorite 56.36 15.69 0.7 8.55 0.1 5.38 6.15 4.37 1.82 0.27
1.8SP9 Basalt porphyry 62.67 13.76 0.98 7.36 0.1 3.46 4.59 3.46 2.26 0.22
1.8SP10 Granite 72.96 13.58 0.11 13.58 0.03 0.21 0.92 4.6 4.88 0.08
KS1 Schist belt 76.69 13.23 0.22 1.25 0.04 0.89 0.69 3.15 3.29 0.18
KS2 Schist belt 62.56 21.02 0.31 4.41 0.07 0.37 0.62 1.79 7.39 0.14
KS3 Schist belt 73.36 16.03 0.25 1.60 0.05 0.72 0.49 2.40 4.57 0.17
KS4 Schist belt 72.05 13.85 0.31 2.36 0.05 1.13 1.96 4.36 1.20 0.19
KS5 Schist belt 77.54 12.57 0.14 1.10 0.03 0.92 3.41 0.81 1.53 0.18
KS6 Schist belt 58.93 16.26 0.81 6.48 0.07 3.65 7.16 4.00 0.77 0.21
KS7 Schist belt 67.90 15.50 0.31 3.35 0.05 0.68 1.26 6.05 2.92 0.15
S. Goswami and P. K. Upadhyay
Tectonic History of the Granitoids and Kadiri Schist Belt in … 265
record a progressive decrease in iron and magnesium with increasing SiO2 and
alkali concentrations. Available FeO and MgO are progressively removed from the
melt through crystallization of olivine, pyroxenes, amphiboles, biotite and iron
oxide minerals such as magnetite. The removal of ferromagnesian minerals enriches
the melt in Na2O, K2O and SiO2. The SiO2 versus FeOt/MgO plot (Miyashiro
1974) and AFM diagram (after Irvine and Barager 1971) of data-points fall in the
calc-alkaline field (Figs. 15 and 16). The progressive changes in chemical com-
position are plotted on the AFM (A, alkali; F, iron; M, magnesium wt%) diagram
(Fig. 16) to study along with the Harker diagrams (Figs. 13 and 14). This is to
demonstrate that the calc-alkaline rocks generated by diversification of andesitic
parent magma. Thus, with increasing fractionation from magma, the calc—alkaline
series exhibits increase in SiO2, Na2O and K2O and decrease in CaO, MgO and the
total iron (Bowen 1928; Miyashiro 1974; Grove and Kinzler 1986). However, due
to successive metamorphism and alterations the addition or removal of certain
oxides can show some discrepancy (Table 2).
As far as tectonic setting is concerned, the plutonic and hypabyssal rocks were
separately studied and interpreted from volcanics. Felsic volcanic data suggest
convergent plate margin for greenstone belt rocks (Fig. 17a) wherein younger
batholiths and composite volcanoes develop above the subduction zone. It is likely
that the parent melts for the acid volcanics produced by partial melting of a rela-
tively undepleted mantle source. Older dioritic to granidioritic peninsular gneiss
equivalent and younger closepet equivalent granitoids along with intermediate
younger hypabyssal porphyries are plotted in the tectonic discrimination diagrams
after Maniar and Piccoli (1989). The older gneiss equivalences plot in the
post-orogenic granitoids (POG) field on the Al2O3 versus SiO2 plot (Fig. 17b).
Thus we can consider the Dharwar batholith to have formed during the crust
stabilizing post-orogenic environment. Whereas the younger granitoids and dykes
plot in the combined field (Fig. 17c), which must be clarified from other supportive
discrimination diagrams. The geochemical data are also plotted into the Batchelor
and Bowden’s diagram (1985) that support the tectonic environment for older
gniess equivalents as post-orogenic. Additionally it shows younger granitoids and
porphyries to associate with post-plate collision uplift (Fig. 17d).
6 Discussions
The greenstone belt rocks near Dorigallu indicate submarine volcanism along fissures
developed in convergent plate boundary involved the transfer of magma (together
with phenocrysts, dissolved volatiles, and country-rock material entrained from the
conduit walls or ground surface) from some depth. The breccias and agglomerates
indicate fissure vent surrounding environment (Goswami et al. 2018a, b). Fragments
in breccias are chiefly felsic volcanics and range up to 1 m across although generally
averaging 10–30 cm. According to Goodwin et al. (1972) the ratio of pyroclastics to
flows increases both with increasing stratigraphic height and with decreasing distance
266 S. Goswami and P. K. Upadhyay
Fig. 13 Variation diagram (Harker Harker 1909) showing multiple oxide (%) plots against SiO2
(%)
to eruptive centers. Thus the studied sections in Kadiri schist belt suggest middle
portion of the greenstone succession. The intensive shearing and brecciation of
granitic apophyses, intruded along schistosity plane in the northern part of Kadiri
greenstone belt, causes pseudo appearance of agglomerate. The actual agglomerate
units are not well-sorted sorted and fragments are generally angular although some
units are composed of well-rounded bombs. The medium to fine ash-flow tuff units are
welded and well bedded. Individual beds are 1 cm to few meters thick.
The basalt, andesite, dacite and rhyolite (BADR) is the signature suite of rocks
indicates calc-alkaline volcanic association. This suite can act as a fingerprint of
convergent plate margin. This setting constitutes possibly one of the most voluminous
rock assemblages on the Earth after MORB (Perfit et al. 1980; Grove and Kinzler
1986). Harker diagram plots of major elements indicate a liquid line of descent from a
common magma source. This BADR rock association are derived from a common
parent magma of basaltic to andesitic composition. These calcalkaline volcanic rocks
Tectonic History of the Granitoids and Kadiri Schist Belt in … 267
Fig. 14 Variation diagrams of selective trace elements (ppm) against SiO2 (%)
are formed at convergent margins where more silicic rocks represent enhanced
fractionation. The calc-alkaline trend with ocean-continent convergence setting of the
present data supports the observation of Chadwick et al. (2000). The calc-alkaline
suite is also reflected from enrichment in SiO2, alkalis (Na2O and K2O), LILE,
volatiles and is relatively depleted in FeO, MgO, HFS concentrations (Miyashiro
1974; Hawkesworth et al. 1993; Pearce and Peate 1995) like the Phanerozoic con-
vergent margins eg., Sierra Nevada (Fliedner et al. 2000; Ducea 2001) and ring of fire
in present day. The peraluminous to metaluminous nature (Fig. 18a, b) with domi-
nantly acidic composition indicates ocean-continent convergence that took place
during Late Archaean time (*2.6 Ga). Oceanic or thinner continental lithosphere in
the overlying plate generally produces metaluminous, mafic to intermediate rocks
(Miyashiro 1974; Gill 1981). Thicker continental lithosphere overlying the subduc-
tion zones commonly yields peraluminous, potassic, intermediate to silicic rocks. The
high K calc-alkaline and shoshonite i.e., high K andesite magma (Fig. 19a) indicate
melting at the base of the lithospheric stack. In Th versus Co plot (Fig. 19b) as per
Hastie et al. (2007), the high K calc-alkaline and shoshonite series indicate late stage
ocean-continent subduction. The progression from tholeiite to calc-alkaline to
shoshonite (trachyandesite) reflects increasing K2O and K2O/Na2O and decreasing
iron enrichment (Jakes and White 1972; Miyashiro 1974). The hot, buoyant Archean
ocean lithosphere possibly favoured shallow low-angle subduction, which can be
explained by Al2O3, K2O and SiO2 rich rhyolites, rhyodacites and shoshonites and
plutonic rocks of increasingly granitic composition. This points to a relatively
low-angle subduction (*25°) of thick continental lithosphere (>25 km) with higher
degrees of partial melting (Davies 1992; Ernst 2007; Stern 2008).
Table 2 Selective trace element (ppm) data of the samples from Kadiri granite-greenstone terrain
S. No. Rock Symbol Ba Co Cr Cu Nb Ni Pb Rb Sr Th Zn Zr Y Ce
S-1.8/6 Altered Basic 153 72 348 <10 37 175 82 <10 62 <10 71 156 31 40
S-1.8/9 Diorite 155 37 133 59 22 97 <10 <10 323 <10 60 74 14 14
1.8SP1 Basalt porphyry 1058 19 13 <10 32 <10 18 188 226 <10 149 595 50 164
1.8SP2 Pink Granite 449 24 16 <10 29 <10 45 367 65 82 37 225 33 124
1.8SP3 Diorite 784 22 77 48 21 33 16 114 781 <10 118 198 46 85
1.8SP4 Basalt porphyry 568 24 95 <10 <10 42 12 66 562 <10 93 163 <10 34
1.8SP5 Basalt porphyry 555 29 212 59 <10 49 10 65 873 <10 115 138 32 48
1.8SP6 Granite 442 47 19 <10 27 <10 31 351 97 41 34 177 18 83
1.8SP7 Gray granite 456 25 14 <10 22 <10 32 381 81 40 33 148 18 58
1.8SP8 Diorite 723 24 59 97 20 35 22 114 431 <10 125 205 55 55
1.8SP9 Basalt porphyry 566 37 63 26 <10 32 25 115 295 <10 45 141 19 27
1.8SP10 Granite 430 26.8 16.3 <10 22 <10 48 315 64 71 43 209 26 180
KS1 Schist belt 919 15 26 11 21 15 21 125 508 35 31 142 16 148
Tectonic History of the Granitoids and Kadiri Schist Belt in …
KS2 Schist belt 1489 22 68 <10 28 43 <10 250 291 29 39 146 23 160
KS3 Schist belt 632 14 31 16 28 19 10 169 306 35 28 144 16 147
KS4 Schist belt 358 23 31 16 14 24 35 60 662 24 69 153 13 123
KS5 Schist belt 172 27 29 <10 17 <10 14 107 622 25 28 150 14 123
KS6 Schist belt 557 26 88 21 33 59 11 24 572 <10 97 170 35 80
KS7 Schist belt 1086 22 44 95 12 15 11 67 555 18 <10 166 16 102
269
270 S. Goswami and P. K. Upadhyay
Fig. 17 a The logarithmic Zr versus Ti plot of data points fall in Island arc lava field which
suggest convergent plate margin (after Pearce, Pearce 1982). b Al2O3 versus SiO2 plot (after
Maniar and Piccoli, Maniar and Piccoli 1989). c AFM ternary (after Maniar and Piccoli, Maniar
and Piccoli 1989). Note: IAG = island arc granitoids, CAG = continental arc granitoids, CCG =
continental collision granitoids, POG = post-orogenic granitoids, RRG = rift-related granitoids,
CEUG = continental epirogenic uplift granitoids, OP = oceanic plagiogranites. d R1–R2 binary
diagram (De La Roche et al. 1980) with geotectonic implications (after Batchelor and Bowden
1985)
Fig. 19 a SiO2 versus K2O binary plot (after Peccerillo and Taylor 1976) to distinguish various
series of tholeiitic, calc-alkaline and shoshonitic rocks. b Confirmation of magma series using less
mobile elements, Co versus Th plot (after Hastie et al. 2007)
7 Tectonic Evolution
According to Dey et al. (2014), the East Dharwar craton have Mesoarchaean and
Palaeoarchaean crust in the E. In fact, the typical BADR trend consists of tholeiitic
basalts, high Mg andesites and dacites-rhyolites (Dey et al. 2013) of the Kadiri
schist belt is the signature volcanic rock suite of a convergent margin (Goswami
et al. 2017). Based on the discussion on the different factors it can be said that the
Kadiri greenstone granite terrain rocks indicate convergent margin magmatism
affected by a number of variables, each of which have diversified magma com-
position (Fig. 20a). The eroded volcanic arc materials deposited as volcaniclastics
in the back arc basin. More specifically, in this present context the entire study have
given clue towards back arc basin environment for greenstone belt development,
which is also discussed in case of the Ramagiri schist belt in the W to the Kadiri
terrain (Goswami et al. 2016). The precipitation of chert and BIF atop the
volcano-sedimentary succession can be interpreted from the model. In fact, the
hydrothermal vents released elements in submarine environment and the free
oxygen in the atmosphere started developing during that time by photo dissociation
and some Archaean blue-green algae (*2.6 Ga). Thus the BIF precipitated in the
oxidation-reduction interface (Fig. 20b). The cartoon (Fig. 20c) of the arc and back
arc basin, proposed in this work, shows a two-stage evolution starting with volcanic
eruption followed by erosion. The local extension in the back arc basin also erupted
lava and surrounding areas too experienced characteristic phases of batholithic
intrusions. Small pockets of plumes might have generated later magmatic eruptions
and thus younger granitic intrusions enriched with large ion lithophile elements
(LILE) like uranium. Uranium anomalies in younger granites are located from the
study area (Fig. 1) during this work. The younger granitic apophyses affected the
schist belt volcanics and late shear and tectonics brecciated the rocks. Thus, after
formation and infilling of back arc basin followed by granitic intrusions, the
long-term erosion allowed the upper chert and BIF-rich portion to get transported
by surface agents when the Cuddapah basin started rifting (Fig. 21). This chert,
jasper and BIFs are the representative of significant time of erosion in the form of
basal conglomerate clast indicating the Eparchean nonconformity. Therefore, it can
be concluded that submarine hydrothermal vent activity in the back arc basin is the
source of iron and silica and thus volcanic hot springs released hot waters con-
taining dissolved iron, which probably precipitated after oxydation and interbedded
with volcanic rocks, graywackes etc. Subsequently, emplacement of dykes and
quartz reefs along extension mode cracks possibly during initiation of rifting and
Cuddapah basin development is another episode of tectonic history. The provinance
Tectonic History of the Granitoids and Kadiri Schist Belt in … 273
Fig. 20 Tectonic model showing ocean-continent convergence setting of greenstone belt in which
the back arc basin environment is more specifically the position of succession during first phase
(Episode from *2.8 to 2.1 Ga). a The regional tectonic setting indicates subducting oceanic
lithosphere below lighter continental plate and associated volcanic arc development. The
deposition of eroded arc material in the back arc basin along with dominantly felsic back arc
magmatism must have provided volcano—sedimentary greenstone belt succession. b Cartoon
depicting a detail zoom in back arc basin and BIF, chert precipitation at the final stage. c Cartoon
represents step wise evolution of the convergence setting in which stage 1 shows arc magmatism
followed by erosion and deposition of arc materials in to the back arc basin with simultaneous back
arc basin volcanism
erosion for the newly generated rift supplied initial basin fill and thus conglomerate
with clasts of granite, jasper, chert, quartz veins and BIFs are derived from
basement.
The SIMS U–Pb age (2556 ± 13 Ma) of the felsic volcanics from the Kadiri
greenstone belt (Jayananda et al. 2013b) is corresponding to the Kenoran orogeny
(>2500 Ma). Emplacement of later developed plume-related granitic magmatism of
*2520 Ma (Jayananda et al. 2013a) is reasonably related to the Hudsonian oro-
geny. In fact, the Hudsonian orogeny (*2500–1750 Ma) is an event resulting
extention by granitic plutonism, emplacements of dykes and veins along fractures,
rifting and associated opening of the Cuddapah basin.
274 S. Goswami and P. K. Upadhyay
Fig. 21 Model shows the second phase of the tectonic evolution in which a switch over from
convergence to divergence after a stable phase in between (episode from *2.0 to 1.9 Ga).
a Initiation of continental rifting and associated extension mode cracks development possibly
related to dyke and quartz reef emplacement. b The erosion from topmost portion of basement
provenance (i.e., granite-greenstone terrain) of the newly developing basin are redeposited as
conglomerate-sandstone beds of Gulcheru Formation
Acknowledgements Authors express sincere gratitude to Shri L. K. Nanda, Director, AMD for
encouragement and infrastructure support to publish the part of the assigned work. The support
and help extended by Dr. Syed Zakaulla (RD/SR, Bangalore), Shri. A. K. Bhatt (Dy. RD/SR,
Bangalore) and Shri. V. Natarajan (SO/H) are thankfully acknowledged. We thank Soumyajit
Mukherjee (IIT Bombay) for reviewing and handling this manuscript. The Springer team is
thanked for proof preparation and other assistance. Vide Vanik et al. (2018) for paleostressanalyses
procedure, and Tripathy et al. (2019) and Mukherjee et al. (in press) for other tectonic updates of
the Cuddapah basin. Mukherjee (2019) summarizes this work.
References
Batchelor RA, Bowden P (1985) Petrogenetic interpretation of granitoid rock series using
multicationic parameters. Chemical Geology 48, 43–55
Bowen NL (1928) The evolution of igneous rocks. Princeton University Press, Princeton NJ,
332pp
Cas RAF, Wright JV (1988) Volcanic succesions, modern and ancient. Chapman & Hall, London,
528pp
Chadwick B, Vasudev VN, Hegde GV (2000) The Dharwar craton, southern India, interpreted as
the result of late Archaean oblique convergence. Precambrian Research 99, 91–101
Chadwick B, Vasudev V, Hegde GV, Nutman AP (2007) Structure and SHRIMP U/Pb zircon ages
of granites adjacent to the Chitradurga schist belt: implications for Neoarchean convergence in
the Dharwar craton, southern India. Journal of Geological Society of India 69, 5–24
Chardon D, Jayananda M (2008) Three-dimensional field perspective on deformation, flow, and
growth of the lower continental crust (Dharwar craton, India). Tectonics 27 TC 1014
Tectonic History of the Granitoids and Kadiri Schist Belt in … 275
Chardon D, Jayananda M, Peucat JJ (2011) Lateral constrictional flow of hot orogenic crust:
insights from the Neoarchean of South India, geological and geophysical implications for
orogenic plateaux. Geochemistry, Geophysics, Geosystems 12, Q02005
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology 125, 561–591
Davies GF (1992) On the emergence of plate tectonics. Geology 20, 963–966
De La Roche H, Leterrier J, Grandclaude P, Marchal M (1980) A classification of volcanic and
plutonic rocks using R1R2-diagram and major element analyses – its relationships with current
nomenclature. Chemical Geology 29, 183–210
Dey S, Nandy J, Choudhary AK, Liu Y, Zong K (2013) Neoarchaean crustal growth by combined
arc–plume action: evidences from the Kadiri greenstone belt, eastern Dharwar craton, India. In:
Roberts N, van Kranendonk M, Parman S, Shirey S, Clift P (eds) Continent formation through
time, vol 389. Geological Society of London, Special Publications, pp 135–163
Dey S, Nandy J, Choudhary AK, Liu Y, Zong K (2014) Origin and evolution of granitoids
associated with the Kadiri greenstone belt, eastern Dharwar craton: A history of orogenic to
anorogenic magmatism. Precambrian Research 246, 64–90
Ducea M (2001) The California arc: thick granitic batholiths, eclogitic residues, lithospheric—
scale thrusting, and magmatic fl are—ups. GSA Today 11, 4–10
Ernst WG (2007) Speculations on evolution of the terrestrial lithosphere—asthenosphere system—
plumes and plates. Gondwana Research 11, 38–49
Fisher RV, Schmincke HU (1984) Pyroclastic rocks. Springer, Heidelberg, 474pp
Fliedner MM, Klemperer SL, Christensen NI (2000) Three-dimensional seismic model of the
Sierra Nevada arc, California, and its implications for crustal and upper mantle compositions.
Journal of Geophysical Research 105, 10899–10921
French JE, Heaman LM (2010) Precise U/Pb dating of Paleoproterozoic mafic dyke swarms of the
Dharwar Craton, India: implications for the existence of the Neoarchean supercraton Sclavia.
Precambrian Research 183, 416–441
Friend CRL (1983) The link between granite production and the formation of charnockites:
evidence from Kabbaldurga, Karnataka. In: Atherton MP, Gribble CD (eds) Migmatities,
melting and metamorphism. Shiva Press, Nantwich, pp 264–276
Friend CRL, Nutman AP (1991) SHRIMP U-Pb geochronology of the closepet Granite and
Peninsular Gneiss, Karnataka, South India. Journal of Geological Society of India 38, 357–368
Gill JB (1981) Orogenic andesites and plate tectonics. Springer-Verlag, New York, p 390
Goodwin AM, Ambrose JW, Ayres LD et al (1972) The superior province. Geological Association
of Canada Special Papers 11, 527–624
Goswami S, Bhattacharjee P, Bhagat S, Kumar S, Zakaulla S (2015) Petrography of chert nodules
in stromatolitic dolostone of Vempalle Formation along Tummalapalle - Motnutalapalle,
Cuddapah Basin, India. Indian J Geosci 69:13–24
Goswami S, Sivasubramaniam R, Bhagat S, Kumar Suresh, Sarbajna C (2016) Algoma type BIF
and associated submarine volcano-sedimentary sequence in Ramagiri granite-greenstone
terrain, Andhra Pradesh, India. Journal of Applied Geochemistry 18, 155–169
Goswami S, Upadhayay PK, Bhattacharjee P, Murugan MG (2017) Tectonic setting of the Kadiri
schist belt, Andhra Pradesh, India. Acta Geologica Sinica-english edition. 91(6):1992–2006
Goswami S, Upadhyay PK, Bhagat S, Zakaulla S, Bhatt AK, Natarajan V, Dey S (2018a) An
approach of understanding acid volcanics and tuffaceous volcaniclastics from field studies: a
case from Tadpatri Formation, Proterozoic Cuddapah basin, Andhra Pradesh, India. J Earth
Syst Sci 127:20. https://doi.org/10.1007/s12040-018-0929-0
Goswami S, Dey S (2018b) Facies analysis of tuffaceous volcaniclastics and felsic volcanics of
Tadpatri Formation, Cuddapah basin, Andhra Pradesh, India. Int J Earth Sci (Geol Rundsch).
https://doi.org/10.1007/s00531-018-1620-z
Grove TL, Kinzler RJ (1986) Petrogenesis of andesites. Annual Review of Earth and Planetary
Sciences 14, 417–454
276 S. Goswami and P. K. Upadhyay
Halls HC, Kumar A, Srinivasan R, Hamilton MA (2007) Palaeomagnetic and U-Pb geochronology
of easterly trending dykes in the Dharwar Craton, India: feldspar clouding, radiating dyke
swarms and the position of India at 2.37 Ga. Precambrian Research 155, 47–68
Harker A (1909) The natural history of igneous rocks. McMillan Publishers, New York, 384pp
Hastie AR, Kerr AC, Pearce JA, Mitchell SF (2007) Classification of altered volcanic island arc
rocks using immobile trace elements: development of the Th, Co discrimination diagram.
Journal of Petrology 48, 2341–2357
Hawkesworth CJ, Gallagher K, Hergt JM, Keynes M (1993) Mantle and slab contributions in arc
magmas. Annual Review of Earth and Planetary Sciences 21, 175–204
Irvine TN, Barager WRA (1971) A guide to the chemical classification of the common volcanic
rocks. Canadian Journal of Earth Sciences 8, 523–548
Jakes P, White AJR (1972) Major and trace element abundances in volcanic rocks of orogenic
areas. Bulletin of Geological Society of America 83, 29–40
Jayananda M, Mahabaleshwar B (1991) Relationship between shear zones and igneous activity:
the Closepet graninte of southern India. Indian Academy of Sciences (Earth and Planetary
Sciences) Proceedings 100, 31–36
Jayananda M, Tsutsumi Y, Miyasaki T, Gireesh RV, Kapfo Kowe-u, Tushipokla Hiroshi Hidaka,
Kano T (2013a) Geochronological constraints on Meso- and Neoarchean regional metamor-
phism and magmatism in the Dharwar craton, southern India. Journal of Asian Earth Sciences
78, 18–38
Jayananda M, Peucat JJ, Chardon D, Krishna Rao B, Fanning CM, Corfu F (2013b) Neoarchean
greenstone volcanism and continental growth, Dharwar craton, southern India: Constraints
from SIMS U-Pb zircon geochronology and Nd isotopes. Precambrian Research 227, 55–76
Maniar PD, Piccoli PM (1989) Tectonic discriminations of granitoids. Geological Society of
America Bulletin 101, 635–643
Mcphie J, Doyle M, Allen R (1993) Volcanic Textures. A guide to the interpretation of textures in
volcanic rocks. Tasmanian Government Printing, Office, Tasmania, p 196pp
Misra AA, Mukherjee S (2015) Tectonic inheritance in continental rifts and passive margins.
Springerbriefs in Earth Sciences. ISBN 978-3-319-20576-2
Misra AA, Mukherjee S (2018) Atlas of structural geological interpretation from seismic
images. Wiley Blackwell. ISBN: 978-1-119-15832-5
Misra AA, Bhattacharya G, Mukherjee S, Bose N (2014) Near N-S paleo-extension in the western
Deccan region in India: does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences 103, 1645–1680
Misra AA, Sinha N, Mukherjee S (2015) Repeat ridge jumps and microcontinent separation:
insights from NE Arabian Sea. Marine and Petroleum Geology 59, 406–428
Misra AA, Sinha N, Mukherjee S (2018a) The gop rift: a paleo slow spreading centre, Offshore
Gujarat, India. In: Misra AA, Mukherjee S (eds) Atlas of structural geological interpretation
from seismic images. Wiley Blackwell, pp 208-212 ISBN: 978-1-119-15832-5
Misra AA, Sinha N, Mukherjee S (2018b) The Ratnagiri fracture zone: a paleo oceanic‐fracture‐
zone in the Mumbai‐Ratnagiri Offshore Region, West India. In: Misra AA, Mukherjee S
(eds) Atlas of structural geological interpretation from seismic images. Wiley Blackwell,
pp. 195-199. ISBN: 978-1-119-15832-5
Miyashiro A (1974) Volcanic rock series in island arcs and active continental margins. American
Journal of Science 274, 321–355
Mukherje S (2015) Atlas of structural geology. Elsevier, Amsterdam
Mukherjee S (2010a) Structures in Meso- and Micro-scales in the Sutlej section of the Higher
Himalayan shear zone, Indian Himalaya. e-Terra 7, 1–27
Mukherjee S (2010b) Microstructures of the Zanskar shear zone. Earth Science India 3, 9–27
Mukherjee S (2011a) Flanking Microstructures from the Zanskar shear zone, NW Indian
Himalaya. YES Bulletin 1, 21–29
Mukherjee S (2011b) Mineral Fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth Sciences 100, 1303–1314
Tectonic History of the Granitoids and Kadiri Schist Belt in … 277
Mukherjee S (2012) Tectonic implications and morphology of trapezoidal mica grains from the
Sutlej section of the Higher Himalayan shear zone, Indian Himalaya. Journal of Geology 120,
575–590
Mukherjee S (2013) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy,
Berlin, pp 1–111. ISBN 978-3-642-25608-0
Mukherjee S (2014a) Mica inclusions inside host mica grains from the Sutlej section of the Higher
Himalayan Crystallines, India- morphology and constrains in genesis. Acta Geologica Sinica
88, 1729–1741
Mukherjee S (2014b) Review of flanking structures in Meso- and Micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Chakraborty R (2007) Pull-apart micro-structures and associated passive folds. In:
Aho J (ed) Annual transactions of the nordic rheology society 15, pp 247–252. 16th Nordic
Rheology Conference, Stavanger, Norway, 13–15 June 2007
Mukherjee S, Koyi HA (2009) Flanking microstructures. Geological Magazine 146, 517–526
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Zanskar Section-microstructural
studies & extrusion mechanism by a combination of simple shear & channel flow. International
Journal of Earth Sciences 99, 1083–1110
Mukherjee S, Koyi HA (2010b) Higher Himalayan Shear Zone, Sutlej Section-structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Mukherjee S, Goswami S, Mukherjee A (In press) Structures and their tectonic implications form
the southern part of the Cuddapah basin, Andhra Pradesh, India. Iranian J Sci Technol Trans A:
Sci. https://doi.org/10.1007/s40995-018-0566-0
Nandy J, Dey S (2013) The mechanism of Neoarchaean granitoid formation: evidence from
Eastern Dharwar Craton, Southern India. American International Journal of Research in
Formal, Applied and Natural Sciences (AIJRFANS) 3, 105–109
Oak KA (1990) The geology and geochemistry of the Closepet granite, Karnataka, South India.
Unpublished Ph.D. thesis, Council for National Academic Awards, Oxford Polytechnic, UK
Pearce JA (1982) Trace element characteristics of lavas from destructive plate boundaries. In:
Thorpe RS (ed) Andesites: orogenic andesites and related rocks. Wiley, Chichester, pp 525–
548. ISBN 0 471 28034 8
Pearce JA, Peate DW (1995) Tectonic implications of the composition of volcanic arc magmas.
Annual Review of Earth and Planetary Sciences 23, 251–285
Peccerillo A, Taylor SR (1976) Geochemistry of Eocene calc-alkaline volcanic rocks from the
Kastamonu area, Northern Turkey. Contributions to Mineralogy and Petrology 58, 63–81
Perfit MR, Gust DA, Bence AE, Arculus RJ, Taylor SR (1980) Chemical characteristics of
island-arc basalts: implications for mantle sources. Chemical Geology 30, 227–256
Rajamani V (1990) Petrogenesis of Metabasites from the schist belts of Dharwar Craton:
implications to Archean Mafic Magmatism. Journal of Geological Society of India 36, 565–
587
Rajamani V, Shivkumar K, Hanson GN, Shirey SB (1985) Geochemistry and petrogenesis of
amphibolites, Kolar Schist Belt, South India; evidence for komatiitic magma derived by low
percent of melting of the mantle. Journal of Petrology 96, 92–123
Rowland SK, Walker GPL (1990) Pahoehoe and aa in Hawaii: volumetric flow rate controls the
lava structure. Bulletin of Volcanology 52, 615–628
Satyanarayana K, Siddilingam J, Jetty J (2000) Geochemistry of Archean Metavolcanic Rocks
from Kadiri Schist Belt, Andhra Pradesh, India. Gondwana Research 3, 235–244
Shackleton RM (1976) Shallow and deep-level exposures of the Archean crust in India and Africa.
In: Windley BF (ed) The Early History of the Earth. Wiley, New York, pp 317–322
Shand SJ (1943) Eruptive rocks. Their genesis, composition, classification, and their relation to
ore-deposits with a chapter on meteorite. Wiley, New York
278 S. Goswami and P. K. Upadhyay
1 Introduction
the recent years (around 2012), hydrocarbon discoveries in the basement at depths
around 1400 m from a few fields have indicated the existence of an additional
basement play which may have a significant role in the overall hydrocarbon
exploration scenario. The present study based on a basement exploration perspec-
tive has been made to decipher different evidences of structural deformation of
basement in outcrops and subcrops in different scales pertaining to the various
episodes of tectonism affecting the basin. These structural signatures derived from
independent G&G studies were subsequently integrated in a GIS platform to build a
conceptual model for basement exploration and thus delineate optimal areas where
the basement can act as a promising hydrocarbon reservoir.
The Cauvery Basin is a peri-cratonic rift basin in the southern part of India
hosting a sedimentary sequence ranging in age from Late Jurassic to Recent
(Mukherjee 2015a, b). The basin is bounded at west by exposures of the
Pre-Cambrian South Indian Shield that rests unconformably below the sedimentary
layers forming the basement to this basin. A number of NE oriented basement highs
divide the basin into alternating ridge and sub basin areas, each of which form
discrete fault-bound blocks (Fig. 1a). In the north of the basin, another basement
high known as the Chingelput High (Mazumder et al. 2013a, b) separates the
Cauvery Basin from the adjacent Palar Basin whereas in the eastern part, the
Cauvery Basin continues underneath the Bay of Bengal as the Cauvery offshore.
Fig. 1 a General tectonic & geographical set up of the Cauvery Basin showing different
sub-basins and their dividing ridges b Generalized stratigraphy encountered in the Cauvery Basin
in outcrops and subcrops depicting the different source rock sequences (After Rangaraju et al.
1993; dghindia.com 2011; Nagendra et al. 2017)
Basement Tectonics and Shear Zones … 281
The basinal area in onshore part is covered by a thick layer of alluvium that
conceals structural and geological information.
The sedimentary sequence of the Cauvery Basin along with its basement (Fig. 1b)
has suffered multiple episodes of tectonism. These different episodes of tectonism
and their effects in the basement are described below.
them probably the southern margin of the Dharwar craton (Chetty and Santosh
2013). The D2-episode is characterized by transpressional tectonic regime with
extensive dextral shear and migmatisation (Chetty et al. 2003). Here it is to be noted
that in general shear fabrics in shear zones are developed ubiquitously in migma-
tites (Mukherjee and Koyi 2010a; Mukherjee 2013, 2014a, 2015a, b). These E-W
individual shear zones are presumably rooted to a decollement zone in the upper
mantle forming a crustal-scale ‘flower structure’ (Chetty and Bhaskar Rao 2006;
Chetty 2015) implying that individual shear zones branch out of a deeper and
thicker zone: vide fig A5 in the Appendix/Repository of Dasgupta and Mukherjee
(2017) for a meso-scale flower structure from a different terrain. Of these shear
zones, the MBSZ in the northern boundary of CSZ dip steeply southward whereas
the other constituent shear zones dip to north with moderate values.
The shear zones of CSZ have been mapped in the outcrop, but their disposition
below the Cauvery Basin plays a major role controlling the basement morphology and
guiding later development of structures in the northern part of the Cauvery Basin
(Murty et al. 2002). As such, CSZ influences basement exploration in the areas of
Ariyalur-Pondicherry Sub Basin, Kumbakonam-Madanam ridge, Tranquebar-
Tanjore Sub Basin, Karaikal ridge and Nagapattinam Sub Basin (Fig. 2).
Fig. 2 Layout of the Cauvery Shear Zone (after Chetty 2015; Chardon et al. 2008; Plavsa 2014;
Gupta et al. 2015) consisting of broadly E-W oriented shear zones with most of it impacting
basement tectonics of Ariyalur-Pondicherry Sub Basin, Kumbakonam-Madanam Ridge,
Tranquebar Sub Basin, Karaikal-Vedaranniyam Ridge and Nagapattinam Sub Basin of Cauvery
Basin. MBSZ Moyar Bhawani Shear Zone, PCSZ Palghat Cauvery Shear Zone, NMB Northern
Madurai Block and SMB Southern Madurai Block
Basement Tectonics and Shear Zones … 283
Other than the formation of shear zones during Precambrian, the basement of
Cauvery Basin has also undergone three major episodes of tectonism that governed
the development of basement structures.
One of the probable episodes of tectonism affecting the basement is the formation
of N-S faults during Lower Gondwana (Late Carboniferous—Permian). Evidences
of these tectonic phase exists as a number of N-S morphotectonic faults, major N-S
lineaments (Subrahmanyam et al. 1995; Saravanavel and Ramasamy 2003;
Prabaharan et al. 2013) and magnetic lineaments mapped in the onland Cauvery
Basin and offshore (Subrahmanyam et al. 1995). These structural trends are sym-
pathetic to the southern part of the adjacent Palar rift that hosts Ongur Formation
(Fig. 3a) dated to be Late Carboniferous to Early Permian (Rangaraju et al. 1993;
Vairavan 1993). In the Cauvery Basin, well P-2 drilled in Kumbakonam Ridge
along the N-S trend has also yielded palynofossil assemblage equivalent to Ongur
Formation (Prasad and Phor 2009). This implies the existence of N-S faults/rifts
continuing from Palar Basin in the north and affecting at least the northern part of
basement of Cauvery Basin (Fig. 3a).
The principal event that initiated and shaped the morphology of Cauvery Basin is
the NE trending rifting associated with the separation of India-Sri Lanka system
from Antarctica and Australia in Late Jurassic to Early Cretaceous (Veevers and
Tewari 1995; Gaina et al. 2007; Veevers 2009; Lal et al. 2009; Nemcok et al. 2013).
The continental separation associated with rifting began at Late Jurassic (160 ma)
and produced genetically related rift systems of Cauvery, Palar, Pennar and
Krishna-Godavari. In Cauvery and Palar Basins the separation probably occurred
along pre-existing NE-SW oriented weaknesses in the Pre Cambrian basement
(linkages between shear zones) that may have guided later structures (Lal et al.
2009; Dasgupta and Mukherjee 2017). Such reactivation along existing fabric is
called “tectonic inheritance” (Misra and Mukherjee 2015). This led to the formation
of a series of major NE-SW horst and graben features in the basement dividing the
basin into a number of fault-bound ridges and sub-basin areas (Lal et al. 2009).
Fig. 3 a N-S oriented morphotectonic faults and lineaments (Subrahmanyam et al. 1995;
Saravanavel and Ramasamy 2003; Prabaharan et al. 2013) probably formed during Lower
Gondwana (Lower Carboniferous—Permian) that are sympathetic to the Ongur rift in Palar Basin
and also host well P-2 in Thanjavur Sub Basin that yielded palynoflora equivalent to Ongur
Formation. b Formation of NW-SE cross trends (Saravanavel and Ramasamy 2003; Prabaharan
et al. 2013) that had affected the basement as a result of simultaneous activity of the NE-SW rifting
and N-S strike slip faulting (Mazumder et al. 2013)
a transcurrent fault (Lal et al. 2009) or a dextral strike-slip transform fault (Nemcok
et al. 2013) along which Antarctica moved southward. The combined and simul-
taneous effect of the NE-SW rift trend and NNW trending strike slip fault probably
led to the development of NW-SE cross trends (Mazumder et al. 2013a, b). This
Basement Tectonics and Shear Zones … 285
also resulted in a rotation of the earlier rift trends (Mazumder et al. 2013a, b)
causing a wide zone of fracturing along the rotational axis (Fig. 3a, b).
These tectonic phases in their respective times have its individual sympathetic
structural signatures that constitute part of the fabric of the basin, controlling its
evolution, sedimentation pattern to even the distribution of hydrocarbon systems.
Results of these phases of tectonism may also have led to parts of the basement
acting as a hydrocarbon reservoir as observed from the results of a number of wells
drilled by ONGC in the basin with depths varying from 2000 to 3500 m.
Now, in order to act as an effective reservoir rock, the basement should have the
following characteristics:
(a) Most basement hydrocarbons are hosted in structural highs of varying but
generally moderate to large elevation. The highs are formed by fault-controlled
blocks, often in rift settings like upthrown sides of faulted block, horsts within a
graben (Chung 1982) or by palaeo-hills buried unconformably below sedi-
mentary cover (Gutmanis 2009).
(b) Younger sediments, which act as hydrocarbon sources, either flank or directly
overlie the basement (Sircar 2004). Charging mostly takes place laterally or by
up-dip migration from adjacent structural lows, however, downward migration
may also take place from overlying source rocks into underlying basement
(Gutmanis 2009).
(c) Most basement rocks are characterized with low matrix porosity (typically
<0.5%) and permeability. As a result, reservoir quality in basement rocks
depends on the development of secondary porosity and the presence of a
connected open fracture network. A number of such fractured reservoirs in
basement exist worldwide e.g. Aerxia Oil Field in China with well-developed
joints, faults and fractures in phyllite, slate and meta-sandstone (Sircar 2004),
White Tiger field, Vietnam constituted of fractured granite basement (Dang
et al. 2011), Dayung Field in South Sumatra with fractured basement com-
prising Permian carbonates intruded by granites (Darmadi et al. 2013) and
fractured basement reservoir in Zeit Bay oil field, Gulf of Suez, Egypt com-
posed of pegmatic granites (Zahran and Askary 1988). This secondary porosity
consists of results of structural deformation like joints, faults, fractures ranging
in scale from microfractures to seismic scale faults and their damage zones
(Sircar 2004; Satyanaryana et al. 2010). As a result, the porosity and perme-
ability values in a basement reservoir are irregularly distributed laterally as well
as in depth, In case of White Tiger field, the largest field of Vietnam for
basement oil, maximum porosity is found to be 4% with permeability occa-
sionally reaching 0.8D (Huy et al. 2012). However, dense micro-cracks asso-
ciated to major fault structures that may be formed as a result of multiple
286 S. Mazumder et al.
The present study presumes that signatures of structural deformation resulting from
a particular episode of tectonics are similar in pattern but may vary in their scale
depending upon the tools and perspective used to view them. Based on this premise,
pattern of signatures of a particular phase of tectonism observed in outcrops may be
repeated in larger scale in subsurface seismic data or in further regional scale in
remote sensing images. In accordance to the above assumption, structural patterns
Basement Tectonics and Shear Zones … 287
in the basement below the Cauvery Basin are extracted from different geophysical
and geological data sets. These are subsequently compared with field analogues and
integrated to prepare a conceptual model for basement exploration (Fig. 4).
In this part of the study, the Cauvery Basin and the Precambrian basement outcrops
beyond it are analysed to delineate structures exposed and manifested on the
surface.
In the basinal area, a neotectonic based structural analysis/morphostructural
analysis is carried out on principles discussed in Mazumder et al. (2011, 2013a, b,
2016) to detect surface manifestations of deep-seated/basement-related structures.
These basement related structures include both faults and structural highs. Both of
these elements reactivate repeatedly in the neotectonic regime and are manifested
on the surface as anomalies in geomorphic elements like topography, soil and
especially drainage, which form the most sensitive parameter. Such neotectonic
manifestations like fluvial anomalies, seismogenic lineaments and seismicity are
well distributed and well-studied in the Cauvery Basin (Murty et al. 2002; Kumaran
and Ramasamy 2005; Ganapathy and Rajarathinam 2010; Prabaharan et al. 2013,
288 S. Mazumder et al.
Basement Tectonics and Shear Zones … 289
JFig. 5 a Drainage features used to interpret probable fault segments that may be manifestations of
basement faults. b Geomorphic highs indicating sub surface structural highs interpreted based on
peripheral drainage geometry. Associations with elements like barbed drainage and rectangular
drainages imply a greater degree of confidence of representing sub surface structural highs
(Trenchard 2007). c Evidences of folding and jointing in the exposed basement near Salem
delineated from LANDSAT ETM+ band combination 7, 4 and 1. d Shear zone exposed in a river
bed in Belur near Salem oriented E-W with earlier NE-SW foliations rotated parallel to sense of
shearing. e Joints developed in Charnockite in the exposed basement. f Fault oriented NW-SE in
the Cuddalore Formation near Neyveli. Geographical positions of the geomorphic, outcrop and
image based features cited above marked in the Fig. 5g
Singh et al. 2016; Resmi et al. 2016 etc.). In this study, features in drainage
geometry (e.g. drainage offsets, rectangular drainages, compressed meanders)
(Fig. 5a), abrupt changes in surface slope and lineaments derived from tonal
variations in images are used to delineate fault segments. Similarly, geomorphic
highs forming surface expression of probable structural highs in basement are
derived from features like radial and peripheral drainages. Associations with fea-
tures like barbed drainages, rectangular drainage patterns and low moisture contents
are used to build up and classify these geomorphic highs as per their degree of
confidence based on their probabilities in representing basement structural highs
(Fig. 5b). These morpho-structural interpretations are subsequently validated by
extensive field checks where the mapped structures are checked for neotectonic
faulting imprints and evidences like unpaired terracing, abnormal incisions and
sudden appearance of braiding.
In areas beyond the basin margin of Cauvery Basin, extensive field work and
image analysis have been carried out to identify structural features in the exposed
basement rocks. LANDSAT ETM+ (EarthExplorer, USGS) band combinations of
band 7, 4 and 1 have been pan-sharpened, contrast enhanced and edge filtered to
identify different regional scale structural features (Fig. 5c). Abrupt termination of
basement outcrops to delineate faults, instances of folding in basement lithology,
prominent lineament patterns have been deciphered from separate image data sets.
As a part of fieldwork in these outcrops, attitude of joint sets and fractures (Fig. 5e)
developed in basement lithology were measured, characterized and used to identify
trends of their causative regional faults. Outcrop scale deformations like folding,
faulting (Fig. 5f) and shearing (Fig. 5d) were also analysed in view of their
geometry and inter-relationships for modelling of subsurface basement features.
These two sets of surface structural data pertaining to the alluvial covered
basinal area and the exposed basement area were correlated and joined in a GIS
platform to understand the disposition of basement structures in the basinal area
(Fig. 6).
290 S. Mazumder et al.
Fig. 6 Map depicting the surface disposition of the principal basement trends and geomorphic
highs in the basinal area overlain on geological map (after Desikachar et al. 1960). The
geomorphic highs are arranged as per degree of confidence to represent subsurface structural
highs. The surface trends in the rectangular area has been correlated with faults at basement top in
Fig. 7
Basement Tectonics and Shear Zones … 291
3D Pre-Stack Time Migrated (PSTM) seismic data used for imaging subsurface
datasets exhibiting complex geological structures recorded in the Cauvery Basin
area has been used to validate and establish the continuity of the surface mor-
phostructural interpretations and to identify the structural effects of different phases
of tectonism.
An overlay analysis of faults mapped on basement top from seismic data in
Kali-Kuthalam area (Fig. 6) with surface basement trends and morphostructural
faults shows a positive correlation. This probably implies the basement related
subsurface faults may have cut through Mesozoic and Tertiary sections as a result
of multiple reactivations of them and have been manifested on the surface due to
present day neotectonism affecting Cauvery Basin (Fig. 7). This also suggests the
interpreted surface trends represent the basement faults.
In a further attempt to correlation, seismic time slices acquired at 2000 ms time
equivalent depth (within basement, near basement top) in Madanam and Kuthalam
area (Fig. 8a) had been digitally filtered in an image processing software to highlight
and extract lines of high seismic amplitude. Here it is to be noted that though crys-
talline formations are associated with lack of acoustic contrast, however structures on
or near basement top can be identified (Gutmanis 2009). These extracted lines appear
to define a similar geometry to the E-W oriented folds observed in image data (Fig. 8c)
and in outcrops (Fig. 8d) and also display an analogous association with the E-W
faults as observed in those areas. As such, it is possible that these features are sheath
folds defined as highly non-cylindrical folds (Reber et al. 2013) that may be produced
as a result of pronounced ductile shear (Mukherjee 2014b; Mukherjee et al. 2015).
Similar sheath folds have been mapped by Chetty et al. (2012) in Mahadevi Hills in the
axial zone of Cauvery Shear Zone that are also characterized by E-W axial traces
(Fig. 9). The E-W trends probably act as manifestations of the shear zones in basement
with the sheath folds forming contemporaneously with the dextral shearing along
them. Fold amplifies during shearing with the fold axes rotating towards shear
direction. At the highest strain level, the fold axes becomes sub-parallel to shear
direction (Burg 2017a, b) and existing foliations of the country rock are found to be
deflected parallel to the E-W trends (Fig. 5d).
In addition, seismic sections are correlated with surface interpretations to validate
the morpho-structural manifestations of basement top morphology (Fig. 5b). The
correlation indicates that faults affecting basement top when extrapolated up-dip
match with the surface fault trends re-iterating the fact that they are reactivation related
manifestations of the basement faults (Fig. 9e). Also, most of the geomorphic highs
spatially coincide with the areas of basement highs associated with a high amount of
structural relief. Since most of the interpreted geomorphic highs are found to be
associated with NE trending surface features, probably the NE faults create basement
highs with a higher structural relief (Fig. 9a). Other than these structural highs, the
basement top is also found to depict a series of smaller undulations with much smaller
variation in basement relief compared to the structural highs. These small undulations
292 S. Mazumder et al.
Fig. 7 Correlation of faults mapped using seismic data at basement top with surface basement
trends corresponding to the area depicted in Fig. 6
are found to correlate with the sheath folded layers delineated from time slices with
individual separated by the E-W trending faults (Fig. 9b). Since both the sheath folded
layers and the E-W trending faults are associated with shearing, it indicates that the
series of small undulations at the basement top are representative of shear zones in a
sectional view. Analogous signatures are also observed in sectional views of shear
Basement Tectonics and Shear Zones … 293
zones exposed beyond the basin margin (Fig. 9c) as well as other shear zones reported
in public domain (Fig. 9d).
Since the shear zones (Mukherjee and Mulchrone 2015) associate with E-W
surface trends, their spatial disposition provides an estimate of the basinal extent
affected by shearing. The results indicate that most of the trends are concentrated in
the area north of the Cauvery Basin (Fig. 10a). Seismic sections near
Ramnathapuram area (Fig. 10b) do not show any effects of the shear zones in
basement like undulations or sheath folds suggesting no such Pre-Cambrian base-
ment deformations in the area. However, basement related structural highs exist and
are manifested as geomorphic highs (Fig. 10b).
Fig. 8 a Seismic time slices in Madanam and Kuthalam area at 2000 ms time equivalent depth
(within basement, near basement top) as shown in the index map. b Lines of high seismic
amplitude extracted from the basement time slices appear to define a folded pattern with an axial
trace that is parallel to the E-W basement trends mapped at the surface. These patterns are similar
to the E-W trending sheath folds defined in image. c (LANDSAT ETM+ band 8) and outcrop scale
(8d) near Salem
294 S. Mazumder et al.
Fig. 8 (continued)
Formation Microimager logs (FMI) recorded in deep wells M-5 and M-6 drilled by
ONGC along the E-W faults were analysed to determine the fracture patterns and
porosity associated with the shear zones. An orientation analysis of fracture azi-
muths derived from FMI logs for every 50 m interval from the basement top
indicates that E-W fractures occur in zones of 200–300 m thickness. These frac-
tures are mostly regularly spaced and found to be oriented along E-W foliations
(Fig. 11a, b and d). The E-W foliated and fractured zones represent the signatures
of shear zones. Observations from FMI data indicate that these zones are separated
by areas of un-sheared rock where other tectonic trends are dominant. Similar
features are observed in cores cut in deep wells drilled by ONGC along the E-W
Basement Tectonics and Shear Zones … 295
Both gravity and magnetic data helps to build a regional representation of basement
configuration. Deep-seated regional scale structural elements affecting basement are
manifested as gravity or magnetic signatures. Based on this premise, E-W signa-
tures were derived from gravity derivative (Kumar et al. 2009) and reconnaissance
scale aeromagnetic data over Cauvery Basin (Rajaram and Anand 2014). Also since
these E-W shear zones are believed to extend offshore (Murty et al. 2002), the E-W
trending faults mapped by Subrahmanyam et al. 1995 had also been extracted from
the magnetic map in offshore areas.
An analysis of the different independent surface and subsurface data sets thus
provided a number of E-W basement related trends that can be regarded as com-
ponents of shear zone. These are (a) E-W surface trends derived from geomorphic
and image data (b) Shear zones and faults mapped in field (Chardon et al. 2008;
Desikachar et al. 1960) (c) E-W trending faults from seismic maps at top of
basement level in both onshore and offshore Cauvery Basin d) E-W trends extracted
from gravity derivative maps (Kumar et al. 2009) and (e) E-W trends derived from
aero-magnetic data (Rajaram and Anand 2014) and offshore magnetic data
(Subrahmanyam et al. 1995). All these components signifying parts of a shear zone
are joined in a GIS platform to define a probable shear zone network in Cauvery
Basin (Fig. 12) maintaining protocols like thickness and spacing constraints, rela-
tionship with sheath folded layers derived from basement time slices and available
well control in terms of both FMI logs and well data.
296 S. Mazumder et al.
Basement Tectonics and Shear Zones … 297
JFig. 9 a Seismic sections along AB, CDE, FG and XY (as in inset map) correlated with surface
faults and geomorphic highs. Most of the geomorphic highs in the surface are found to overlie
areas of basement high that are associated with a high amount of structural relief. b Basement top
in seismic section also depicts a series of smaller undulations characterized with a much smaller
variation in relief. These small undulations (encircled) correlate with the sheath folded layers
extracted from time slice and are found to be separated by the E-W trending faults (in red). c Shear
zone exposed in sectional view in Manmedu near Tiruchurapalli displays similar such undulation.
d Undulations also observed in the cross section of Homestake Shear Zone, Hornsilver Ridge,
Colorado similar to those interpreted in seismic data (Shaw et al. 2002). e Faults affecting
basement top when extrapolated up-dip are found to correlate with interpreted surface fault trends
(especially oriented E-W) implying that these surface trends are reactivation related manifestations
of the basement faults
Fig. 9 (continued)
3 Discussions
The Precambrian shear zones are considered products of deep and ductile defor-
mation with the concentration of large strain. However, with decrease in depth and
associated decrease in temperature and pressure, wider zones of ductile displace-
ment are transformed into discrete planes and narrow zones of brittle displacement
(Burg 2017a, b). The areas affected by the shearing show higher fracture density
than the host rock, with the highest fracture densities restricted to fine-grained and
strongly foliated sections (Laws et al. 2003).
In the present case, dextral shearing (D2) leads to the formation of a shear
foliation and sheath folding aligned along the E-W shear zones. Most of these shear
zones are continuous from the basement exposures to the basinal part and show
298 S. Mazumder et al.
Fig. 10 a E-W trends appear to be concentrated in the northern part of the Cauvery Basin
indicating that the effect of shear zone on basement is restricted to solely that area. b Seismic
sections along lines AB and CD in southern part of Cauvery Basin showing no indications of
shearing (GH stands for Geomorphic Highs)
similar type of sheath folding and foliation properties both in outcrops and subcrops
(Fig. 13c). However, upon reaching a critical length, the shear zones link obliquely
along step over areas (Schrank et al. 2008) forming NE-SW oriented linkages
between the adjacent shear belts. The shear foliations along the E-W shear belts
exposed at the basement top level changes to a weakly bonded planar fabric.
Deformations occurring subsequent to the formation of the shear zones reactivated
these planes of weakness, both along the shear zones and their linkages eventually
forming fractures oriented along the foliations. These areas of fracturing due to the
formation of shear zones are about 250–300 m in orthogonal thickness and are
associated with a moderate to steep dip. As such, these shear zones could have
branched out of a thicker, gentler and a more ductile layer in the deeper part of
basement (Fig. 13a, b). Hence it is possible that more than one shear zone could be
encountered in a particular drilled well that are separated by an interval of
un-sheared and less fractured rock thus increasing the probability of encountering a
zone of high porosity (secondary).
The isotopic boundary probably forms the limit of the area where the basement
is affected by shearing.
Basement Tectonics and Shear Zones … 299
Other than the Precambrian tectonism creating E-W shear zones, the basement is
deformed by to the three episodes of Phanerozoic tectonism as discussed earlier.
Each of these phases probably inherited and reactivated older fracture patterns
within basement depending upon the state of stress prevalent at the time. Of these
episodes, the NE-SW trending main rifting phase in Late Jurassic-Early Cretaceous
had the most dominant effect on basement. Additionally, each deformation phase is
instrumental in reactivation of faults/fractures that were created in an earlier phase.
The areas of intersection of Phanerozoic basement trends demarcated from their
surface manifestations with the shear zones will form the areas of maximum
fracture porosity in Cauvery Basin (Figs. 11a, d and 14). These are also the areas of
maximum reactivation resulting in extensive fracturing along the shear zones with
deep, open and connected fracture networks increasing secondary porosity in the
area. The areas of intersections of the different sets of fractures form an intercon-
nected fracture mesh increasing the areas of permeability in the region.
The Cauvery Basin forms a neotectonically active area with number of earthquake
epicentres and seismic prone lineaments mapped in its geographical extent
(Saravanavel and Ramasamy 2003; Ganapathy and Rajarathinam 2010).
Accordingly, most of its structural and tectonic elements can be considered to be
neotectonically reactivated. The in situ tectonic stress acting on the northern part of
the basin as deduced from different sources like Drilling Induced Fractures in
different wells (Fig. 15a), Focal Plane Solutions (Murty et al. 2002), World Stress
Maps (Heidbach et al. 2016) and works of Chatterjee and Mukhopadhyay (2008),
Subrahmanyam (2014) indicate SHmax direction oriented ENE-WSW to NW-SE.
In the present analyses, a trend based correlation of morphostructural faults at the
surface with faults mapped at the top of basement from seismic data in the
Madanam area indicates that the E-W trends are more reflected in the surface
manifestations (Fig. 15b). This specifies that under the present stress regime the
E-W faults are the most reactivated. Consequently, the E-W shear zones turn out to
be the most neotectonically reactivated structures in the basement. This implies that
the shear zones are associated with a greater damage zone constituted of a cluster of
incipient fractures with the damage zone width increasing with each phase of
reactivation associated displacement. This further enhances the reservoir character
of the shear zone area.
Although the shear zone affected areas of basement can effectively act as
reservoir with secondary porosity, it needs to be associated with a younger sedi-
mentary source for charging. These areas of basement either need to directly overlie
the hydrocarbon source horizon or needs to be connected to them by faults acting as
300 S. Mazumder et al.
Basement Tectonics and Shear Zones … 301
JFig. 11 a Trend analysis of fractures of wells M-5 and M-6 at every 50 m interval from basement
top indicates 200–250 m thick zones with dominance of E-W fracture sets. b FMI of well M-6
shows regularly spaced E-W fractures oriented along a zone of E-W foliations. c DSI logs of well
P-6 correlated with FMI indicate that the E-W oriented fractures are open and deep. d Field analog
showing E-W oriented open fractures associated with shear foliations in Attur (Fig. 9). Imprints of
fracture sets oriented NE-SW (along linkages) and E-W lead to an increase of fracture porosity in
the area
Fig. 12 Probable shear zone network derived from different components in the Cauvery Basin
(pink indicates ridges and blue indicates sub-basins)
conduits. For Cauvery Basin, shales within Andimadam Formation and Sattapadi
Shales of Early Cretaceous are considered the main source kitchen for the area. The
critical moment or the time of maximum hydrocarbon generation prior to migration
is observed to be around 65.5 Ma (Phaye et al. 2011). Since after Early Cretaceous,
no major tectonic event occurred in the basin other than subsidence and tilting
(Nagendra et al. 2017; DGH India), the in situ stress direction is considered similar
to that prevailing at the time of hydrocarbon migration. With this assumption, based
on principles in Zhang et al. (2011), E-W faults being parallel to the in situ stress
direction tend to remain open and act as conduits of migration in
Eocene-Oligocene. As a result they play a significant role in charging basement
reservoirs by updip/lateral migration from adjacent kitchen areas into structural
highs (Chandrasekhar et al. 2015). This is corroborated by adsorbed gas anomalies
302 S. Mazumder et al.
in onland areas and natural hydrocarbon seepages in offshore (Dave et al. 2012)
both of which are considered as manifestations of migration/remigration conduits
that show a positive correlation with E-W shear zones (Fig. 15c). This suggests that
the E-W shear zones and associated faults are most instrumental in migration/
Fig. 13 a Probable model of arrangement of the shear zones branching off from a thicker and
deeper shear zone b Field analog from Manmedu showing shear zones in a cliff section branching
from each other enveloping unsheared rock in between c E-W oriented shear zones found to be
continuous from the basement exposures to the basinal part, showing similar type of fracturing
properties in outcrop in Manamedu and in FMI of subcrop in well M-5 (pink indicates ridges and
blue indicates sub-basins)
Basement Tectonics and Shear Zones … 303
Fig. 13 (continued)
The target areas for hydrocarbon exploration in the Cauvery Basin are the areas of
overlap of (1) structural highs (2) areas of maximum fracture porosity overlying
shear zones and (3) source pods of hydrocarbon generation that either overlie these
regions or are connected with them by E-W faults formed by reactivations of shear
zones. The areas of structural highs are manifested on the surface as geomorphic
highs and are classified as per their degree of confidence. Geomorphic highs of high
and medium confidence in the ridge areas are used to represent structural highs.
Areas of positive correlation between all the elements in a GIS platform indicate the
promising areas of basement exploration in Cauvery Basin (Fig. 16a, b).
The structural highs are mostly products of the main rift tectonics of Late
Jurassic-Early Cretaceous that uplifted the basement with high structural relief.
Areas of basement, affected by shear zones are also uplifted in these basement highs
and form zones of high fracture porosity as described earlier. The post rift
Cretaceous shales in Portonovo, Kudavasal, Sattapadi and Synrift Andimadam
Formation form the potential source rocks (Singh et al. 2007) that are mapped as the
304 S. Mazumder et al.
Fig. 14 Areas of maximum fracture porosity delineated from Phanerozoic basement trends
associated with shear zones in the Cauvery Basin. Wells P-6, MT-9, M-5 and M-6 are hydrocarbon
bearing wells producing from basement in these areas
source pods. These are deposited in the lows adjacent to the basement highs or
directly overlie the basement highs. Shales within Kudavasal and Kamlapuram
Formation act as seals. On attaining critical moment, the hydrocarbon expulsed
migrate updip or migrate laterally along E-W reactivated faults and charge the areas
of fractured basement in the highs. In case of these source rocks directly overlying
the basement highs, downward migration resulting from hydrocarbon expelled
during compaction of the source sediments may also charge the reservoir areas
(Sircar 2004). The fractures serve as essential porosity for the hydrocarbon that may
move downdip along them and form accumulations (Fig. 16b).
Basement Tectonics and Shear Zones … 305
Fig. 15 a Drilling Induced Fractures (DIFs) interpreted from well M-5 indicating SHmax oriented
E-W to NW-SE. b Comparison of rose diagrams of surface morphostructural trends and faults
mapped at the basement top in Madanam field (Fig. 7) indicate that the E-W trends are the most
reactivated. c Adsorbed gas anomalies in onshore Cauvery Basin and natural hydrocarbon
seepages in offshore areas (Dave et al. 2012) associated with E-W oriented shear zones indicating
that the E-W trending faults act as conduits of migration/remigration under present stress regime
306 S. Mazumder et al.
Fig. 16 a Promising areas of basement exploration extracted from the spatial correlation of the
areas of maximum porosity in shear zones, geomorphic highs and source pods that spatially
overlap each other. Wells P-6, MT-9, M-5, M-6 and P-9 are hydrocarbon bearing wells producing
from basement in these areas. b Schematic model (not to scale) of (i) Initial basement rock with
probable NE-SW trending foliations (ii) Formation of E-W trending shear zones (in red lines)
associated with dextral shearing resulting in re-orientation of earlier foliations, formation of E-W
fractures at basement top and NE-SW oriented linkages connecting the shear zones (iii) Probable
formation of N-S trending faults resulting in reactivations of earlier fractures (iv) Main phase of
rifting due to NE-SW trending faults forming a number of basement highs and lows. Basement
highs associated with shear fractures and fault intersections act as promising areas c Block diagram
of conceptual model for basement exploration
Basement Tectonics and Shear Zones … 307
4 Conclusions
The study emphasizes on the use of tectonics and structures in basement exploration
in Cauvery Basin. A GIS based correlation of different surface and subsurface
geological & geophysical datasets using outcrop structural analogues model the
disposition of shear zones in the northern part of Cauvery Basin. The mapped shear
zones occur as individual bands of 250–400 m thick spaced apart by 500 m to a
few km and dipping moderately to steeply near the basement top thereby fracturing
extensively basement top areas. These shear zone fractures develop in an E-W
orientation parallel to the shear foliations that opened due to repeated reactivations.
Imprints of subsequent basement tectonism on these zones increases fracture
porosity and creates connectivity with different fracture sets thus enhancing the
reservoir character of the basement. Structural highs created during the main phase
of rifting and associated with fractured basement forms promising areas of base-
ment exploration. Neotectonic reactivations due to changes in in-situ stress regime
results in hydrocarbon charging into the basement reservoirs thus creating hydro-
carbon prospects. A similar procedure may be effective delineating basement
hydrocarbon in other similar basins that have undergone multiple episodes of
tectonism.
Acknowledgements The authors express their gratitude to Soumyajit Mukherjee (IIT Bombay)
for handing and reviewing this manuscript. The other anonymous reviewer is also thanked.
Mukherjee (2019) summarizes this work.
The views expressed by the authors in this manuscript are not necessarily similar to the views of
the organization they represent.
References
Bhutani R, Balakrishnan S, Nevin CG, Jeyabal S (2007) Sm–Nd isochron ages from Southern
Granulite Terrain, South India: Age of protolith and metamorphism. In: Goldschmidt
conference abstracts 2007, A89
Burg JP (2017a) Ductile Faults: Shear Zones. http://www.files.ethz.ch/structuralgeology/JPB/files/
English/13shearzones.pdf
Burg JP (2017b). Folds. http://www.files.ethz.ch/structuralgeology/JPB/files/English/8folds.pdf
Cao J, Liu S, He X (2011) Earthquake-induced secondary hydrocarbon migration and
accumulation in Longmanshan region, China. In: American Geophysical Union, fall meeting
2011, abstract #T42B-04
Chandrasekhar N, Mane PH, Rajappan P (2015) Basement hydrocarbon exploration-overview of
basement exploration. In: Multicomponent seismic analysis, identification of fractures and
status of Indian basin exploration. Geohorizons July, pp 36–46
Chardon D, Jayananda M, Chetty TRK, Peucat JJ (2008) Precambrian continental strain and shear
zone patterns: South Indian case. Journal of Geophysical Research 113, B08402
Chatterjee R, Mukhopadhyay M (2008) In-situ stress results from east coast sedimentary basins of
India. In: Oral session VI: stress field interpretation at regional scale, 3rd world stress map
conference 15–17. Potsdam
308 S. Mazumder et al.
Chetty TRK, Bhaskar Rao YJ (2006) The Cauvery Shear Zone, Southern Granulite Terrain, India:
A crustal-scale flower structure. Gondwana Res 10(1–2):77–85
Chetty TRK, Santosh M (2013) Proterozoic orogens in southern Peninsular India: contiguities and
complexities. Journal of Asian Earth Sciences 78, 39–53
Chetty TRK (2015) The Cauvery suture zone: map of structural architecture and recent advances.
Journal of Geological Society of India 85, 37–44
Chetty TRK, Bhaskar Rao YJ, Narayana BL (2003) A structural cross section along Krishnagiri–
Palani Corridor, Southern Granulite Terrain of India. Memoir Geological Society of India 50,
255–277
Chetty TRK, Yellappa T, Mohanty DP, Nagesh P, Sivappa VV, Santosh M, Tsunogae T (2012)
Mega sheath fold of the Mahadevi Hills, Cauvery Suture Zone, Southern India: implication for
accretionary tectonics. Journal Geological Society of India 80, 747–758
Chung HP (1982) Petroleum in basement rocks. American Association of Petroleum Geologists
66, 1597–1643
Dang CTQ, Chen Z, Nguyen NTB, Bae W, Phung TH (2011) Improved oil recovery for fractured
granite basement reservoirs: historical lessons, successful application, and possibility for
improvement. In: Society of petroleum engineers, SPE-144148-MS, SPE European formation
damage conference, 7–10 June, Noordwijk, The Netherlands
Darmadi Y, Harahap A, Achidat R, Ginanjar M, Hughes J (2013) Reservoir characterization of
fractured basement using seismic attributes, Dayung field case study, South Sumatra Indonesia.
In: Proceedings, Indonesian petroleum association 37th annual convention and exhibition, May
2013
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology 125, 561–591
Dave HD, Mazumder S, Samal JK, Pangtey KKS, Mitra DS (2012) Mapping Hydrocarbon
seepages using satellite SAR data in Eastern offshore—essential input in oil exploration, P-137.
In: Proceedings of 9th Biennial international conference and exposition on petroleum
geophysics, SPG Hyderabad, 2012
Desikachar SV, Ramanathan S, Babu PVLP (1960) Progress report of geological exploration of
Cauvery Basin. 1959–60. ONGC Unpublished report
DGH India (2015) Estimated resources of crude oil & natural gas, exp.about.resources2015.pdf.
https://www.ndrdgh.gov.in/NDR/?page_id=1251
EarthExplorer, USGS. https://earthexplorer.usgs.gov/
Gaina C, Müller RD, Brown B et al (2007) Breakup and early seafloor spreading between India
and Antarctica. Geophysical Journal International, 170, 151–169
Ganapathy GP, Rajarathinam (2010) Use of remote sensing and seismotectonic parameters to
identify seismogenic sources of Tamil Nadu State. International Journal of Applied
Engineering Research 1, 59–76
Gupta P, Rathore SS, Raza S, Uniyal GC (2015) Radiometric dating and textural characterization
of Basement rocks of East Coast, KG and Cauvery Basins, ONGC report, unpublished
Gutmanis J (2009) Basement reservoirs—a review of their geological and production character-
istics. In: International petroleum technology conference, IPTC 13156
Harinarayana T, Naganjaneyulu K, Patro BPK (2006) Detection of a collision zone in south Indian
shield region from magnetotelluric studies, Gondwana Research 10, 48–56
Heidbach O, Rajabi M, Reiter K, Ziegler M, Team WSM (2016) World stress map database
release 2016. GFZ Data Services. https://doi.org/10.5880/WSM.2016.001
Hua B (1995) Stress field, seismic pumping and oil-gas migration. Acta Sedimentological Sinica
13, 77–85
Huy XN, Bae W, San TN, Xuan VT, SungMin J, Kim DY (2012) Fractured basement reservoirs
and oil displacement mechanism in white tiger field. In: Offshore Vietnam, APG Search and
Discovery Article #90155©2012 AAPG international conference & exhibition, Singapore,
16–19 September
Janardhan AS (1999) Southern Granulite Terrain, south of the Palghat Cauvery Shear Zone:
implications for India Madagascar connection. Gondwana Research 2, 463–469
Basement Tectonics and Shear Zones … 309
Kumar N, Singh AP, Singh B, (2009) Structural fabric of the Southern Indian shield as defined by
gravity trends. Journal of Asian Earth Sciences 34, 577–585
Kumaran CJ, Ramasamy SM, (2005) Fluvial anomalies and Neotectonics of parts of Western
Ghats, Tamil Nadu, India. In: Ramaswami SM (ed) Remote Sensing in Geomorphology,
Chapter: Chapter-7, Publisher: New India Publishing Agency. pp 81–88
Lal NK, Siawal A, Kaul AK (2009) Evolution of east Coast of India—a plate tectonic
reconstruction. Journal of Geological Society of India 73, 249–260
Laws S, Eberhardt E, Loew S, Descoeudres F (2003) Geomechanical properties of shear zones in
the Eastern Aar Massif, Switzerland and their Implication on Tunnelling. Rock Mechanics and
Rock Engineering 36, 271–303
Mazumder S, Dave HD, Samal JK, Mitra DS (2011) Delineation of shallow structures as
exploratory input in Ganga Basin based on Morphotectonic studies. ONGC Bulletin 46,
190–198
Mazumder S, Pangtey KKS, Mitra DS (2013a) Delineation of a possible subsurface ridge in
Onshore Palar Basin based on morphotectonic studies and its implications. In: Proceedings of
10th Biennial international conference & exposition, SPG, Kochi
Mazumder S, Tep B, Pangtey KKS, Mitra DS (2013b) A Morphotectonic based approach to derive
tectonic framework and possible areas of hydrocarbon accumulation in onshore Palar Basin.
ONGC Bulletin 48, 113–123
Mazumder S, Adhikari K, Mitra DS, Mahapatra S, Pangtey KKS (2016) A Neotectonic based
geomorphic analysis using remote sensing data to delineate potential areas of hydrocarbon
exploration: Cachar Area, Assam. Journal Geological Society of India 88, 87–97
Meert JG, Pandit MK, Pradhan VR, Banks J, Sirianni R, Stroud M, Newstead B, Gifford J (2010)
Precambrian crustal evolution of Peninsular India: A 3.0 billion year odyssey, Journal of Asian
Earth Sciences 39, 483–515
Misra AA, Mukherjee S (2015) Tectonic Inheritance in Continental Rifts and Passive Margins.
Springerbriefs in Earth Sciences. ISBN 978-3-319-20576-2
Mukherjee S (2012) Tectonic implications and morphology of trapezoidal mica grains from the
Sutlej section of the Higher Himalayan Shear Zone, Indian Himalaya. The Journal of Geology
120, 575–590
Mukherjee S (2013) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy,
Berlin, pp 1–111. ISBN 978-3-642-25608-0
Mukherjee S (2014a) Atlas of shear zone structures in meso-scale. Springer Geology, Cham,
pp 1–124. ISBN 978-3-319-0088-6
Mukherjee S (2014b) Review of flanking structures in meso- and micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2015a) Atlas of structural geology. Elsevier, Amsterdam. ISBN: 978-0-12-420152-1
Mukherjee S (2015b) Petroleum geosciences: Indian contexts. Springer Geology. ISBN
978-3-319-03119-4
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Zanskar section-microstructural
studies & extrusion mechanism by a combination of simple shear & channel flow. International
Journal of Earth Sciences 99, 1083–1110
Mukherjee S, Koyi HA (2010b) Higher Himalayan Shear Zone, Sutlej section- structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Mukherjee S, Mulchrone KF (2015) Ductile Shear Zones: From Micro-to Macro-scales. Wiley
Blackwell. ISBN: 978-1-118-84496-0
Mukherjee S, Punekar J, Mahadani T, Mukherjee R (2015) A review on intrafolial folds and their
morphologies from the detachments of the western Indian Higher Himalaya. In: Mukherjee S,
Mulchrone KF (eds) Ductile shear zones: from micro- to macro-scales. Wiley
310 S. Mazumder et al.
Murty GPS, Subrahmaniyam AS, Murthy KSR, Sarma KVLNS (2002) Evidence of fault
reactivation off Pondicherry coast from marine geophysical data. Current Science 83,
1446–1449
Nagendra R, Reddy AN (2017) Major geologic events of the Cauvery Basin, India and their
correlation with global signatures a review. Journal of Palaeogeography 6, 69–83
Nemcok M, Sinha ST, Stuart CJ et al (2013) East Indian margin evolution and crustal architecture:
integration of deep reflection seismic interpretation and gravity modelling. In: Mohriak WU,
Danforth A, Post PJ et al (eds) Conjugate divergent margins. Special Publications 369, The
Geological Society, London, pp 477–496
Phaye DK, Nambiar MV, Srivastava DK (2011) Evaluation of petroleum systems of
Ariyalur-Pondicherry Sub-basin (Bhuvangiri area) of Cauvery Basin, India: a two dimensional
(2-D) Basin modeling study. In: AAPG search and discovery Article #90118
Plavsa D (2014) The tectonic evolution of the Southern Granulite Terrane of India and its role in
the amalgamation of Gondwana, Ph.D. Thesis, University of Adelaide
Prabaharan S, Subramani T, Manonmani R, Ramalingam M (2013) Satellite lineaments and Subtle
structures in Cauvery Basin, Tamil Nadu. International Journal of Remote Sensing &
Geoscience 2, 8–11
Prasad B, Phor L (2009) Palynostratigraphy of the subsurface Gondwana and post-Gondwana
Mesozoics of the Cauvery Basin, India. Journal of the Palaeontological Society of India 54,
41–71
Rajaram M, Anand SP (2014) Aeromagnetic signatures of Precambrian shield and suture zones of
Peninsular India. Geoscience Frontiers 5, 3–14
Rajendra Prasad B, Rao GK, Mall DM, Rao PK, Raju S, Reddy MS, Rao GSP, Sridher V,
Prasad ASSSRS (2007) Tectonic implications of seismic reflectivity pattern observed over the
Precambrian Southern Granulite Terrain, India. Precambrian Research 153, 1–10
Rangaraju MK, Agrawal A, Prabhakar KN (1993) Tectono-Stratigraphy, structural styles,
evolutionary model and hydrocarbon habitat, Cauvery and Palar basins. In: Proceedings of
second seminar on Petroliferous Basins of India, Vol I, pp 371–388
Rao VV, Prasad BR (2006) Structure and evolution of the Cauvery Shear Zone system,
SouthernGranulite Terrain, India: evidence from deep seismic and other geophysical studies.
Gondwana Research 10, 29–40
Rao VV, Sain K, Reddy PR, Mooney WD (2006) Crustal structure and tectonics of the northern
part of the Southern Granulite Terrane, India. Earth and Planetary Science Letters 251, 90–103
Ravindra Kumar GR (2005) Lithology and metamorphic evolution of granulite-facies segments of
Kerala, Southern India. Journal of Geological Society of India 66, 253–254
Reber JE, Dabrowski M, Galland O, Schmid DW (2013) Sheath fold morphology in simple shear.
Journal of Structural Geology 53, 15–26
Resmi MR, Achyuthan H, Jaiswal MK (2016) Middle to late Holocene paleochannels and
migration of the Palar River, Tamil Nadu: implications of neotectonic activity, Quaternary
International 443, 211–222
Santosh M, Maruyama S, Sato K (2009) Anatomy of a Cambrian suture in Gondwana: Pacific-type
orogeny in southern India? Gondwana Research 16, 321–34
Santosh M, Xiao WJ, Tsunogaed T, Chetty TRK, Yellappae T (2012) The Neoproterozoic
subduction complex in southern India: SIMS zircon U–Pb ages and implications for Gondwana
assembly. Precambrian Research 192–195, 190–208
Saravanavel J, Ramasamy SM (2003) Probable Seismo Tectonic Lineaments in South India. In:
Proceedings of annual convection of ISRS and national conference on remote sensing, CESS,
At Trivandrum
Satyanaryana P, Sinha PK, Gupta DK, Sathe AV, Katiyar GC (2010). Hydrocarbon prospectivity
of the Basement of Mumbai High Field, P-374. In: Proceedings of 8th Biennial international
conference and exposition on petroleum geophysics, SPG 2010, Hyderabad
Schrank CE, Handy MR, Fusseis F (2008) Multiscaling of shear zones and the evolution of the
brittle-to-viscous transition in continental crust. Journal of Geophysical Research 113, B01407
Basement Tectonics and Shear Zones … 311
Shaw CA, Karlstrom KE, McCoy A, Williams ML, Jercinovic MJ, Dueker K (2002) Proterozoic
Shear Zones in the Colorado Rocky Mountains: From Continental Assembly to
Intracontinental Reactivation. In: Lageson D (ed) Science at the highest level. Geological
Society of America Field Guide 3, 102–117
Sibson RH, Moore JMM, Rankin AH (1975) Seismic pumping—a hydrothermal fluid transport
mechanism. Journal of the Geological Society 131, 653–659
Singh H, Goswami BG, Pahari S, Prasad IVSV, Singh RR (2007) Petroleum Geochemistry of the
Cauvery Basin, India. In: AAPG Search and Discover Article #90063©2007 AAPG Annual
Convention, Long Beach, California
Singh Y, John B, Ganapathy GP, George A, Harisanth S, Divyalakshmi KS, Kesavan S (2016)
Geomorphic observations from southwestern terminus of Palghat Gap, south India and their
tectonic implications. Journal of Earth System Sciences 125, 821–839
Sircar A (2004) Hydrocarbon production from fractured basement formations. Current Science 87,
147–151
Subrahmanyam AS, Lakshminarayana S, Chandrasekhar DV, Murthy KSR, Rao TCS (1995)
Offshore Structural Trends from Magnetic data over Cauvery Basin, East Coast of India.
Journal of Geological Society of India 46, 269–273
Subrahmanyam DS (2014) Stress Provinces of India–contribution to world stress
map. International Journal of Advanced Earth Science and Engineering 3, 108–113
Trenchard S (2007) Geomorphology as applied to oil exploration in Northwest Kansas. http://
www.docstoc.com/docs/96655562/Geomorphology-as-Applied-to-Oil-Exploration-in-
Northwest-Kansas
Trice R (2014) Fractured basement reservoirs: a new play for the UK, Offshore Engineer. http://
www.oedigital.com/energy/item/5363-fractured-basement-reservoirs-a-new-p-lay-for-the-uk
Vairavan V (1993) Tectonic Hisyory and Hydrocarbon prospects of Palar and Pennar Basins,
India. In: Proceedings of second seminar on Petroliferous Basins of India Vol I, 308–396
Vasudevan K, Rao PH, Vairavan V (2012) Role of tectonics in development of fractured basement
reservoir in Mumbai High Field, Western Offshore Basin, India. In: 9th Biennial International
Conference and Exposition on Petroleum Geophysics, p 482
Veevers JJ (2009) Palinspastic (pre-rift and -drift) fit of India and conjugate Antarctica and
geological connections across the suture. Gondwana Research 16, 90–108
Zahran I, Askary S (1988) Basement reservoir in Zeit Bay oil field, Gulf of Suez. https://www.
researchgate.net/publication/236487721_Basement_reservoir_in_Zeit_Bay_oil_field_Gulf_of_
Suez
Zhang L, Luo X, Guy V, Yu C, Yang W, Lei Y, Song C, Yu L, Yan J (2011) Evaluation of
geological factors in characterizing fault connectivity during hydrocarbon migration:
Application to Bohai Bay basin, Marine and Petroleum Geology, 28, 1634–1647
Implication of Transfer Zones in Rift
Fault Propagation: Example
from Cauvery Basin, Indian East Coast
Swagato Dasgupta
1 Introduction
The Cauvery basin, a rift zone located in the southern part of the eastern India,
records a poly-phase rift-drift transition (Veevers and Tewari 1995; Veevers 2009;
Lal et al. 2009; Sinha et al. 2010; Nemčok et al. 2013; Mukherjee 2018). The basin
evolved by Late Jurassic-Early Cretaceous rifting of the Indian plate from the
Antarctic plate (Sastri et al. 1981; Veevers 2009; Sinha et al. 2010). It was a part of
the rift system that included Palar–Pennar, Krishna–Godavari and Mahanadi rift
zones, and dextral Coromondal, North Vizag and Konark transform faults (Veevers
2009; Sinha et al. 2010). The Cauvery rift zone evolved as a system of rift units
controlled by NE striking normal faults, having a trend similar to that of Krishna–
Godavari and Mahanadi rift zones. The Cauvery basin, in the southern part, initi-
ated as India separated from Sri Lanka due to India–Africa break–up (Katz 1978;
Chari Narasimha et al. 1995; Gaina et al. 2007; Veevers 2009). Thus the basin
experienced a poly phase rifting episode which ended by dextral strike-slip and
oblique-slip separation of Antarctic plate along the Coromondal transform margin
(Veevers 2009; Sinha et al. 2010; Nemčok et al. 2013). This *NNW trending
transform fault (*200–300 km length) comprise of horse-tail structure, which
overprints the previously developed normal fault pattern of the Cauvery offshore
basin in the southern part of India (Nemčok et al. 2013). The northern horse-tail
structure (*NNE-NE trend, *150–200 km length) of the NNW striking
Coromondal transform margin got linked with the Krishna–Godavari rift zone
(Sinha et al. 2010; Nemčok et al. 2013). Some small-scale pull-apart basins exist
along the Coromondal transfer zone, immediately N of the Cauvery basin
(Dasgupta 2018b).
S. Dasgupta (&)
301-Platinum Aura, Sector-20, Roadpali, Navi Mumbai 410218, India
e-mail: swagato.dg@gmail.com
Some authors document that the Cauvery basin is a failed arm rift between the
Peninsular India in NW and the Sri Lankan landmass towards SE with structural
trend sub-parallel to the adjacent Precambrian Eastern Ghat lineaments (Katz 1978;
Chari Narasimha et al. 1995; Lal et al. 2009; see Misra and Mukherjee 2015 on
inheritance issue). Precisely this basin, having different rift-drift transition history,
can be termed as a pericratonic rift basin considering its setting within the Indian
carton (Biswas et al. 1993).
Like other extensional rift settings, relay structures and transfer zones play
important roles in rift fault propagation in the Cauvery basin (Dasgupta and Maitra
2018; Dasgupta 2018a). In a rifted setting the strain transfer occurs along adjacent
faults. This transfer of displacement along a set of faults is normally gradual, a
typical characteristic of a relay structure or accommodation zone (Rosendahl 1987;
Morley et al. 1990; Morley 1999a). Pre-existing basement lineations further govern
the formation of transfer zones and segment the rift basin into different sub-basins
(Morley 1999b; Morley et al. 2004; Fossen and Rotevatn 2016; Misra and
Mukherjee 2015; Dasgupta and Mukherjee 2017).
The Cauvery basin is subdivided into a number of fault-controlled horst and graben
structures with sediment thickness ranging from 1 to >6 km (Chari Narasimha et al.
1995). The basement rock of the Cauvery basin comprise of Archean–Proterozoic
granite gneiss complex, a part of the Southern Granulite Terrain (Janardhan 1999;
Rajaram and Anand 2003; Rao et al. 2006; Mazumder et al. 2019). The major rift
faults orient NE. The *NW trending extension initiated the rifting between the
southern part of the Indian peninsular and Sri Lanka along pre-existing Precambrian
*E trending weak zones (Rao et al. 2006; Chetty 2015; Mazumder et al. 2019).
Further, extension along NW sub-divided the Cauvery basin into five sub–basins
corroborated by a number of NW trending transverse faults, which played a key
role in the basin evolution. Rift extension and subsequent anticlockwise rotation of
the peninsular India and Sri Lanka (Katz 1978; Chari Narasimha et al. 1995; Lal
et al. 2009) developed the Gulf of Mannar basin towards south of the Cauvery
basin.
In the onland portion and towards south, the major depressions are the Ariyalur–
Pondicherry sub-basin, the Tranquebar sub-basin, the Tanjavur/Tanjore sub-basin,
the Nagapattinam sub-basin and the Ramnad-Palk bay sub-basin (Fig. 1; Sastri
et al. 1979; Chari Narasimha et al. 1995; Murthy et al. 2008; Chaudhuri et al. 2010).
The Mandapam-Delft Ridge separates it in the south from the Gulf of Mannar
Basin. The onshore and offshore portions of the Cauvery basin are separated by
three major horsts, which include the Madanam/Portonovo Ridge in north, Karaikal
Ridge and the Vedaranniyam Terrace towards south (Fig. 1; Chari Narasimha et al.
1995; Murthy et al. 2008). These basement highs and sub-basins, shear zones
correlate well with the gravity anomaly map derived from the Bouguer gravity
Implication of Transfer Zones in Rift Fault Propagation … 315
Fig. 1 Tectonic map of the Cauvery basin. Basement highs and sub-basins associated with
various rift fault geometries shown that are overlaid on the present day physiographic map of
southern India. Modified from Ram Babu and Lakshmi (2004), Murthy et al. (2008), Radhakrishna
et al. (2012), Bastia and Radhakrishna (2012), Nemčok et al. (2013), Dasgupta and Maitra (2018)
anomaly and the free-air anomaly data (Verma et al. 1993; Ram Babu and Lakshmi
2004; Bastia and Radhakrishna 2012; Twinkle et al. 2016). The aeromagnetic map
derived from aeromagnetic survey (by Geological Survey of India), also depicts
similar type of basement morphology (Rajaram and Anand 2003). Interpretation of
subsurface seismic data from offshore area also reveal the *NE trending of rift
grabens and fault pattern of the Cauvery basin (Sinha et al. 2010; Nemčok et al.
2013; Dasgupta and Maitra. 2018; Misra and Dasgupta 2018).
The Ramnad and the Mannar sub-basin appear to have come into being in the
earliest stage of basin development (Katz 1978; Chari Narasimha et al. 1995).
During the early-mid rift stage, the northern grabens of the Cauvery basin e.g., the
Tranquebar, the Tanjavur, the Nagapattinam and the Ariyalur–Pondicherry
remained nascent due to shear and rifting that deepened the basin. These sub-basins
316 S. Dasgupta
probably started with the deposition of ill-sorted Early Cretaceous sediments at the
base. The oldest sediments of the Upper Gondwana expose in the western margin of
the Cauvery basin known as the Sivaganga bed/Therani Formation comprising of
coarse gritty conglomerates with occasionally argillaceous sandstones and shales
(Roy Moulik and Prasad 2007; Watkinson et al. 2007; Kale 2010). The oldest
unit of the sub-crop stratigraphy is of Jurassic age drilled in few sub-basins
(Chaudhuri et al. 2010).
This tectonic phase was followed by continued deepening of the grabens mainly
in response to the tensional forces and partly to thermal subsidence (Sag Phase: Roy
Moulik and Prasad 2007; Nagendra et al. 2011). A cumulative thickness of *3 km
of Lower Cretaceous sediments in the northern part of the block suggests a long
period of basin deepening followed by a differentiated sedimentation history (Raju
et al. 2005; Nagendra et al. 2011). During the Cenomanian–Turonian (*100–
89 Ma) as Madagascar was getting separated from the Indian west-coast, it tilted
the Cauvery basin towards SE (Chari Narasimha et al. 1995; Storey et al. 1995; Roy
Moulik and Prasad 2007; Reeves 2013). This generated a set of prograding chan-
nelized system in the deep-waters. Widespread marine sedimentation happened in
the Late Cretaceous (Rangaraju et al. 1993; Roy Moulik and Prasad 2007; Bastia
and Radhakrishna 2012). It essentially marks the development of the shelf-slope
system with the outer shelf sinking or tilting relatively faster (Nagendra et al. 2011).
Differential compaction (Mukherjee and Kumar submitted for mechanism) and fast
rate of sedimentation in the Late Cretaceous formed a polygonal fault system that
are mainly fluid escape features (Dasgupta and Misra 2018; also see Dasgupta and
Mukherjee submitted for general concepts).
Transfer zone terminologies were initially applied to explain the structures formed
in between the overlapping tips of the thrusts (Dahlstrom 1970). These terms were
used subsequently in extensional settings as well to define the zone of accommo-
dation between the two overlapping normal fault tips (Rosendahl 1987).
Accommodation zones are areas enclosed by terminating segments of overlapping
faults that dip in same or opposite direction (Rosendahl 1987; Faulds and Varga
1998). The amount of overlap between the normal faults controls the geometry of
the accommodation zone (Rosendahl 1987). The term ‘transfer zone’ or ‘relay
structure’ is preferred by many authors over accommodation zone in extensional
regimes (Morley et al. 1990; Gawthorpe and Hurst 1993; Morley 1999a; Fossen
2016; Fossen and Rotevatn 2016). These transfer zones preserve both the strain
amount as well as the elevation difference between the adjacent footwalls
(Morley et al. 1990).
Transfer faults have been defined as transverse to oblique faults with a strong
strike-slip component (Gibbs 1984). Strain transfer facilitates in extensional
regimes from one normal fault to another along these transfer faults (Gibbs 1984;
Implication of Transfer Zones in Rift Fault Propagation … 317
Fig. 2 Transfer zone nomenclature. Modified from Morley et al. (1990), Morley (1999a)
Morley et al. 1990). The dip of the normal faults on either side of a transfer fault are
generally in same direction, although opposite dips are also common (e.g., Morley
et al. 1990; Morley 1999a).
The classification of transfer zones by Morley et al. (1990; Fig. 2) is relatively
robust than that of other authors. It encompasses both extents of overlap as well as
changes in dip direction of the adjoining set of normal faults. The primary subdi-
vision is based on the direction of throw of the two overlapping normal faults;
(a) conjugate—boundary faults having opposite dips and (b) synthetic—boundary
faults with similar dip directions. Conjugate transfer zones can be sub-divided: (a1)
convergent—boundary faults dipping towards each other; and (a2) divergent—
boundary faults dipping oppositely. Transfer zone comprising of approaching
normal fault tips usually associate with transfer faults or transverse faults across
which the slip transfers. This nomenclature of transfer zone as proposed by Morley
et al. (1990) has been incorporated in the present work.
4 Discussions
In a typical rift system, the rate of occurrence and dip direction of secondary faults
might change on successive dip lines leading formation of different type of transfer
zones (Morley et al. 1990; Morley 1999a). This is well observed in the Cauvery
offshore basin (as in Dasgupta and Maitra 2018). Switch of polarity of the
rift-related faults separated the syn-rift grabens by basement highs (Morley et al.
1990; Withjack et al. 2002; Fossen 2016; Fossen and Rotevatn 2016). Variation in
dip of the secondary faults is complemented by change in polarity of the major
318 S. Dasgupta
faults. The polarity switch over zones of the secondary faults is normally associated
with convergent overlapping transfer zones (Morley et al. 1990). The transfer faults
generally originate at much later stage of rifting, even after the overlapping transfer
zones formed.
In onland and offshore areas of Cauvery basin, transfer faults segregate different
sub- basins (Chari Narasimha et al. 1995; Twinkle et al. 2016). The overall rift fault
trend is *NE from the onland to the offshore areas. However, the rose diagram plot
of the fault strike depicts that the rift fault trends has certain variations across the
different sub-basins (Fig. 3). Also, there is a major variation of the transfer or
transverse fault trend (Fig. 3). The WNW trend of the transfer faults is prevalent in
the Ramnad-Palk Bay, the Ariyalur–Pondicherry and the NE offshore (Tranquebar
offshore) sub-basins. While the NW trend is the dominant one in the Tanjavur, the
Nagapattinum and the Tranquebar sub-basins (Figs. 1 and 3). There is also some
good proportion of NW trending transfer faults in the Ariyalur–Pondicherry
sub-basin. Such variation of rift fault geometry is mainly due to transfer zones
guided by pre-existing Precambrian basement shear fabrics and basement rheology
(Withjack and Jamison 1986; Morley 1999b; Withjack et al. 2002; Morley et al.
2004; Robertson et al. 2016; Dasgupta and Mukherjee 2017; Dasgupta and
Mukherjee, in press). Few of the *E trending faults (Fig. 3) were probably affected
by the pre-existing basement lineations (Mazumder et al. 2019). These transfer
faults also alter the basin depo-centers in the onland and the offshore parts.
The Ariyalur–Pondicherry, the Tranquebar and the Nagapattinam sub-basins
extend into the deep-water areas towards NE. The Nagapattinam sub-basin termi-
nates against the Karaikal high in NE (Fig. 1). The Portonovo high dies out towards
NE through a set of relay faults associated with some *NW trending transfer faults.
This provides the sediment entry points to the deeper basins. These transfer faults
probably associate with strike slip or oblique slip motion (e.g. Morley 1999a, b;
Morley et al. 2004). The basin opening as well as deepening is influenced by the
transfer zones guided rift faults. In each of the sub-basins, the *NW trending
extension separates the basement highs. Such separation commonly associate with a
set of secondary fault systems across which slip transfers through different transfer
zones (Figs. 4 and 5). Figure 4 explains this as a cartoon, which depicts that the
basin opens towards NE as explained by the dip lines 1–3 (Fig. 5). Transfer zones
can be easily identified and named in Fig. 4 (highlighted by yellow; also see Fig. 2).
The major Mesozoic discoveries and producing fields of the basin are in vicinity
to the basement highs where the transfer zones are present to a great extent (Fig. 1;
Chakraborty et al. 2011; www.dghindia.org). This indicates that the transfer zones
act as good fluid migration pathway as well as create hydrocarbon entrapments.
Near the basin margin area the rock physical parameters indicate a mixed type
lithology with considerable variation of porosity and elastic properties (Chatterjee
et al. 2013). This is obvious as the sediments undergo short distance of transport,
dominated by transverse drainage, thereby having poor textural and mineralogical
maturity (Ravnås and Steel 1998). Reservoir quality is a key concern in such areas.
Thus in order to get a better reservoir quality (of Early Cretaceous units) and
considerable hydrocarbon pool, the relatively basin interior areas over/adjacent to
Implication of Transfer Zones in Rift Fault Propagation … 319
Fig. 3 Rose diagram plot of rift fault trends of different sub basins in Cauvery basin. Note the
variation among the trends of the rift faults as well as the transfer faults across different sub basins.
Data gathered from Murthy et al. (2008), Nemčok et al. (2013), Dasgupta and Maitra (2018)
320 S. Dasgupta
Fig. 4 Tectonic map (cartoon) of a rift system, depicting a typical basin opening process due to
extension guided by a set of transfer zones
the transfer zones are potential targets. Such areas possess good structural entrap-
ment guided by the transfer zones. Also, prominent sediment fairway leads to better
reservoir quality.
The crustal architecture varies from the proximal to the distal margin of the
basin. At the proximal margin, the sediment accommodation space in the
half-grabens is controlled by high-angle faults having steeply inclined shear
(Fig. 6). Thus, the syn-rift sediments can be quite thick (e.g., sedimentary fills of
the Ariyalur-Pondicherry and the Ramnad sub-basins). In these areas, especially the
Nagapattinum, the Tranquebar and the Ramnad–Palk Bay sub basins, rift faults
interacted with the transfer faults producing stress locking geometries leading to
inverted structures (Fig. 6; Murthy et al. 2008; Chakraborty et al. 2011). At the
distal margin, the rifting geometries are relatively younger and are governed by the
low-angle detachments (Nemčok et al. 2013; Misra and Dasgupta 2018). The dip of
the normal rift faults becomes more listric towards the deep water offshore areas
(Fig. 7; Mukherjee and Agarwal, in press, for listric fault kinematics). Differential
stretching of the continental crust from the proximal to the distal margin may thin
the crust and exhume partially the the continental lithosphere (Sinha et al. 2010;
Nemčok et al. 2013).
Implication of Transfer Zones in Rift Fault Propagation … 321
Fig. 5 Schematic dip sections of three lines from SW to NE (shown in Fig. 4) in the direction of
basin opening
322 S. Dasgupta
Fig. 7 Around 30 km long seismo-geological cross-section from offshore deep-water area of the
Cauvery basin adjacent to the outboard part of the Karaikal high. Note the low-angle listric faults
with typical syn-sedimentary growth features. Modified from Dasgupta and Maitra (2018)
Implication of Transfer Zones in Rift Fault Propagation … 323
The vertical movements of rifted fault blocks and isostatic uplifts (Mukherjee
2017 for mechanisms) in consequence to continental break-up (Nemčok et al. 2013;
Dasgupta and Mukherjee 2017) control the distribution of trap, reservoir and source
rock elements. Thus, understanding the driving dynamics plays an important role in
petroleum system evaluation at rifted and sheared continental margin segments like
those of the Cauvery basin.
5 Conclusions
It is observed that the basin expands towards NE associated with change in dip of
basin bounding faults. Rift fault propagation is guided by the transfer zones.
Switching of polarity of the basin-bounding rift faults link with the transfer zone
geometry. The transfer faults present in the basin are possibly influenced by the
older basement lineaments. This is well observed in the Cauvery rift basin where
the Southern Granulite Terrain is the basement rock. The transfer faults occur at
late-stage of rifting. In a typical rift system, transfer zone and transfer faults
associate with changes in slip form one rift fault to another. These act as major
sediment entry as well as barrier zones. Thus a variation in the transfer zone
geometry can affect the sediment depositional system and the fluid migration
pathway. The transfer zone also plays a key role in forming structural- and strati-
graphic traps.
Ackowledgements SDG thanks Soumyajit Mukherjee for handling this article and provide two
rounds of review comments. SDG thanks Abhimanyu Maitra for helping in preparation of Fig. 5 in
this article. The work is solely based on published literature survey. No data has been taken from
any companies/organizations. Mukherjee (2019) summarized this work.
References
Bastia R, Radhakrishna M (2012) Basin evolution and petroleum prospectivity of the continental
margins of India, vol 59. Elsevier, p 11–200. ISBN: 978-0-444-53604-4
Biswas SK, Bhasin AL, Ram J (1993) Classification of Indian sedimentary basins in the
framework of plate tectonics. In: Biswas SK, Pomdery J, Dave A, Maithani A, Garg P,
Thomas NJ (eds) Proceedings of the 2nd seminar on petroliferous basins of India, vol 1. Indian
Petroleum Publishers, Dehradun, p 1–46
Chakraborty C, Srivastava DK, Rana MS, Saxena VM, Sati GC (2011) Paradigm shift in
exploration strategy: identification of prospective corridors in Ramnad-Palk Bay sub basin,
Cauvery basin, India. In: The 2nd South Asian geoscience conference and exhibition, GEO
India
Chari Narasimha MV, Sahu JN, Banerjee B, Zutshi PL, Chandra K (1995) Evolution of the
Cauvery basin. India from subsidence modelling, Marine and Petroleum Geology 12, 667–675
Chatterjee R, Manoharan K, Mukhopadhyay M (2013) Petrophysical and mechanical properties of
cretaceous sedimentary rocks of Cauvery basin, eastern continental margin of India. Journal of
the Indian Geophysical Union 17, 349–359
324 S. Dasgupta
Chaudhuri A, Rao MV, Dobriyal JP, Saha GC, Chidambaram L, Mehta AK, Ramana LV,
Murthy KS (2010) Prospectivity of Cauvery Basin in deep syn-rift sequences, SE India; Search
and Discovery Article #10232, Adapted from oral presentation at AAPG Annual Convention
and Exhibition, Denver, Colorado, USA, June 7–10
Chetty TRK (2015) The Cauvery suture zone: map of structural architecture and recent advances.
Journal of Geological Society of India 85, 37–44
Dahlstrom CDA (1970) Structural geology in the eastern margin of the Canadian Rocky
Mountains. Bulletin of Canadian Petroleum Geology 18, 332–406
Dasgupta S (2018a) Sedimentary deformation features produced by differential compaction and
buoyancy in the offshore Palar Basin, East Coast of India (Chapter 24). In: Misra AA,
Mukherjee S (eds) Atlas of structural geological interpretation from seismic images. Wiley
Blackwell, Hoboken, pp 131–133. ISBN: 978-1-119-15832-5
Dasgupta S (2018b) Pull-apart basin in the offshore Cauvery–Palar Basin, India (Chapter 23). In:
Misra AA, Mukherjee S (eds) Atlas of structural geological interpretation from seismic images.
Wiley Blackwell, Hoboken, pp 127–129. ISBN: 978-1-119-15832-5
Dasgupta S, Maitra A (2018) Transfer zone geometry in the offshore Cauvery Basin, India. In:
Misra AA, Mukherjee S (eds) Atlas of structural geological interpretation from seismic images.
Wiley Blackwell, Hoboken, pp 117–120. ISBN: 978-1-119-15832-5
Dasgupta S, Misra AA (2018) Polygonal fault system in late cretaceous sediments of Cauvery
deepwater Basin, India. In: Misra AA, Mukherjee S (eds) Atlas of structural geological
interpretation from seismic images. Wiley Blackwell, Hoboken, pp 113–116. ISBN:
978-1-119-15832-5
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology 125, 561–591
Dasgupta S, Mukherjee S (in press) Remote sensing in lineament identification: examples from
western India. In: Billi A, Fagereng A (eds) Problems and solutions in structural geology and
tectonics. Developments in structural geology and tectonics book series. Series Editor:
Mukherjee S, Elsevier ISSN: 2542-9000
Dasgupta T, Mukherjee S (Submitted) Sediment compaction and applications in petroleum
geoscience. Springer. Series: Advances in oil and gas exploration & production. Series Editor:
Swenner R. ISSN: 2509-372X
Faulds JE, Varga RJ (1998) The role of accommodation zones and transfer zones in the regional
segmentation of extended terranes. Geological Society of America, Special Paper, p 323
Fossen H (2016) Structural geology, 2nd edn. Cambridge University Press, Cambridge. ISBN:
9781107057647
Fossen H, Rotevatn A (2016) Fault linkage and relay structures in extensional settings—a review.
Earth-Science Reviews 154, 14–28
Gaina C, Müller RD, Brown B, Ishihara T, Ivanov S (2007) Breakup and early seafloor spreading
between India and Antarctica. Geophysical Journal International 170, 151–169
Gawthorpe RL, Hurst JM (1993) Transfer zones in extensional basins: their structural style and
influence on drainage development and stratigraphy. Geological Society of London 150, 1137–
1152
Gibbs AD (1984) Structural evolution of extensional basin margins. Geological Society of London
141, 609–620
Janardhan AS (1999) Southern granulite terrain, south of the Palghat-Cauvery shear zone:
implications for India-Madagascar connection. Gondwana Research 2, 463–469
Kale A (2010) Comments on ‘Sequence surfaces and paleobathymetric trends in Albian to
Maastrichtian sediments of Ariyalur area, Cauvery Basin, India’ from Nagendra, Kannan, Sen,
Gilbert, Bakkiaraj, Reddy, and Jaiprakash. Mar Pet Geol 28:1252–1259
Katz MB (1978) Sri Lanka in Gondwanaland and the evolution of the Indian Ocean. Geological
Magazine 115, 237–244
Lal NK, Siwal A, Kaul AK (2009) Evolution of East Coast of India—a plate tectonic
reconstruction. Journal Geological Society of India 73, 249–260
Implication of Transfer Zones in Rift Fault Propagation … 325
Mazumder S, Tep B, Pangtey KKS et al (2019) Basement tectonics and shear zones in Cauvery
basin: implications in hydrocarbon exploration. In: Mukherjee S (ed) Tectonics and structural
geology: Indian context. Springer International Publishing AG, Cham, pp 279–311. ISBN
978-3-319-99340-9
Misra AA, Dasgupta S (2018) Shallow detachment along a transform margin in Cauvery Basin,
India. In: Misra AA, Mukherjee S (eds) Atlas of structural geological interpretation from
seismic images. Wiley Blackwell, Hoboken, pp 101–105. ISBN: 978-1-119-15832-5
Misra AA, Mukherjee S (2015) Tectonic inheritance in continental rifts and passive margins.
Springer briefs in Earth Sciences, Berlin. ISBN 978-3-319-20576-2
Morley CK (1999a) Aspects of transfer zone geometry and evolution in East African rifts. In:
Morley CK (ed) Geoscience of rift systems—evolution of East Africa, AAPG Studies in
Geology No. 44, p 161–171
Morley CK (1999b) How successful are analogue models in addressing the influence of
pre-existing fabrics on rift structure? Journal of Structural Geology 21, 1267–1274
Morley CK, Nelson RA, Patton TL, Munn SG (1990) Transfer zones in the East African rift
systems and their relevance to hydrocarbon exploration in rifts. AAPG Bulletin 74, 1234–1253
Morley CK, Haranya C, Phoosongsee W, Pongwappe S, Kornsawan A, Wonganan N (2004)
Activation of rift oblique and rift parallel pre-existing fabrics during extension and their effect
on deformation style: examples from the rifts of Thailand. Journal of Structural Geology 26,
1803–1829
Mukherjee S (2017) Airy’s isostatic model: a proposal for a realistic case. Arabian Journal of
Geosciences 10, 268
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Agarwal I (in press) Shear heat model for gouge free dip-slip listric normal faults.
Marine and petroleum geology. https://doi.org/10.1016/j.marpetgeo.2018.09.004
Murthy KS, Chaudhuri A, Ramana LV, Rao MV, Dobriyal JP (2008) Hydrocarbon exploration of
syn rift sediments in Nagapattinam Sub Basin, Cauvery Basin—a case study. In: 7th Biennial
international conference and exposition on petroleum geophysics, society of petroleum
geophysicists (SPG), p 443
Nagendra R, Kamalak Kannan BV, Sen G, Gilbert H, Bakkiaraj D, Reddy AN, Jaiprakash BC
(2011) Sequence surfaces and paleobathymetric trends in Albian to Maastrichtian sediments of
Ariyalur area, Cauvery Basin, India. Marine and Petroleum Geology 28, 895–905
Nemčok M, Sinha ST, Stuart CJ et al (2013) East Indian margin evolution and crustal architecture:
integration of deep reflection seismic interpretation and gravity modelling. In: Mohriak WU,
Danforth A, Post PJ et al (eds) Conjugate divergent margins, special publications, vol 369.
Geological Society London, pp 477–496
Radhakrishna M, Twinkle D, Nayak S, Bastia R, Rao S (2012) Crystal structure and rift architecture
across the Krishna–Godavari basin in the central Eastern Continental Margin of India based on
analysis of gravity and seismic data. Marine and Petroleum Geology 37, 129–146
Rajaram M, Anand SP (2003) Crustal structure of south India from aeromagnetic data. Journal of
the Virtual Explorer 12, 72–82
Raju DSN, Ramesh P, Mohan SKG, Uppal S (2005) Sequence and bio-chronostratigraphic
subdivisions of the Cretaceous and Cenozoic of India with notes on pre-Cretaceous events: an
overview. In: Raju DSN, Peters J, Shankar R, Kumar G (eds) Association of petroleum
geologists special publication No. 1, pp 97–103
Ram Babu HV, Lakshmi MP (2004) A reappraisal of the structure and tectonics of the Cauvery
Basin (India) from aeromagnetics and gravity. In: 5th conference & exposition on petroleum
geophysics. SPG Hyderabad, India, pp 19–23
Rangaraju MK, Aggarwal A, Prabhakar KN (1993) Tectono-stratigraphy, structural styles,
evolutionary model and hydrocarbon prospects of Cauvery and Palar Basins, India. In:
Biswas SK, Alak D, Garg P, Pandey J, Maithani A, Thomas NJ (eds) Proceedings of the 2nd
326 S. Dasgupta
seminar on petroliferous basins of India, vol 1. Indian Petroleum Publishers, Dehradun, India,
pp 331–354
Rao VV, Sain K, Reddy PR, Mooney WD (2006) Crustal structure and tectonics of the northern
part of the Southern Granulite Terrane, India. Earth and Planetary Science Letters 251, 90–103
Ravnås R, Steel RJ (1998) Architecture of marine rift-basin successions. AAPG Bulletin 82(1),
110–146
Reeves CV (2013) The global tectonics of the Indian Ocean and its relevance to India’s western
margin. J Geophys 34:87–94
Robertson EAM, Biggs J, Cashman KV, Floyd MA, Vye-Brown C (2016) Influence of regional
tectonics and pre-existing structures on the formation of elliptical calderas in the Kenyan Rift.
In: Wright TJ, Ayele A, Ferguson DJ, Kidane T, Vye-Brown C (eds) Magmatic rifting and
active volcanism, vol 420. Geological Society of London Special Publications, pp 43–67
Rosendahl BR (1987) Architecture of continental rifts with special reference to East Africa.
Annual review of Earth and Planetory Science Letters 15, 445–504
Roy Moulik SK, Prasad GK (2007) Seismic expression of the canyon fill facies and its geological
significance—a case study from Ariyalur-Pondicherry Sub-basin, Cauvery Basin, India. AAPG
Annual Convention, Long Beach, California
Sastri VV, Raju DSN, Venkatachala BS, Acharyya SK (1979) Sedimentary basin map of India,
stratigraphic correlation between sedimentary basins of the ESCAP region. Mineral Resource
Development Series, United Nations, New York, p 45
Sastri VV, Venkatachala BS, Narayanan V (1981) The evolution of the east coast of India.
Palaeogeography, Palaeoclimatology, Palaeoecology 36, 23–54
Sinha ST, Nemčok M, Choudhuri M, Misra AA, Sharma SP, Sinha N, Venkatraman S (2010) The
crustal architecture and continental break up of East India Passive Margin: an integrated study
of deep reflection seismic interpretation and gravity modeling; Search and Discovery Article
40611
Storey M, Mahoney JJ, Saunders AD, Duncan RA, Kelley SP, Coffin MF (1995) Timing of
hot-spot related volcanism and the breakup of Madagascar and India. Sci 267:852–855
Twinkle D, Rao GS, Radhakrishna M, Murthy KSR (2016) Crustal structure and rift tectonics
across the Cauvery–Palar basin, eastern continental margin of India based on seismic and
potential field modelling. Journal of Earth System Science 125, 329–342
Veevers JJ (2009) Palinspastic (pre-rift and -drift) fit of India and conjugate Antarctica and
geological connections across the suture. Gondwana Research 16, 90–108
Veevers JJ, Tewari RC (1995) Gondwana master basin of peninsular india between Tethys and the
interior of the Gondwanaland Province of Pangea. Geological Society of America Memoir,
pp 72–187
Verma RK, Rao SC, Satyanarayana Y (1993) Gravity field and evolution of Krishna–Godavari and
Cauvery basins of India. Indian Journal of Petroleum Geology 2, 39–52
Watkinson MP, Hart MB, Joshi A (2007) Cretaceous tectonostratigraphy and the development of
the Cauvery Basin, Southeast India. Petroleum Geoscience 13, 181–191
Withjack MO, Jamison WR (1986) Deformation produced by oblique rifting. Tectonophysics 126,
99–124
Withjack MO, Schlische RW, Olsen PE (2002) Rift-basin structure and its influence on
sedimentary systems. In: Renaut RW, Ashley GM (eds) Sedimentation of continental rifts.
SEPM Special Publication 73, pp 57–81
Remote Sensing, Structural and Rock
Magnetic Analyses of the Ramgarh
Structure of SE Rajasthan, Central
India-Further Clues to Its Impact Origin
and Time of Genesis
1 Introduction
The Ramgarh structure (centered at 25° 20′N, 76° 38′E) of SE Rajasthan, central
India (Figs. 1 and 2a), is a ring-shaped structure of an outer diameter of 4 km,
and is situated within the Vindhyan Supergroup of sedimentary rocks of Meso- to
Neoproterozoic age (Sharma 1973; Prasad 1984; Ramasamy 1987). Although, it is
a potential candidate of asteroid impact crater for last many decades (e.g., Crawford
1972; Grieve et al. 1988; Sisodia et al. 2006a, b), the supporting evidences in favour
of this idea are slowly accumulating. A re-investigation of its origin is thus
important since it is the only known potential candidate for asteroid impact crater in
India excavated in the sedimentary target rocks. The other comparable terrestrial
asteroid impact craters on sedimentary target-rocks of equivalent diameter are:
B. P. structure, Libya (25° 19′N, 24° 20′E); Goyder Crater, northern Australia
(13° 9′S, 135° 2′E), and Gusev Crater, Russia (48° 26′N, 40° 32′E) (Earth Impact
database, http://www.unb.ca/passc/ImpactDatabase).
The Ramgarh structure has many of the physical attributes of an impact crater
with a raised rim (Fig. 2a), as described below, but its origin has been debated and
more than one hypotheses have been proposed. Sharma (1973) first suggested that
S. Misra (&)
Discipline of Geological Sciences, SAEES, University of KwaZulu-Natal,
Durban 4000, South Africa
e-mail: misras@ukzn.ac.za
P. K. Srivastava
Department of Petroleum Engineering and Earth Sciences,
University of Petroleum and Energy Studies, Dehradun 248007, India
Md. Arif
Birbal Sahni Institute of Palaeosciences, Lucknow 226007, India
Fig. 1 Sketch map of a India, and b Vindhyan basin (in light grey shade), central India, showing
the position of the Ramgarh structure
Remote Sensing, Structural and Rock Magnetic Analyses … 329
this structure could have been resulted due to magmatic activity. However, this
hypothesis can safely be discarded because no igneous plug/activity has been
described within this structure, including in borehole studies reaching to *450 m
depth and in surrounding areas (Ramasamy 1987). Furthermore, the regional
aeromagnetic anomaly map does not support the presence of any magnetic high
below the structure (Kumar et al. 2011) suggestive of any magmatism.
Alternatively, it has been suggested that the Ramgarh structure formed due to
tectonic activity by ‘rheid’ flow of shales and associated horizontal compression in
subordinate scale (Ramasamy 1987). But there is no evidence of structures dictated
by compressive forces in the bed rocks surrounding the Ramgarh structure. This
model, therefore, fails to explain the origin of this unique structure within the
undeformed, flat-lying sedimentary country rocks.
The third upcoming hypothesis is that this structure could have been formed by
asteroid impact. Some of the important observations favouring the impact
hypothesis are (a) presence of a prominent conical central peak (Grieve et al. 1988;
Master and Pandit 1999), (b) shatter-cone like structure in colluvium near its center
(Crawford 1972); (c) presence of closed spaced Planar Fractures (PF) and granu-
lation in quartz grains in shocked sandstones from this structure (Balasundaram and
Dube 1973; Das et al. 2011), and (d) occurrences of impact spherule-looking
materials, diaplectic glasses, and possible Planer Deformation Fracture (PDF) in
quartz grains from this structure (Sisodia et al. 2006a, b; Purohot and Sisodia 2013;
Rana and Agrawal 2016), although the true nature of these PDFs remain ques-
tionable (Reimold et al. 2006; also cf. Vernooij and Langenhorst 2005; French and
Koeberl 2010).
The iron-silica rich mm-sized particles/spherules from within the alluvium inside
the Ramgarh structure have been primarily investigated in Misra et al. (2008b). The
Back Scattered Electron (BSE) images of these magnetic particles are similar to
impact-induced accretionary lapilli reported from other established impact struc-
tures (cf. Graup 1981; Koeberl et al. 2007; Newsom et al. 2016). These particles/
spherules have cores consisting of coarse grained quartz fragments surrounded by
multiple concentric layers of fine-grained materials with smaller quartz fragments
(Misra et al. 2008b; Das et al. 2011). Our studies further suggest that these
iron-silica rich particles have very high Natural Remnant Magnetization
(NRM) (*2–19 Am−1) and high REM (after Yu 2006 and references therein)
(*7–145%) that could indicate the presence of a high magnetic field during their
formation, much higher than the ambient Earth magnetic field (*40 Am−1) (Das
et al. 2009). The exact reason of this local high magnetic field in and around the
Ramgarh structure is not clearly known, but it could be resulted by asteroid impact
(cf. Crawford and Schultz 1988, 1999). More recently, a few fragments of iron-rich
Ca–Al silicate glassy pieces, recovered from the soil inside the Ramgarh structure
and nearby outside the western rim of the structure, contain dendritic magnetite
with occasional inclusions of relict native iron containing high proportions of Co
(*350–3000 ppm), Ni (*200–4000 ppm) and Cu (*2200–7000 ppm) (with
similar Co/Ni ratios), and these iron globules are thought to be extra-terrestrial in
origin (Misra et al. 2013).
Remote Sensing, Structural and Rock Magnetic Analyses … 331
Hence, the emerging geological evidences favour more towards the asteroid
impact origin of the Ramgarh structure. The only lacuna is that the bedrock
sandstones from the rim of this structure do not show any confirming PDF under the
transmitted light polarizing microscope satisfying the definition described in French
and Koeberl (2010). However, the general absence of PDF in quartz grains within
the Ramgarh structure sandstones does not preclude the asteroid impact origin of
this structure. This is because this microstructure could also be formed by terrestrial
lightning strikes (Chen et al. 2017; Melosh 2017), and hence cannot be considered
as a confirming criterion for identification of an impact crater. On the other hand,
although some geochemical clues on possible impact origin of the Ramgarh
structure have been documented (Ahmed et al. 1974; Sisodia and Lashkari 2003;
Sisodia et al. 2006a; Misra et al. 2008b, 2013), no detail information on the
structural geology and/or rock magnetic properties on this structure in favour or
against of its asteroid impact origin presently exist. In the present study, we thus
examine the available remote sensing images, and structural data of the Ramgarh
structure in more detail (cf. Misra et al. 2008a). The rock magnetic properties of a
few glassy samples from this structure (cf. Das et al. 2009) have also been
examined to re-evaluate the asteroid impact origin of the Ramgarh structure.
The Vindhyan Supergroup, into which the Ramgarh structure is situated, exposes as a
sickle-shaped, east-west elongated outcrop in the central India (Oldham 1856; Mallet
1869; Auden 1933), and occupies *178,000 km2 (Tandon et al. 1991) with a
thickness up to *4.5 km (Ahmad 1958) (Fig. 1). This Supergroup of rocks are
unmetamorphosed and only mildly deformed. Geophysical investigations revealed
that the rifting across E-W faults initiated the formation of the Vindhyan basin (Bose
et al. 2001). The N-S rifting was accompanied by a dextral shear and as a result the
basin segmented into several NW-SE elongated sub-basins. Between the two major
sectors of the Vindhyan Supergroup, the Ramgarh structure occurs in the western
(Rajasthan) sector of this huge sedimentary outcrop (Ramasamy 1987). This structure
is excavated in the flat-lying sandstone, shale, and minor limestone horizons of the
Lower Bhander Group of the Vindhyan Supergroup (Prasad 1984; Ramasamy 1987).
The oldest to youngest stratigraphic sequences of the Lower Bhander Group that are
exposed within the Ramgarh structure include the Ganurgarh Shale (thickness
*630 m), Lower Bhander Limestone (*26 m), Samaria Shale (*52 m) and Lower
Bhander sandstone (*400 m) (cf. Fig. 3 in Ramasamy 1987).
The ring-like, prominently rectangular shaped Ramgarh structure has a
rim-to-rim diameter of *2.37 and 2.48 km along E-W and N-S respectively
(measured from Google Earth Imagery 2017), and a more or less continuous raised
rim along its periphery except to the southwest, which rises *200 m above the
surrounding undisturbed, flat-lying plain (cf. Sisodia et al. 2006a; Misra et al.
2008b) (Fig. 2a). Earlier reports indicate the presence of a central conical peak
332 S. Misra et al.
within this structure (Grieve et al. 1988; Master and Pandit 1999; Sisodia et al.
2006a). The Upper Bhander sandstone, constituting the youngest Formation of the
Vindhyan Supergroup, has been dated as 1000–1070 Ma (U-Pb zircon age, Malone
et al 2008), which could be the upper age limit of formation of the Ramgarh
structure.
The sandstone exposures on the rim of the Ramgarh structure have dips between
*45° and 78°, which gradually reduce to *20°–30° at the lower outer slope, and
to *5°–15° at the foot of the outer rim (Prasad 1984). The rim of the structure is
only discontinuous at the SW corner and displaced by a set of NE-SW dextral strike
slip faults (Prasad 1984; Ramasamy 1987; Sisodia et al. 2006a; Misra et al. 2008a).
Additional joint/fault systems that cross-cut the Ramgarh structure trend E-W, N-S
and NW-SE (Ramasamy 1987; Misra et al. 2008a). Geomorphologically, the
Ramgarh structure appears to have more degraded state (i.e. modified by erosion) as
compared to those of the Arizona Crater, USA, and Lonar Crater, India (Grant
1999). This conclusion is well supported by the presence of reworked materials (i.e.
materials derived during weathering of the rim crest rocks of the structure) within a
very limited zone (*770 m) in the outer flank of the structure rim (Misra et al.
2008a). Although the Ramgarh structure is eroded, the sandstone exposed on its rim
shows quaquaversal dips (Sharma 1973; Ramasamy 1987) like other impact craters
(e.g., Lonar Crater, Misra et al. 2010).
3 Analytical Techniques
The rock magnetic characterization experiments were carried out on a dark reddish
coloured, glassy silicate sample (RJ-26) of a maximum size of 3 cm (Fig. 9 in
Appendix), which was recovered from within the rocky debris derived from the rim
crest rocks and deposited on the outer slope of the western rim of the Ramgarh
structure (Fig. 3). The glassy piece was magnetic, and under transmitted light,
polarizing, optical microscope it was looking similar to the glassy sample R-82
reported in Misra et al. (2013).
Our experimental studies include measurements of Natural Remanent
Magnetization (NRM), REM (=NRM/SIRM expressed in %, where SIRM—
Saturation Isothermal Remanent Magnetization) (after Yu 2006 and references
therein), Isothermal Remanent Magnetization (IRM) acquisition and backfield
SIRM dc demagnetization, and Lowrie-Fuller test (methods and techniques of our
experiments after McElhinny and McFadden 2000). All these measurements were
carried out at the Palaeomagnetism Laboratory, Birbal Sahni Institute of
Palaeosciences (Lucknow, India). Four mm-sized sub-samples were cut from the
original glassy silicate piece for our present measurements. The NRM was mea-
sured by an Agico JR-6 spinner magnetometer. The IRM was imparted in pro-
gressively increasing magnetic fields up to 1 T and their back field application to
the saturation IRM for evaluating remanence coercive force (Hcr) by an ASC
scientific IM-10–30 impulse magnetizer; magnetic remanence after each IRM step
Remote Sensing, Structural and Rock Magnetic Analyses … 333
Fig. 3 The panchromatic band Landsat-7 grey shed image of Ramgarh structure (source GLCF
2007) showing three prominent sets of lineament transecting the structure’s rim, abbreviations:
S.R.-structure’s rim, Kul- Kul River, 1, 2, 3-lineaments cross-cutting rim of Ramgarh structure,
NE-SW dextral slip and E-W sinistral slip shown by arrows. The rim and adjoining area of
Ramgarh structure are segmented into twelve geographic sectors, average orientations of
sedimentary country rocks (sandstones) on structure’s rim in each sector shown for reference (read
as follows: strike in degree/dip amount in degree, right-hand rule should be followed to interpret
dip direction), location of sample RJ 26 also shown. Black arrows at north eastern part of the map
indicating displacement of Kul River channel along NE-SW lineament
4 Analyses
The sedimentary country rocks surrounding the Ramgarh structure are essentially
flat (Fig. 4a). Our observations in the Parvati and Kul River sections to the W and
NNE respectively of the Ramgarh structure (Fig. 2), and around *19.4 km south
of this structure show wide-spaced (*2 m) cross-cutting fractures trending NE-SW
and NW-SE (Figs. 4b and 6a). The cross-cutting field relationship at the Parvati
River section suggests that the NE-SW fracture could have developed early in
sequence, followed by the formation of NW-SE fracture.
The bedding plane (So) in sandstone on the rim crest of the Ramgarh structure
show quaquaversal dips at moderate angle (Fig. 4c) (cf. Sharma 1973; Ramasamy
1987). In places, the rim sandstones also dip *vertical (Fig. 4d). The sandstones
examined on the rim crest (Fig. 4e, f), the structure wall (Fig. 4g), and the shale
exposed at the base inside the structure (Fig. 4h) are extremely fractured. The
sequence of fracture development within the Ramgarh structure appears to be
Remote Sensing, Structural and Rock Magnetic Analyses … 335
JFig. 4 a Cross sectional view of flat-lying Vindhyan Supergroup of sedimentary rocks including
limestone (lst), sandstone (sst) and shale at Parvati River section at *4 km W of Ramgarh
structure [width of photograph is *30 m, Ramgarh structure (RS) to E in background], b plan
view of flat-lying Vindhyan Supergroup of sedimentary rocks at Parvati River section showing
development of well-spaced (*2 m) orthogonal fractures where NW-SE fracture cross-cuts
NE-SW fracture (length of hammer *27.9 cm), c quaquaversal dip of bedding plane (So) in
sandstone on the western rim crest of Ramgarh structure, white arrow indicates towards the center
of structure to E. d sub-vertical orientation of So in sandstone on NNE rim crest [hammer
(*27.9 cm length) placed parallel to So]. e plan view of N-S subvertical fracture (shown by arrow)
and cross fracture on sandstones exposed at SSW sector of rim crest (hammer’s head points to N).
f plan view of sandstone exposed on SSW rim crest showing development of N-S, E-W and
NE-SW trending fractures intersecting each other (hammer’s head pointed towards E).
g cross-sectional view of inner wall of Ramgarh structure at the southern sector of rim crest
showing highly fractured sandstone with flowage of Mn-rich solution along few selected fractures
(shown by arrow). h plan view of Ganurgarh Shale exposed at the base inside the Ramgarh
structure towards SW close to the entrance showing development of extensive cross-cutting
fractures. i plan view of sandstone on SSW structure rim showing NE-SW fracture crosscuts
NW-SE fracture, both fractures are cross-cut by E-W fracture. j a view of rock sample at Kul River
section towards NE of Ramgarh structure showing gneissic foliation (diameter of coin shown as
scale is 25 mm). k cross-sectional view of inner wall of Ramgarh structure at southern sector
showing flowage of Mn-rich solution along few fracture planes (shown by arrow) (diameter of
coin shown as scale is 23 mm). l plan view on northern rim crest showing injection of Mn-solution
along So plane of tilted sandstone (shown by arrow). m plan view on SE rim crest showing
injection of Mn-solution along So plane of sandstone and fracture planes (shown by arrow, width
of photograph *50 cm). n a shatter cone-like structure on the inner wall of Ramgarh structure at
NW rim crest. o view of *6 m high central uplift with an ancient temple at top (shown by arrow),
and. p cross sectional view of extremely fractured sandstone in central uplift (width of photograph
*20 cm)
Unlike the surrounding undeformed sedimentary country rocks (Fig. 6a), the
rocks inside the Ramgarh structure show extreme development of fractures
(Fig. 6b). These fractures, including both the joints and faults, could primarily be
classified into two types: vertical and moderately dipping fractures. The former
fractures trend in all directions, while the dipping fractures trend *NE-SW and
NW-SE. In the N-NNE-ENE sectors of the Ramgarh structure (Fig. 3), the dipping
fractures trend *NE-SW and dip at moderate angle both towards NW and SE
(Fig. 6c, d). Two subordinate sets of fractures are also identified having (1) trend of
NW-SE and dipping moderately towards SW, and (2) trend E-W and dipping
moderately towards S. The second important sector, where sufficient data were
available, is the S-SSW-WSW sectors (Fig. 3). Here, two sets of fractures are
noticed, e.g. the NE-SW trending fracture dipping towards NW, and the NW-SE
trending fracture dipping towards NE (Fig. 6g, h). In the W-WNW-NNW sectors,
we have the NE-SW trending fracture dipping both towards NW and SE, and the
NW-SE trending fracture dips towards NE. There is also an E-W trending fracture
dips towards the south.
338 S. Misra et al.
Fig. 4 (continued)
Remote Sensing, Structural and Rock Magnetic Analyses … 339
An examination of the ASTER image of the greater areal extent suggests that the
Ramgarh structure is most likely situated on the palaeo-channel of the Parvati
River, which is indicated by the presence of a NNW-SSE trending abundant
channel of this river to the south and west of this structure (Fig. 7). The present
course of Parvati River to the west of the Ramgarh structure is extremely irregular
compared to the rest of the course of this river. The SRTM-DEM image further
shows that the river marked ‘4’ was previously a tributary of Parvati River
(Fig. 2c). Presently this river changes its course leaving its palaeo-channel to the
south of Ramgarh structure (channel marked 6) and merges with the Kul River.
The NRM of all four sub-samples from sample RJ26 (Fig. 3) ranges between 9 and
19 Am−1, and their REM (i.e., NRM/SIRM ratio in %) varies between 40 and 78%.
In all sub-samples, the NRM decay curves are above the SIRM decay curves for the
lower range of AF demagnetization fields, whereas there are always overlap
between these two curves in the higher AF fields (Fig. 8a–d) indicating a mixture of
Single Domain (SD), Pseudo-SD and Multi Domain (MD) grains in them. The IRM
acquisition curves of all our sub-samples seem to saturate at higher fields of
500 mT indicating high coercivity magnetic mineral as the remanence carrier
(Fig. 8e). Their IRM backfield curves indicated a range of coercivity of remanence
(Hcr) between 30 and 45 mT (Fig. 8f).
5 Discussion
Fig. 5 Stereographic projections of bedding planes (So) in sandstones exposed along the rim crest
of the Ramgarh structure, the sectors selected along the rim (Fig. 3) is purely geographical after the
convention described in Misra et al. (2010), for detail see text
Remote Sensing, Structural and Rock Magnetic Analyses … 341
Fig. 5 (continued)
342 S. Misra et al.
(Prasad 1984), and in area surrounding the Ramgarh structure (Figs. 4b and 6a).
The evidence of extensive dextral shear along the NE-SW fracture plane
cross-cutting the Ramgarh structure, and following minor sinistral shear along the
E-W fracture planes are only seen within the Ramgarh structure (Figs. 2a and 3).
All these brittle deformation features suggest that the sedimentary country rocks
within the Ramgarh structure could have experienced a sudden shock that generated
new fractures and reactivated the pre-existing fractures promoting limited slip.
The quaquaversal dips of the country rock sandstone along the rim of the
Ramgarh structure (Fig. 5), along with its extreme brittle deformations resemble in
several aspects to those observed in the Lonar impact Crater, India (Kumar 2005;
Misra et al. 2010). The Ramgarh structure is * rectangular (Figs. 2a and 3), which
is similar in shape to the Arizona Crater, USA (Grieve et al. 1988). The present
diameter/depth ratio of the Ramgarh structure is *12 (measured from Google Earth
Imagery 2017), which is within the range (10–20) of the terrestrial complex asteroid
344 S. Misra et al.
Fig. 8 a–d Alternating field demagnetization of NRM and SIRM of sub-samples from RJ26
glassy sample recovered from Ramgarh structure (Fig. 3), note position of NRM decay curves
above SIRM decay curve for lower range of AF demagnetization field. e IRM acquisition curves of
all four glassy samples showing saturation at higher field of *500 mT. f IRM backfield curves of
glassy sub-samples
impact craters (cf. Smith 1971; Koeberl and Sharpton 2017). Like other terrestrial
complex impact crater, the Ramgarh structure has a prominent conical central peak
(Fig. 4o) (also Grieve et al. 1988; Master and Pandit 1999), which consists of
sandstone and has a present height of *6 m (Misra et al. 2008a). The subsurface
presence of this central peak below the present alluvium is also modelled based on
borehole data (Ramasamy 1987). Hence it can be modelled that the extreme brittle
deformation of the country rocks observed within the Ramgarh structure, the
development of shatter cone like structure on the wall of the Ramgarh structure
(Fig. 4n) and central uplift within the structure (Fig. 4o) were resulted due to a
sudden shock effect caused by asteroid impact (cf. French 1998).
Remote Sensing, Structural and Rock Magnetic Analyses … 345
Our previous study on magnetic particles recovered from the Ramgarh structure
(Das et al. 2009) and our present study on four small (mm-size) sub-samples from a
cm-size magnetic glassy rock, recovered from the rocky debris lying on the outer
slope of western rim of this crater (Figs. 3 and 8), confirm that the glassy sample/
particles from this structure have very high NRM (*2–19 Am−1) and high REM
(*7–145%) indicating a high magnetic field during their formation. The REM
values provide an estimation of the paleomagnetic field (Kletetschka et al. 2003)
and a ratio close to 1.5% indicates an Earth-magnetic field (Gattacceca and
Rochette 2004; Kletetschka et al. 2004; Yu 2006). The low-field processes other
than thermal remanent magnetization, e.g., viscous or chemical remanent magne-
tization, yield lower REM values for the same paleofield (Fuller et al. 1988),
whereas the high-field processes, e.g., lightning-induced or plasma-induced mag-
netization, yield REM values above 10% (Wasilewski and Dickinson 2000). As the
lightening alone cannot produce the observed ring-like shape of the Ramgarh
structure and its brittle deformations, the increase of initial NRM intensity of the
Ramgarh magnetic particles (Das et al. 2009, and Fig. 8a–d) could only be acquired
by any secondary magnetization process such as shock metamorphism by asteroid
impact (cf. Gold and Soter 1976; Schultz and Srnka 1980). This impact also led to
brittle deformations of sedimentary country rocks. The role of the impact was to
create a high magnetic field during the formation of the Ramgarh structure, which
perhaps covers an area >*20 km2 during its formation (cf. Crawford and Schultz
1988, 1999). Hence, our present observations on structural geology and rock
magnetics, and previous observations on petrography (Misra et al. 2008b; Das et al.
2011; Purohit and Sisodia, 2013) and geochemistry (Misra et al. 2008b; 2013) have
enhanced the possibility of an asteroid impact origin for the Ramgarh structure.
One of the against-hypothesis on the asteroid impact origin for the Ramgarh
structure could be the general absence of PDF in the sandstones in this structure (cf.
Reimold et al. 2006). However, a detailed observation under the optical microscope
shows that the sandstones from this structure is extremely fine grained (grain size
200 lm) (Fig. 10 in Appendix), and such fine-grained rocks would inhibit the
formation of PDF than their coarse-grained counterparts (cf. French and Koeberl
2010 and references therein). The absence of any PDF under the microscope,
therefore, does not negate the asteroid impact origin of the Ramgarh structure (cf.
Chen et al. 2017; Melosh 2017). In fact, a few of quartz grains within the bed rock
sandstone from the rim of the Ramgarh structure show one set very closed-spaced
( 20 lm) fractures (PFs) (Das et al. 2011), which could favour the asteroid impact
origin of this structure (cf. French 1998). Further, more recent observation of Purohot
and Sisodia (2013) on the universal stage measurements probably identified a few
relict PDFs within the bed rock sandstone from this structure.
346 S. Misra et al.
The quaquaversal dips of the rim crest sandstones are extremely disturbed in the
NNE, ENE and E sectors, and in the WNW and NNW sectors (Fig. 5). In the
former three sectors, the dips of So planes in sandstones vary from N to E, while in
the latter two sectors it varies in wide range from SW to NNE. The deviation of dips
of the rim crest sandstones from its typical quaquaversal dips in the NNE, ENE and
E sectors is definitely due to the effect of the dextral NE-SW faulting that transected
and partly displaced the structure rim (Figs. 2a and 3) (cf. Sharma 1973). However,
the reason behind the same type of deviation of dips for the rim sandstone in the
WNW and NNW sectors is not clearly understood in the present study. It could be
due to another NW-SE dextral faulting that was early in sequence to that of the
major NE-SW fault and cross-cut the Ramgarh structure (Ramasamy 1987). Hence
the rectangular shape of the Ramgarh structure (Figs. 2a and 3) is a combined result
of slip along dextral NW-SE, dextral NE-SW and sinistral E-W faults in sequence.
It appears that the NE-SW and NW-SE fracture sets within the surrounding
Vindhyan Supergroup were reactivated within the Ramgarh structure due to the
shock waves generated by the impact (cf. French 1998). A similar example could be
the Arizona Crater, USA. The rectangular shape of this crater was resulted due to
impact on sedimentary target rocks with pre-existing joint sets and faults (Poelchau
et al. 2009).
The ‘rheid flow of shale’ model of Ramasamy (1987) to explain the origin of the
Ramgarh structure was mainly based on the morphology of the folds observed in
shale-siltstone sequence in central borehole between 20 and 210 m depth.
According to his description, these shale and siltstone layers show intricate
disharmonic folds. The limbs and axial surfaces of the folds are parallel and vertical
with either sharp or rounded hinges. These folds are refolded with sub-horizontal
axial surfaces. To explain this folded sequence, Ramasamy (1987) introduced the
concept of vertical injection of rheid shales in the axial portion of the Ramgarh
structure. However, no mechanism is suggested to explain why this vertical
injection of solid rocks took place at the central part of this structure. It is under-
stood from the impact cratering process that during the end part of excavation stage
and the following modification stage of evolution of complex impact crater, the
solid materials flow vertically upward close to the central uplift of the impact crater
(French 1998). This type of material flow could thus explain the vertical folding
developed in the shale and siltstone in the central borehole within the Ramgarh
structure.
The Ramgarh structure is situated on the palaeo-channel of Parvati River
(Figs. 2c and 7) and hence its formation must has post-dated the formation of this
river. It appears that the impact has resulted the shift of river courses of Parvati and
its tributary adjacent to Ramgarh structure. Impact on river channel, hence,
enhances the possibility of involvement of surface water during the evolution of this
structure. Our previous microprobe analyses on spherules recovered from this
structure consistently show total major oxide around 80 wt% supporting possible
Remote Sensing, Structural and Rock Magnetic Analyses … 347
6 Conclusions
(f) Observation on remote sensing images confirms that the asteroid impact
took place on the palaeo-channel of Parvati River, which is now displaced
towards W.
Acknowledgements The first author (S. M.) is grateful to PLANEX, Indian Space Research
Organization, and NRF, South Africa (grant no. 91089) for supporting this research work. Special
thanks to Anand Dube of India, for helpful guidance to the first author during the field work, to
Horton Newsom of USA for his continuous encouragement during the progress of this research,
and to Tesfaye Kidane for helping in stereoplot software. We are indebted to Dr. Soumyajit
Mukherjee and an anonymous reviewer for their constructive comments on the early version of the
manuscript.
Appendix
Fig. 10 a, b Sandstone sample from Kul River section N of Ramgarh structure (location R 38),
the dark matters among the grains (shown by arrow in b) are manganese encrustation. c, d impure
limestone from Ramgarh structure (R 18) containing numerous fine grain quartz sand particles. e,
f percolation of manganese solution within the sandstone (shown in box) sample from the rim of
Ramgarh structure (R 11). g, h manganese encrustation containing numerous sub-angular to
sub-rounded quartz grains (R 15). GPS locations of samples: R 11: 25° 19.378′N, 76° 37.933′E; R
15: 25° 19.582′N, 76° 37.570′E; R 18: 25° 19.406′N, 76° 37.071′E; R 38: 25° 22.929′N,
76° 36.961′E. Note fineness of grain size of sandstones occurring on rim crest of Ramgarh
structure
350 S. Misra et al.
References
Ahmad K (1958) Paleogeography of Central India in the Vindhyan period. Geological Suvey of
India Records 87, 531–548
Ahmed N, Bhardwaj BD, Sajid HA, Hasnain I (1974) Ramgarh meteorite crater. Current Science
43, 598
Auden JB (1933) Vindhyan sedimentation in the Son valley, Mirzapur district. Geological Survey
of India Memoirs 62, 141–250
Balasundaram MS, Dube A (1973) Ramgarh structure, India. Nature 242, 40
Bose PK, Sarkar S, Chakrabarty S, Banerjee S (2001) Overview of the Meso- to Neoproterozoic
evolution of the Vindhyan basin, central India. Sedimentary Geology 41–142, 395–419
Chen J, Elmi C, Goldsby DL, Gieré R (2017) Generation of shock lamellae and melting in rocks
by lightning-induced shock waves and electrical heating. Geophysical Research Letters 44,
8757–8768
Crawford AR (1972) Possible impact structure in India. Nature 237, 96
Crawford DA, Schultz PH (1988) Laboratory observations of impact–generated magnetic fields.
Nature 336, 50–52
Crawford DA, Schultz PH (1999) Electromagnetic properties of impact-generated plasma, vapor
and debris. International Journal of Impact Engineering 23, 169–180
Das PK, Misra S, Basavaiah N, Newsom H, Dube A (2009). Rock magnetic evidence of asteroid
impact origin of Ramgarh structure. India. In: 40th Lunar and planetary science conference.
Abstract no. 1466
Das PK, Misra S, Newsom HE, Sisodia MS (2011) Possible planer fractures, coesite, and
accretionary lapilli from Ramgarh structure, India: new evidence suggesting an impact origin
of the crater. In: 42nd Lunar and planetary science conference. Abstract no. 1294
French BM (1998) Traces of catastrophe: a handbook of shock-metamorphic effects in terrestrial
meteorite impact structures. In: Lunar and planetary institute contribution series. vol 954, p 120
French BM, Koeberl C (2010) The convincing identification of terrestrial meteorite impact
structures: what works, what doesn’t, and why. Earth-Science Reviews 98, 123–170
Fuller M, Cisowski S, Hart M, Haston R, Schmidtke E, Jarrard R (1988) NRM:IRM(S)
demagnetization plots; an aid to the interpretation of natural remanent magnetization.
Geophysical Research Letters 15, 518–521
Gattacceca J, Rochette P (2004) Toward a robust normalized magnetic paleointensity method
applied to meteorites. Earth and Planetary Science Letters 227, 377–393
GLCF (2007) Global land cover facility. University of Maryland, USA. Available online at http://
glcf.umiacs.umd.edu
Gold T, Soter S (1976) Cometary impact and the magnetization of the Moon. Planetary and Space
Science 24, 45–54
Grant JA (1999) Evaluating the evolution of process specific degradation signatures around impact
craters. International Journal of Impact Engineering 23, 331–340
Graup G (1981) Terrestrial chondrules, glass spherules and accretionary lapilli from the suevite,
Ries Crater, Germany. Earth and Planetary Science Letters 55, 407–418
Grieve RAF, Wood CA, Garvin JB, Mclaughlin G, McHone JF (1988) Astronaut’s guide to
terrestrial impact craters. Lunar and planetary institute technical report 88-03, Houston, p 89
Kletetschka G, Kohout T, Wasilewski PJ (2003) Magnetic remanence in the Murchison meteorite.
Meteoritics and Planetary Science 38, 399–405
Kletetschka G, Acuna MH, Kohout T, Wasilewski PJ, Connerney JEP (2004) An empirical scaling
law for acquisition of thermoremanent magnetization. Earth and Planetary Science Letters 226,
521–528
Koeberl C, Sharpton VL (2017) Terrestrial impact craters, 2nd edn. www.lpi.usra.edu/
publications/slidesets/craters
Remote Sensing, Structural and Rock Magnetic Analyses … 351
Koeberl C, Brandstätter F, Glass BP, Hecht L, Mader D, Reimold WU (2007) Uppermost impact
fall back layer in the Bosumtwi crater (Ghana): mineralogy, geochemistry, and comparison
with ivory coast tektites. Meteoritics and Planetary Science 42, 709–729
Kumar PS (2005) Structural effects of meteorite impact on basalt: evidence from lonar crater.
Journal of Geophysical Research 110, B12402
Kumar J, Negi MS, Sharma R, Saha D, Mayor S, Asthana M (2011) Ramgarh magnetic anomaly
in the Chambal valley sector of Vindhyan basin: a possible meteorite impact structure and its
implications in hydrocarbon exploration. In: American Association of Petroleum Geologists,
Search and Discovery. Article #80145
Lowrie W, Fuller M (1971) On the alternating field demagnetization characteristics of
multidomain thermoremanent magnetization in magnetite. Journal of Geophysical Research
76, 6339–6349
Mallet FR (1869) On the Vindhyan series, as exhibited in the North-western and Central Province
of India. Memoir of Geological Survey of India 7(Part 1), 129
Malone SJ, Meert JG, Banerjee DM, Pandit MK, Tamrat E, Kamenov GD, Pradhan VR, Sohl LE
(2008) Paleomagnetism and detrital zircon geochronology of the Upper Vindhyan sequence,
Son valley and Rajasthan, India: a ca. 1000 Ma closure age for the Purana basins? Precambrian
Research 164, 137–159
Master S, Pandit MK (1999) New evidence for an impact origin of the Ramgarh structure.
Meteoritics and Planetary Science 34, 4
McElhinny MW, McFadden PL (2000) Paleomagnetism: continents and oceans. In: International
geophysics series, vol 73. Academic Press, San Diego, CA, p 386. ISBN: 0124833551
Melosh HJ (2017) Impact geologists, beware! Geophys Res Lett 44, 8873–8874
Misra S, Dube A, Srivastava PK, Newsom HE (2008a) Time of formation of Ramgarh crater,
India-constraints from geological structures. In: 39th Lunar and planetary science conference.
Abstract. no. 1502
Misra S, Lashkari G, Panda D, Dube A, Sisodia MS, Newsom HE, Sengupta D (2008b)
Geochemical evidence for the meteorite impact origin of Ramgarh structure, India. 39th Lunar
and planetary science conference. Abstract. no. 1499
Misra S, Arif M, Basavaiah N, Srivastava PK, Dube A (2010) Structural and anisotropy of
magnetic susceptibility (AMS) evidence for oblique impact on terrestrial basalt flows: Lonar
crater, India. Bulletin of Geological Society of America 122, 563–574
Misra S, Panda D, Ray D, Newsom H, Dube A, Sisodia MS (2013) Geochemistry of glassy rocks
from Ramgarh structure, India. In: 44th Lunar and planetary science conference. Abstract. no.
1020
Newsom H, Gasnault O, Le Mouelic S, Mangold N, Le Deit L, Wiens R, Anderson R, Edgar L,
Herkenhoff K, Johnson JR, Bridges N, Grotzinger JP, Gupta S, Jacob S (2016) Long distance
observation with the ChemCam remote micro-imager: mount sharp and related deposits on gale
crater floor? Geological Society of America, Denver, 25–28 Sept 2016
Oldham T (1856) Remarks on the classification of the rocks of central India resulting from the
investigation of the Geological Survey. Journal the Asiatic Society, Calcutta 25, 224–256
Poelchau MH, Kenkmann T, Kring DA (2009) Rim uplift and crater shape in meteor crater: effects
of target heterogeneities and trajectory obliquity. Journal of Geophysical Research 114,
E01006
Prasad B (1984) Geology, sedimentation and palageography of the Vindhyan Supergroup,
Southeast Rajasthan. Geological Survey of India Memoirs 116, 1–107
Purohot V, Sisodia MS (2013) Universal-stage measurements of planar deformation features in
shocked quartz grains recovered from Ramgarh structure. In: 44th Lunar and planetary science
conference. Abstract. no. 1151
Ramasamy SM (1987) Evolution of Ramgarh dome, Rajasthan: India. Records of Geological
Survey of India 113, 13–22
Rana S, Agrawal V (2016) Microscopic evidences for the impact origin of Ramgarh structure,
Rajasthan, India. Journal of Indian Geophysical Union 20, 544–550
352 S. Misra et al.
Reimold WU, Trepmann C, Simonson B (2006) Comment on “Impact origin of the Ramgarh
structure, Rajasthan: some new evidences by M. S. Sisodia, G. Lashkari and N. Bhandari.
Journal of Geological Society of India, v. 67, pp. 423–431”. Journal of Geological Society of
India 68, 561–563
Schultz PH, Srnka LJ (1980) Cometary collision on the moon and mercury. Nature 284, 22–26
Sharma HS (1973) Ramgarh structure, India. Nature 242, 39–40
Sisodia MS, Lashkari G (2003) Ramgarh structure, Rajasthan, India: meteorite impact evidences.
Workshop on Impact cratering: bridging the gap between modeling and observations, Houston.
Abstract no. 8008
Sisodia MS, Lashkari G, Bhandari N (2006a) Impact origin of the Ramgarh structure, Rajasthan:
Some new evidences. Journal of Geological Society of India 67, 423–431
Sisodia MS, Lashkari G, Bhandari N (2006b) Reply to “The comment on Impact origin of the
Ramgarh structure, Rajasthan: Some new evidences by W. U. Reimold, C. Trepmann and B.
Simonson”. Journal of Geological Society of India 68, 561–563
Smith EI (1971) Determination of origin of small lunar and terrestrial craters by depth diameter
ratio. Journal of Geophysical Research, 76, 5683–5689
Tandon SK, Pant CC, Casshyap SM (1991) Sedimentary basins of India-Tectonic context.
Gyanodaya Prakashan, Nainital
Vernooij MGC, Langenhorst F (2005) Experimental reproduction of tectonic deformation lamellae
in quartz and comparison to shock-induced planar deformation features. Meteoritics and
Planetary Science 40, 1353–1361
Wasilewski P, Dickinson T (2000) Aspects of the validation of magnetic remanence in meteorites.
Meteoritics and Planetary Science 35, 537–544
Yu YJ (2006) How accurately can NRM⁄SIRM determine the ancient planetary magnetic field
intensity? Earth and Planetary Science Letters 250, 27–37
Geology, Structural Architecture
and Tectonic Framework of the Rocks
of Southern Lalitpur District Uttar
Pradesh, India: An Epitome
of the Indian Peninsular Shield
1 Introduction
The southern part of Uttar Pradesh is geologically known for exposing a part of the
Indian peninsular shield. Both basement and cover rocks are exposed at several
areas. The Lalitpur district of this region typically exposes several geological ele-
ments of the peninsular shield. The southern part of Lalitpur district (Fig. 1) epit-
omizes the peninsular shield to a larger extent. This area has been studied by various
workers mainly from the viewpoints of geology, rock types, and mineral resources.
The early studies were confined to identifying the major rock groups. Later studies
focused on identifying various rock types together with the associated metasedi-
mentary sequences. Recent studies focus amongst others, on mineralogical, geo-
chemical and geochronological studies to understand the evolution of the ancient
crust and crustal growth in the context of the peninsular shield. Despite geological
studies on several aspects, detailed lithostratigraphic and structural-tectonic studies
of south Lalitpur district are scanty. This paper fills up this gap.
2 Geological Set up
Fig. 1 Map showing the location of the study area. The inset shows the state Uttar Pradesh
On a regional basis, the area has been referred to, directly or indirectly, by
several workers. Medlicott (1859) was possibly the first to make a geological study
of the rocks of the Bundelkhand region. He made a detailed study of the Banded
Iron Formation (Baraitha) and various schistose rocks from the southwestern part of
the Bundelkhand region and grouped them as crystallines of Bundelkhand. The
metasedimentary belt around Gwalior was studied by Hacket (1870). Much later,
Heron (1935) studied the crystalline rocks from several areas of the Bundelkhand
region and he referred these rocks as “Bundelkhand Granite”.
During the above period, study remained confined only to some general aspects
of the Bundelkhand region. Thereafter, the trend of study gradually shifted to some
specific aspects of the region. Srivastava (1951–1952) was one of the earlier
workers who carried out detailed geological and mineralogical investigations in
several areas of the Bundelkhand region and published his new findings on geo-
logical and petrological characters of the Bundelkhand granites. He also reported
pyrophyllite deposits in some parts of the Bundelkhand massif. By that time, the
granitic rocks of the Bundelkhand massif became the major subject matter of study
mainly because a variety of granitic rocks were reported till then. Jhingran (1958)
Geology, Structural Architecture and Tectonic Framework … 355
identified two major types of granites: pink granite and grey granite and he sug-
gested that the former (pink granite) is younger than the latter. He also identified ten
types of granites within the great batholithic massif on the basis of grain size, colour
of feldspar and presence or absence of ferromagnesian minerals. In addition to
these, he also reported gneisses, quartz–reefs, basic dikesand tuffaceous serpentinite
rocks within the massif. Saxena (1961) suggested a lithostratigraphic correlation of
the metasediments of the Bundelkhand massif with the middle Dharwarian rocks of
South India and the Bundelkhand granitic activity as equivalent of the Closepet
granite.
During the nineteen hundred seventies, the different rock types of the
Bundelkhand region including lithostratigraphic sequences of some areas came to
limelight. Prakash et al. (1975) carried out a detailed lithostratigraphic study of the
southwestern part of Bundelkhand massif and distinguished three tectonic blocks in
this massif: a northern block (Jhansi–Lalitpur) consisting of various types of
granitoids with enclaves of metasedimentary rocks, a southern block (south of
Lalitpur) comprising of Bundelkhand granitoids including the low grade
metasediments (Mehroni Group) together with the overlying Vindhyan sediments,
and a central block mainly exposing the pink, massive granites. Misra and Sharma
(1975) identified metasedimentary rocks within the Bundelkhand complex and
suggested that the Bundelkhand rocks can be divided into four formations; from
lower to upper these are Kuraicha Formation of high grade metamorphic rocks,
Palar Formation with low grade metamorphic rocks, Bundelkhand Granite, and
Bundelkhand basic intrusives.
The nineteen hundred eighties witnessed a multifarious progress on several yet
untouched aspects on the geology of the region. In this connection, the contribu-
tions of Basu (1986, 2001, 2007, 2010) deserve special mention. He carried out
extensive field investigations and mapped almost the entire area of the Bundelkhand
Craton. He proposed an evolutionary model for the Precambrian crust on the basis
of petrological, structural and chemical characteristics of various types of granitoids
and basic rocks. He suggested that the metallogenic and tectonometamorphic events
of the massif were regionally dominated by plutonic and hypabyssal rocks.
Despite the available information on several aspects of the geology of the region,
studies from structural angles were almost lacking. Bhattacharya (1985) estimated
tectonic strain for the rocks of the Bundelkhand massif and has shown that the
mineralogical behaviour of the rocks has controlled the deformation patterns of the
massif. He (Bhattacharya 1986) developed a semi-quantitative method for identi-
fying different fold episodes of deformed terrains and on applying this method to
the rocks of the Bundelkhand massif, he identified five different folding episodes
for the massif.
On a much larger perspective, Radhakrishna (1989) identified a “Central Indian
Tectonic Zone” encompassing the SONATA Belt that marks the junction between
the Bundelkhand block in the north and the Peninsular block in the south. Sarkar
et al. (1989, 1995, 1997) on the basis of the geochronology and petrology of
various rock types of the Bundelkhand complex suggested that the granitic rocks
belong to three distinct intrusive phases. They also suggested that the trondhjemitic
356 G. K. Dinkar et al.
gneiss containing Rb-Sr isotopes may be considered as 3.5 Ga old (Baghora TTG
suite). The older crust of the region (Archaean to Palaeoproterozoic) is comprised
of amphibolites, quartzites, BIF, schists, marbles and calc–silicate rocks and that
these rocks are intruded by 3.3 Ga old syn-to-late tectonic Na-rich
Tonalite-Trondjhemite-Granodiorite (TTG) rocks. The oldest metamorphic events
of the Bundelkhand region have been dated at 3.3 Ga, while the oldest relics of
younger geneisses may be 2.7–2.5 Ga old (Mondal et al. 2002). Attempts have
been made to work out the crustal evolutionary pattern of the Bundelkhand region
(Singh et al. 2007; Singh and Dwivedi 2009; Singh 2012).
Our knowledge on the hitherto unknown crustal shear systems of the
Bundelkhand massif is definitely a new addition as came into light by the joint work
of A. R. Bhattacharya and S. P. Singh (Singh and Bhattacharya 2010, 2017;
Bhattacharya and Singh 2013). They have identified the following four different
crustal shear systems in the Bundelkhand massif. (1) The E-W shear system: It is
mainly confined to the central parts of the massif and is represented by mylonites
showing vertical foliation together with a greenstone belt developed at several
places with dimensions ranging from about 10 30 m to 30 50 m. This is an
example of crustal scale vertical ductile shear zone and is possibly the first report of
vertical crustal shear zone in the Indian subcontinent. (2) The NE-SW shear system:
It is represented by the quartz reefs that constitute the most spectacular structure in
the Bundelkhand terrain. The quartz reefs constitute strike-slip dominated vertical
to subvertical shear zones with dominant sinistral sense of shear and they represent
Proterozoic crustal scale extensional events in the Bundelkhand massif. (3) The
NW-SE shear system: These are in the form of a number of minor basic igneous
dikes emplaced in the fractures developed due to extensional processes followed by
strike-slip shearing. (4) N-S shear system: This is a major tectono–magmatic event
of the Bundelkhand craton and is represented by N-S trending quartz and
quarto-feldspathic veins. This is the youngest shear system that cross-cuts all the
earlier (E-W, NE-SW and NW-SE) shear systems.
In the context of vertical shear zones of the Bundelkhand massif, it may be
mentioned here that elsewhere in the Indian subcontinent, shear zones with
sub-vertical foliations are reported from Karakoram Shear Zone (Mukherjee 2011)
and vertical brittle shear zones from trap rocks near Mumbai (Misra et a. 2014).
Recently, Dinkar (2016) carried out a detailed study on the structure and
deformation patterns of the rocks of the area around Sonrai. He presented a geo-
logical and a structural map of the area both on the 1:50,000 scale and worked out
the structure and structural evolution of the rocks of the area
In a nutshell, south Lalitpur district is characterized by three different geological
subdivisions. The northern part is represented by the Bundelkhand Gneissic
Complex, the central part by the rocks of the Bijawar Group and the southern part
by the Vindhyan Supergroup. In addition, there are a few minor rock units also that
fall in any of these three major subdivisions. The geological information brought to
light through this paper, and the results arrived at, should thus advance our
knowledge on the Indian peninsular shield.
Geology, Structural Architecture and Tectonic Framework … 357
About 460 km2 has been mapped on 1:50,000 scale covering the area around
Digwar in the northwest, Balla in the northeast, Madanpur in the southwest and
Kurrat in the southeast. The study area (Fig. 1) encompasses the river valley areas
of the Rohini River in the west, Bandai River in the central part and Dhasan River
in the east.
4 Physiography
The physiography of the southern Lalitpur district is rather typical from the
remaining parts of the southern Uttar Pradesh. The main reason is that in this part
the physiography appears to be related to the geology of the area. The general
surface level of flat lying regions is 400 m above the mean sea level with hillocks
rising up to about 566 m from the surrounding areas. The southern part (Fig. 2) of
the area represents plateaus and ridges while the vast expanse of the central (Fig. 2)
and northern parts is mostly alluvial exposing scattered hillocks. In general, we
identify three physiographic subdivisions (Fig. 2). The northern subdivision is
mostly alluvial exposing scattered hillocks. The central subdivision mainly includes
E-W trending arcuate to elongated hills. The southern part of the area is occupied
by plateau with E-W trending ridges rising up to about 566 m.
Fig. 2 Photographs showing the physiographic features of the southern Lalitpur district. Our
studies suggest that the area shows three different geological subdivisions: the northern subdivision
exposing the Bundelkhand Gneissic Complex, the central subdivision exposing a sedimentary/
metasedimentary belt consisting of the Bijawar Group of rocks while the southern one exposes the
rocks of the Vindhyan Supergroup. Each of these geological subdivisions has been observed to
show its typical physiography. The northern subdivision is mostly alluvial exposing scattered
hillocks. The central subdivision mainly includes E-W trending arcuate to elongated, discontin-
uous hills. The southern one shows plateau with long, E-W trending ridges. The photographs
shown here are representative of the latter two types of physiography, i.e. those shown by the rocks
of the Bijawar Group (a) and those of the Vindhyan Supergroup (b)
358 G. K. Dinkar et al.
5 Lithostratigraphy
From the above discussion, we identify three distinct belts in southern Lalitpur
district as described above. The basement rocks are represented by the BGC, the
metasedimentary sequence by the Bijawar Group while the cover sediments by the
Vindhyan Supergroup. In addition, a few outcrops of Deccan Trap also expose at
the southern parts of the area. A generalized lithostratigraphic succession has been
presented in Table 1. Detailed lithostratigraphic description and the various rock
types of the study area are given below. The lithological characteristics and
lithostratigraphic details of the rocks of the area are described here:
(A) Bundelkhand Gneissic Complex (BGC), (B) Bijawar Group, and (C) Vindhyan
Supergroup (Fig. 3).
Fig. 3 Photographs showing the physiographic features of the study area. a Central part with
east-west trending ridges, b southern part dominated by plateau. These physiographic subdivisions
are characterized by their typical geological attributes (see text)
Fig. 4 Geological map of the southern Lalitpur a district, Uttar Pradesh. Modified after Prakash
et al. (1975) and Sharma (1982)
In the course of our field study, we have identified a few lithologic associations
within this complex as described below.
(i) Bundelkhand Gneisses
The Bundelkhand Gneisses (Fig. 5) are exposed in the northern parts of the study area.
These rocks are dominantly constituted of moderate to high grade metamorphic rocks.
The major rock types include granite-gneiss, migmatite, biotite-gneiss, tonlite-
trondhjemite–granodiorite (TTG), hornblende-biotite-gneiss, garnet-sillimanite-gneiss,
garnet-sileimanite–cordierite-gneiss and amphibolites. The foliation of the gneissic rocks
generally trend ENE-WSW to E-W with dips of 30°–50° southwards.
Fig. 5 Field photograph showing outcrop of gneissic rock. Loc. North of Balla
Geology, Structural Architecture and Tectonic Framework … 361
Fig. 6 Field photograph of the coarse grained pink granite. The diameter of the coin is 2.5 cm.
Loc. Near Manpura
362 G. K. Dinkar et al.
Fig. 7 Field photograph of the medium grained pink granite. The diameter of the coin is 2.5 cm.
Loc. 200 m east of Imaliya Khurd
Fig. 8 Field photograph of the grey granite. The diameter of the coin is 2.5 cm. Loc. Near Balla
Geology, Structural Architecture and Tectonic Framework … 363
(e) Pegmatites
5–12 cm wide and 1–10 m long pegmatite veins have been noticed at several
places. Most of them trend N-S, though variations are also there at many places.
The pegmatite commonly shows large crystal of orthoclase and quartz, e.g. north of
Sonrai (Fig. 10). Occasionally, the veins show large crystals of biotite and
Fig. 10 Field photograph of pegmatite showing large crystals of orthoclase and quartz. The
diameter of the coin is 2.5 cm. Loc. North of Sonrai
364 G. K. Dinkar et al.
muscovite, e.g. near Nathikhera. It appears that the pegmatites have intruded into
the preexisting rocks in phases as based on cross-cutting relations. Basu (1986) has
reported three generations of pegmatite veins in the Bundelkhand massif.
(iii) Mafic and Ultramafic Suite (Metavolcanic Suite)
A variety of mafic and ultramafic (Fig. 11a, b) rocks are exposed in various parts of
the BGC. We assign these rocks to constitute a metavolcanic suite that is mainly
represented by metaperidotite, metapyroxinite and coarse to medium grained gab-
bro. This suite of rocks is intrusive into different types of granites as well as
gabbroic rocks. Earlier Prakash et al. (1975) and Sharma (1982) have reported a
suite of ultrabasic and basic rocks around madaura and Sonrai. Basu (1986)
reported medium grained and dark grey gabbro associated with these ultrabasics to
the east of Madaura.
Fig. 11 Field photographs of metavolcanic suite of rocks. The dashed lines show the boundaries
of a metavolcanic rock (in each photo) intruded into the host granitic rocks. Loc. NW of Balla
Geology, Structural Architecture and Tectonic Framework … 365
Fig. 12 Field photograph of a mafic dike. The yellow line demarcates the upper and lower
boundaries of the dike. Loc. Near Girar
Barwar Formation represent the Mahrauni Group. In this work we name the
Mehroni Group as Mahrauni Group following the recent toposheets. Occurrence of
banded hematite quartzite seems to be a characteristic feature of the Barwar
Formation in the study area and so also for the Mahrauni Group at least in the
present area.
The Barwar Formation is exposed in the eastern part of the study area. Major
rock types of the Barwar Formation include banded hematite quartzite (BHQ),
conglomerate, fuchsite quartzite, grey to green chloritic schists. These rocks are
typically exposed in the Girar and Barwar areas (Fig. 13). The banded ferruginous
rocks with their parallel layering appear to be product of marine deposition. During
sedimentation, the basin also witnessed sedimentation of clastic constituents in the
form of arenaceous and politic sediments. Occurrence of chloritic schists and
quartzites indicate that the rocks were subsequently subjected to low grade meta-
morphism. Folding in the BHQ layers represents deformation of the rocks of this
Formation.
The earlier workers have considered the Mahrauni Group as part of the base-
ment. In the present area, only a small part of the Barwar Formation (Mahrauni
Group) is exposed to northeast. Both to the north (against BGC) and to the south
(against Bijawars), the Barwar Formation bears thrust contact. As such, we consider
the Barwar Formation to constitute a thrust slice.
Geology, Structural Architecture and Tectonic Framework … 367
Fig. 13 Field photograph showing outcrop of folded banded hematite quartzite of Barwar
Formation. Loc. East of Barwar
On the basis of lithological characteristics, this Group has been subdivided in the
present work into two formations; a lower Sonrai Formation and an upper Solda
Formation.
(i) Sonrai Formation
The Sonrai Formation is mostly a carbonate sequence without tuffaceous material.
The various rock types include carbonates (Fig. 14a), shales (Fig. 14b) and sand-
stones (Fig. 14c) with some basaltic and gabbroic intrusive. The rocks occasionally
show abundant bedding-controlled sulphides. Prakash et al. (1972) have subdivided
the Sonrai Formation into five members; from base to top, these members are
Jamuni Carbonate Member, Gora Kalan Shale Member, Rohini Carbonate Member,
Bandaui Sandstone Member, and Kurrat Member. However, in the study area, these
members are not fully developed. As such, it is not possible to subdivide this
formation into different members.
(ii) Solda Formation
The rocks of the Solda Formation are exposed with NW-SE to E-W trends and
occasionally show steep dips. This formation is mainly constituted of ferruginous
shale with quartz arenite interbands together with tuffaceous material and some
sandstone (Fig. 14d). The various rock types of this formation include ferruginous
shale, ferruginous quartzite, iron stone, calcareous sandstone and chloritic shale.
Prasad et al. (1999) considered this formation to represent a volcano-sedimentary
sequence.
Fig. 14 Field photographs showing outcrops of Bijawar Group of rocks. a Carbonate rocks of the
Sonrai Formation. Loc. South of Tori. b Shales of Sonrai Formation. Loc. North of Dhori Sagar.
c Sandstone of Sonrai Formation. Loc. Near Gora Kalan. d Ferruginous sandstone of Solda
Formation. Loc. Near Solda
Geology, Structural Architecture and Tectonic Framework … 369
Fig. 15 Field photographs of the rocks of the Vindhyan Supergroup (Lower Vindhyan). a Kaimur
sandstone. Loc. South of Uldana Khurd. b Fragile sandstone of Semri Group. Loc. Near
Madanpur. c Bedded sandstone of Semri Group. Loc. South of Kurrat. d Contact of Semri and
Kaimur Groups. Loc. Near Madanpur
6 Structural Architecture
(A) Structures
The large-scale structure of the rocks of the area has been observed to vary with
different lithostratigraahic units as well as, to a larger extent, on the overall phys-
iographic set-up. As mentioned above, the area can be subdivided into three
physiographic–lithostratigraphic subdivisions, we can describe the structure under
three domains as described below.
Structure of Domain-I
The Domain-I is considered here including the BGC. Structurally, the rocks of this
domain do not show any visible large-scale structure excepting the dominantly
south-dipping foliation. The rocks however occasionally show complications in
their internal domain as reflected in the mesoscopic structures described below.
Structure of Domain-II
Structurally, the central part of the study area has been identified as Domain-II. This
domain is mainly represented by the rocks of the Bijawar Group. In the study area,
the rocks of the Bijawar Group are folded into an anticline to the south and a
corresponding syncline to the north. The axes of these folds trend approximately
E-W to ENE-WSW and are slightly curved in nature with gentle plunges towards
east. These folds are traversed by a number of roughly N-S trending steep to
vertical faults of varying dimensions.
Structure of Domain-III
Domain-III is constituted of the rocks of the Vindhyan Supergroup. This lithounit
includes a massive succession of sandstones shales. Structurally, the Vindhyan
rocks show a simple dipping structure with southward dipping strata. The Vindhyan
rocks are believed to rest unconformably over the Bijawar rocks. At some places,
the Vindhyan and Bijawar strata show some effects of deformation along their
contacts in the form of some small scale folds and faults.
A prominent outcrop of Deccan Trap also occurs in the southwestern part of the
area. This outcrop trends almost N-S and is about 8 km N-S and about 4 km E-W.
In addition to this, four scattered outcrops of Deccan Trap also occur at the southern
fringe of the study area. These outcrops extend for a few tens meters only. The
Deccan Trap occurs as flat cap rocks and thus adds a special feature to the phys-
iography of the area. The Deccan Trap rocks are constituted of basaltic rocks. As
such, we consider these rocks as discordant igneous bodies and possibly represent
the last igneous (volcanic) event in the study area. The Deccan Trap rocks appar-
ently do not show any structure in their internal domain.
(B) Mesoscopic Structures
The rocks of the area show a variety of mesoscopic structures that are distributed in
the area rather sporadically. Most of these structures occur in the BGC followed by
the Bijawar Group. In the rocks of the Vindhyan Supergroup, occurrence of
Geology, Structural Architecture and Tectonic Framework … 371
Fig. 16 Geological cross-sections of the rocks of the study area in three traverses (marked in
Fig. 4) across the strike of the rock formations: T1: upper diagram, T2: middle diagram, T3: lower
diagram
372 G. K. Dinkar et al.
(i) F1 Folds
The first phase of deformation structures are represented by the F1 folds (Fig. 17a,
b). These folds are well developed in the quartzite (banded hematite quartzite) of
Mahrauni Group. In the BGC, these folds are commonly traced by silica bands of
the gneissic rocks. The F1 folds are generally tight recumbent isoclinal with
interlimb angle of about 20°–30°. They commonly show thickened hinge zones and
thin limbs. The F1 folds vary in size from a few centimeters to a few meters and are
characterized by high amplitude/wavelength ratio which is up to about 4 at some
places. The axes commonly trend between NW-SE and E-W but occasionally show
minor variations due to the effect of later deformations. In general, the higher
degree of flattening at the hinge zone is associated with higher thinning of the limbs.
The F1 folds are occasionally associated with an axial plane foliation designated as
S1 which is parallel to F1 fold axes. However, S1 transects bedding at the closures of
the F1 folds.
(ii) F2 Folds
The second generation folds (F2) are commonly developed in the banded hematite
quartzite (Fig. 17c) and in the banded hematite jasper of Barwar Formation. The F2
folds are generally coaxial to F1. In comparison with the F1 folds, the F2 folds are
more open, asymmetrical, and show low amplitude/wavelength ratio and higher
interlimb angle up to 60°. These folds generally show almost constant orthogonal
thickness in the competent folded layers.
Fig. 17 Photograph showing minor folds of the study area. a, b F1 folds in the banded hematite
quartzite affected by ductile shearing showing thinning of limb. Loc.1.8 km west of Girar. c F2
folds in banded hematite jasper, 1.8 km southwest of Girar. d F3 folds in carbonate rocks together
with ferruginous layering. Loc. 0.5 km south of Gorakalan
Geology, Structural Architecture and Tectonic Framework … 373
(iii) F3 Folds
The third generation (F3) folds (Fig. 17d) are relatively much less common in the
area. These are open folds with rounded hinges. Generally these are parallel folds
thus showing class 1B of Ramsay (1967). The amplitude/wavelength ratios are low
as compared with the F1 and F2 folds. In the schistose rocks, these folds commonly
occur as crenulation folds. The F3 folds commonly show vertical axial planes
striking NNE-SSW to NNW-SSE direction. These folds are noticed in the basic
dikes as well as in some rocks of the Bijawar Group. The F3 folds possibly rep-
resent signatures of some large scale open folds in the form of warps that developed
a curvilinear pattern to the some of the ridges of the area.
7 Tectonic Framework
Our knowledge on the tectonic framework of the rocks of the area is as yet meager.
The main reason is that the nature of contact of a few lithostratigraphic units (e.g.
Sonrai-Solda, Bijawar-Vindhyan) is either not known properly or is debatable (e.g.
Jha et al. 2012; Kumar et al. 1990; Prakash et al. 1972). In the course of our present
work, we have made efforts to study the nature of contact of important lithos-
tratigraphic units and the related structural features, if any.
(i) BGC-Bijawar contact
The contact of crystalline rocks with sedimentary/metasedimentary (cover) rocks in
any area is always of fundamental significance. The main reason, amongst others, is
that the study of the contact zone helps us to unravel the geological and palaeo-
geograhical conditions of the past. In this context, the presence of the contact of the
BGC and the Bijawars is also of great significance not only to the present area but
also to the peninsular shield in general. When we talk of the contact of basement
rocks with cover rocks, there appears to be two possibilities: the contact is either
unconformable or tectonic. In the first case, i.e. unconformable contact, commonly
the occurrence of conglomerates is expected in which the conglomerate horizon
should contain pebbles, cobbles, etc. derived from the basement rocks. In the
second case, i.e. tectonic contact, development of structural features indicating the
presence of a fault or thrust is expected.
In the study area (Fig. 4), the BGC-Bijawar contact is exposed in the northern
parts of the area with almost an E-W trend. It passes from the south of Sonrai in the
west through the south of Pisnari in the central part, and south of Girar in the
eastern part of the area. Unfortunately the direct contact of these two rock units is
exposed only at a few localities. Also, in most cases the contact zone is rather
vaguely exposed. We have carried out detailed study of the contact zone at Sonrai,
Pisnari, Tori, Barwar, Girar and their adjoining areas. So far, we have not observed
any conglomerate horizon at the contact. We thus rule out any unconformable
contact between the BGC and the Bijawars. We have instead observed the
374 G. K. Dinkar et al.
• South Lalitpur district exposes some typical rock units of the Indian peninsular
shield.The area shows three distinct physiographic features: scattered hillocks to
the northern parts, nearly continuous hills in the central parts and plateau in the
southern parts. This physiographic set-up is closely related to the geological
setup of the area. The northern part exposes crystalline-metamorphic rocks of
the Bundelkhand Gneissic Complex (BGC), the central part exposes Bijawar
Group and the southern part exposes rocks of the Vindhyan Supergroup.
• The BGC consists of a variety of granitoids, gneisses and schists together with
mylonitised granite gneisses, migmatites and amphibolite schists. Of the gran-
itoids, pink and grey granites are the dominating types. In the study area, we
have identified the following lithologic associations: gneisses designated here as
Bundelkhand Gneisses, Bundelkhand Granites that include coarse grained pink
porphyritic granite, medium grained pink porphyritic granite, grey granite,
leucogranite, aplites and pegmatites. A few NE-SW trending quartz reefs and
NW-SE trending mafic dikes are also present.
• South of the BGC, a group of schistose rocks is exposed. This group was earlier
referred to as Mehroni Group, which we rename as Mahrauni Group following
376 G. K. Dinkar et al.
recent toposheets. The Mehroni Group was divided into four Formations.
However our work reveals that in the study area only one formation, i.e. Barwar
Formation, is exposed. This Formation is constituted of banded hematite
quartzite, conglomerate, fuchsite quartzite and grey to green chlorite schists.
• South of the BGC occurs the Bijawar Group that extends almost E-W in the
central parts of the area. This group includes a sequence of sedimentary rocks
that are affected by very low metamorphism. In the study area, this group is
represented by two Formations: a Lower Sonrai Formation and an Upper Solda
Formation. The Sonrai Formation includes carbonates, shales and sandstones
with some basaltic and gabbroic intrusive while the Solda Formation is mainly
constituted of ferruginous shales with quartz arenite interbands together with
tuffaceous material and some sandstone.
• The southern part of the area is mostly occupied by the rocks of the Vindhyan
Supergroup and is represented by only two formations, viz. Semri and Kaimur
Formations.
• In addition to the above rock units, a few outcrops of Deccan Trap also occur in
the southern parts of the area. The Trap is constituted of fine to medium grained
basaltic rocks that are light to dark grey in colour.
• The structure of the area seems to be related to the individual lithostratigraphic
units. Thus we can identify three different structural domains towards south. The
Domain-I is assigned to the BGC, which apparently does not show any visible
large-scale structure excepting a dominantly south-dipping foliation. Domain-II
is represented mainly by the Bijawar Group folded into an anticline to the south
and a corresponding syncline to the north. The axes of these folds trend
approximately E-W to ENE-WSW. Domain-III mainly includes the Vindhyan
Supergroup. The rocks of this rock units show a simple dipping structure with
southward dipping strata.
• The rocks of the area show a variety of mesoscopic structures rather irregularly
distributed in different parts. We have identified a variety of minor folds, planar
structures and linear structures. Study of minor folds indicates that the rocks of
the area have undergone at least three generations of folds. The first generation
(F1) folds are commonly traced by quartz bands of the associated rocks and are
generally tight recumbent isoclinal with interlimb angle of about 20°–30°. The
second generation (F2) folds are generally co-axial to F1 but they differ in
geometry from F1 folds in being more open, asymmetrical and occasionally
greater interlimb angle up to 60°. The third generation (F3) folds are relatively
less common in the area and occur as open folds with high
wavelength-amplitude ratios.
• The tectonic framework of the rocks of the area is as yet not clearly known
mainly because the nature of contact of the lithostratigraphic units is either not
known properly or is debatable. One practical problem is that the contact of
most of the major rock units is not clearly exposed. In the light of our obser-
vations, it appears that the BGC and the Bijawars bear a thrust contact while the
Bijawars and the Vindhyans bear a faulted contact. The Mahrauni Group
Geology, Structural Architecture and Tectonic Framework … 377
appears to bear a thrust contact with the BGC to the north as well as with the
Bijawars to the south. We thus consider this group to occur as a thrust slice.
• It seems possible that the contacts of the major lithotectonic units were tec-
tonized due to tectonic disturbances in different parts of the peninsular shield
possibly due to the differences in the overall lithologic, and so rheologic,
characters of the individual lithostratigraphic units. As a result, the individual
lithostratigraphic units behaved differently to the deforming stresses. Possibly
because of this the overall tectonic framework of the lithostratigraphic units of
the study area does not fully reflect the general framework elsewhere in the
peninsular shield.
Acknowledgements We express our sincere thanks to the Head, Centre of Advanced Study in
Geology, University of Lucknow, for providing working facilities. GKD extends his thanks to the
Director, Directorate of Geology and Mining, U.P., for granting him permission to carry out this
work as part of his doctorate thesis. We thank Prof. Soumyajit Mukherjee (IIT Bombay) for his
editorial and reviewing works that significantly improved the quality of the manuscript. Mukherjee
(2019) summarizes this work.
References
Dinkar GK (2016) Deformation pattern of the rocks exposed around Sonrai, district Lalitpur, Uttar
Pradesh. Unpublished Ph.D. thesis, University of Lucknow, Lucknow
Hacket CA (1870) Geology of Gwalior and vicinity. Records of Geological Survey of India 3:33–
42
Heron AM (1935) Synopsis of the Pre-Vindhyan geology of Rajasthan, Rajputana. Transactions of
the National Institute of Science 1, 17
Jha SK, Shrivastava JP, Bhairam CL (2012) Clay mineralogical studies on Bijawars of the Sonrai
Basin: Palaeoenvironmental implecations and influences on the uranium mineralization.
Journal of Geological Society of India 79, 117–134
Jhingran AG (1958) The problem of Bundelkhand granites and gneisses. In: Presidential Address,
Proceedings of Indian Science Congress 45th session, Madras, pp 48–120
Kumar B, Srivastava RK, Jha DK, Pant NC, Bhandaru BK (1990) A revised stratigraphy of the
rocks of type area of the Bijawar Group in Central India. Indian Minerals 44, 303–314
Misra RC (1969) The Vindhyan System. Presidential Address (Geology & Geography Section)
56th Session, Bombay
Misra RC, Sharma RP (1975) New data on the geology of Bundelkhand complex of Central India.
In: Verma VK et al (ed) Recent research in geology. Hindustan PubIishing Co., Delhi, vol 2,
pp 311–346
Misra AA, Mukherjee S (2015) Tectonic inheritance in continental Rifts and Passive Margins.
Springer-Briefs in Earth Sciences. Springer International Publishing, Berlin. ISBN
978-3-319-20576-2
Misra AA, Bhattacharya G, Mukherjee S, Bose N (2014) Near N-S paleo-extension in the western
Deccan region in India: does it link strike-slip tectonics with India-seychelles rifting?
International Journal of Earth Science 103, 1645–1680
Misra AA, Sinha N, Mukherjee S (2015) Repeat ridge jumps and microcontinent separation:
insights from NE Arabian Sea. Marine and Petroleum Geology 59, 406–428
Medlicott HB (1859) Vindhyan rocks and their associates in Bundelkhand. Memoir of Geological
Survey of India II, pp 1–95
Mondal MEA, Goswami JN, Deomurari MP, Sharma KK (2002) Ion microprobe 207Pb/Pb206 ages
of Zircons from Bundelkhand massif, northern India, implications for crustal evolution of the
Bundelkhand-Aravalli Proto Continent. Precambrian Research 117, 85–110
Mondal MEA (2010) Geochemical evolution of the Archaean-paleoproterozoic Bundelkhand
Craton, Central India shield. Journal of Economic Geology and Georesource Management 7,
69–80
Mukherjee S, Misra AA, Calvès G, Nemčok M (2017) Tectonics of the Deccan large igneous
province: an introduction. In: Mukherjee S, Misra AA, Calvès G, Nemčok M (eds) Tectonics
of the Deccan large igneous province. Geological Society, London, Special Publication 445,
pp 1–9
Mukherjee S (2011) Mineral fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth sciences 100, 1303–1314
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Pascoe EH (1950) A manual of the Geology of India and Burma. Geological Survey of lndia,
Calcutta, 1
Pati JK, Patel SC, Pruseth KL, Malviya VP, Arima M, Raju S, Pati P, Prakash K (2007) Geology
and geochemistry of gaint quartz veins from the Bundelkhand Craton, Central India and their
implications. Journal of Earth System Science 116, 497–510
Prakash R, Srivastava RN, Saharma DP, Bhatt GD (1972) Progress of geological investigation on
Sonarai copper belt, Mehroni Tehsil, Jhansi, Uttar Pradesh. DGM, U.P.
Prakash R, Swarup P, Srivastava RN (1975) Geology and mineralization in the southern parts of
Bundelkhand in Lalitpur district, Uttar Pradesh. Journal of Geological Society of India 16,
143–156
Geology, Structural Architecture and Tectonic Framework … 379
Prasad MH, Hakim A, Rao BK (1999) Metavolcanic and Metasedimentary Inclusions in the
Bundelkhand Granitic Complex in Tikamgarh District, Madhya Pradesh. Journal of Geological
Society of India 54, 359–368
Radhakrishna BV (1989) Suspect Tectono-stratigraphic terrain elements in the Indian
Subcontinent. Journal of Geological Society of India 34, 1–24
Ramsay JG (1967) Folding and fracturing of rocks. McGraw-Hill, New York
Roday PP, Diwan P, Singh S (1995) A kinematic model of emplacement of quartz reef and
subsequent deformation pattern in Central Indian Bundelkhand batholiths. Proceedings of the
Indian Academy of Science—Earth and Planetary Science 104, 465–488
Sarkar A, BhaIIa JK, Bishui PK, Gupta SN, Srimal N (1989) Geochemistry and geochronology of
the Early Proterozoic Bundelkhand granitic complex, Central India. In: Symposium on
Precambrian, Granitoids, Helsinki, Finland, Geological Survey of Finland, p 117 (Abstract 8)
Sarkar A, Paul DK, Potts PJ (1995) Geochronology and geochemistry of the Mid-Archean,
Torndhjemitic gneisses from the Bundelkhand Craton, Central India. In: Saha AK (ed) Recent
researches in geology, vol 16, pp 76–92
Sarkar A, Ghosh S, Singhsi RK, Gupta SN (1997). Rb-Sr geochronology of the Dargawan sill:
constraint on the age of the type Bijawar sequence of Central India. In: International conference
on isotopes in solar system, pp 100–101 (Abstract pp 11–14)
Saxena MN (1961) Bundelkhand granites and associated rocks from Kabrai and Mauranipur area
of Harnirpur and Jhasi district; U.P., India. Research Bulletin of Panjab University 12, 85–107
Sharma RP (1982) Lithostratigraphy, structure and petrology of Bundelkhand Group. In:
Valdia KS, Bhatia SB, Gaur VK (eds) Geology of Vindhyanchal. Hindustan Publishing
Corporation, pp 30–46
Singh SP (2012) Geochemical signatures of Archaean felsic volcanism in Central part of
Bundelkhand Craton, Central India. International Journal of Advance Earth Sciences 1, 20–32
Singh SP, Bhattacharya AR (2017) N-S crustal shear system in the Bundelkhand massif: a unique
crustal evolution signature in the northern Indian peninsula. Journal of Earth System Science
126, 121
Singh SP, Bhattacharya AR (2010) Signature of Archaean E-W crustal-scale shears in the
Bundelkhand massif, Central India, an example of vertical ductile shearing. Earth Science India
3, 217–225
Singh SP, Singh MM, Srivastava GS, Basu AK (2007) Crustal evolution in Bundelkhand area,
Central India. Journal of Himalayan Geology 28, 79–101
Singh SP, Dwivedi SB (2009) Garnet-silliminite-cordierite-quartz bearing assemblages from the
early Archaean supracrustal rocks of Bundelkhand massif, Central India. Current Science 97,
103–107
Srivastava JP (1951–1952) Rocks of Probable Bijawar age overlying the Bundelkhand granite in
Tikamgarh district. Geological Survey of India, Unpublished Progress Report
Swaroop P, Saxena AK (1969) A report on the preliminary investigation of iron-formations near
Obra, District Mirzapur, Uttar Pradesh. Report of the Directorate of Geology and Mining, U.P.
Deformation in the Kangra Reentrant,
Himachal Pradesh of NW-Sub Himalaya
of India: A Paradox
1 Introduction
The Himalayan orogeny is geologically very complex and heterogeneous along its
entire length even though it consists of longitudinal curvilinear stripes of similar
lithologies (Mukherjee 2013a). For the ease of correlation and understanding, broad
sub-divisions have been made based on comparable litho-tectonic units (Fig. 1).
These sub-divisions have been further verified through detailed lab (geochemical)
and field studies (Valdiya 1984). Accordingly, the southern (outer)most
litho-tectonic sub-division of the Himalaya is called the Sub-Himalayan belt. The
Sub-Himalayan belt in the NW Himalaya has been studied in detail and classified
based on energy sequences, which are equivalent to various rock types (Raiverman
2002). This belt is the youngest and most actively deforming tectonic sub-division
of the Himalayan orogen that accommodates about 20% (*4–11 mm/yr) of the
total India-Asia convergence (Larson et al. 1999; Lave and Avouac 2000; Bilham
et al. 2001). Note that the Sub-Himalayan belt is assumed to be tectonically col-
located with the external locked zone of the Himalayas (Bilham et al. 2001; Stevens
and Avouac 2015). This locked zone has an approximate width of 100 km and
tends to accumulate the potential tectonic slip almost entirely as elastic (>90%)
rather than inelastic (<10%) strain, which would permanently deform the rock
T. Singh (&)
CSIR-Central Scientific Instruments Organisation, Chandigarh 160030, India
e-mail: tejpal@csio.res.in; geotejpal@yahoo.co.in
A. K. Awasthi
Graphic Era (Deemed to be University), Dehradun 248002, India
A. K. Awasthi
Formerly at Department of Earth Sciences, Indian Institute of Technology
Roorkee, Roorkee 247667, India
(Bilham et al. 2001). The cover sequences in this locked zone are variably coupled
with the basement and the coupling tends to maximum near the deformation front
i.e. Himalayan Frontal Thrust: HFT (Stevens and Avouac 2015).
The Sub-Himalayan belt is a typical example of a “cover sequence”, which is
faulted and folded over a rigid basement. The structures characterize the thin skin
deformation of Fold-Thrust belts (FTBs). There has been a considerable progress in
our understanding of thin-skin deformation of FTBs over the last few decades (Ruh
et al. 2012) and as more and more data and new tools become available, refined
interpretations are coming up (e.g., Rajendra Prasad et al. 2011; Carrillo et al.
2017). The initial models are commonly based on surface geology and structure.
However, the current understanding largely draws from the theoretical and exper-
imental models (Suppe 1983; Suppe and Medwedeff 1990; Stokmal et al. 2007).
Most of these models are based on (i) different material properties of the detach-
ment e.g., frictional vs. viscous detachments (Cotton and Koyi 2000; Bahroudi and
Koyi 2003), (ii) inherent properties of the cover sequences such as cohesion and
pore pressure, and (iii) interaction between the detachment and cover sequences
such as their coupling (Withjack and Callaway 2000; Paul and Mitra 2015;
Borderie et al. 2018). The results are commonly interpreted in terms of evolution of
(i) structural geometries such as imbricate thrusts, back thrusts, open folds etc.,
(ii) topographic slope (a), and (iii) detachment dip (b) (Singh et al. 2012). In most
cases, the models are tuned compatible with the available seismic sections, well
logs, structural- and topographic maps.
The width of the Sub-Himalayan belt varies along the length of the orogen. It
reaches to a maximum of over 300 km in the western sections of Pakistan whereas
Fig. 1 Broad litho-tectonic subdivisions of the ‘Himalaya’ (after Valdiya 1984, 1998)
Deformation in the Kangra Reentrant, Himachal Pradesh … 383
it thins to <10 km in the eastern sections, especially east of Nepal. The variable
width to a large extent is attributed to the original basin geometry and rheology of
the basement. The basins of deposition towards the west have been extremely wide
and sediments have been deformed over rheologically weak basement in salt (Davis
and Lillie 1994). This has not been the case eastwards where, the rheologically
weak salt is not found in the basement and the belts are narrower. However, there is
an interesting paradox in the Kanrga Reentrant of Western India where the width of
the Sub-Himalayan belt is *80 km: much wider than that at east. The topographic
slope (a) of the Kangra Reenetrant is *1° and together with the basement dip (b)
of 3°, it forms a wedge taper of around 4°. This wedge taper is comparable to those
of Salt Range-Potwar Plateau in Pakistan (Davis and Lillie 1994) and Fars and
Lorestan segments of the Zagros FTB in Iran (McQuarrie 2004). Notably the FTBs
in Pakistan and Iran are extremely wide and have also evolved over a weak
detachment in salt which justifies a low wedge taper. However, there have been no
reports on the presence of salt at the surface or sub-surface in the Kangra Reentrant
which could justify a relatively low wedge taper spread over a width of 80 km. This
article focuses on this paradoxical reentrant to elaborate on the geological structure
and to understand the reasons for its relatively larger width.
2 Litho-Tectonic Setting
The Sub-Himalayan belt, in general comprises the Tertiary rock units, which are
classified into the Lower and the Upper Tertiaries. In western India, it consists of
parautochthonous and the autochthonous units (Table 1). The former units include
the Subathu and the Dagshai (=Lower Dharamshala) and the Kasauli (=Upper
Dharamshala) Formation of Lower Tertiary age. The autochthonous unit includes
the Siwalik Supergroup of Upper Tertiary to Quaternary age (Raiverman 2002).
The Siwalik Supergroup is conventionally sub-divided into the Lower, Middle and
Upper Siwalik Groups (Table 1). Both parautochthonous and autochthonous units
are unconformably overlain by Quaternary terraces along the stream network
(Singh et al. 2011). The Sub-Himalayan rock units thrust over the sediments of the
Indo-Gangetic plains along the Himalayan Frontal Thrust (HFT). The HFT defines
the present-day tectonic boundary of active tectonic convergence (Yeats et al. 1992;
Powers et al. 1998; Delcaillau et al. 2006; Kumar et al. 2006; Singh and Jain 2009).
Deformation style and kinematics vary considerably across the fold-thrust belts in
orogenic systems (e.g., Mukherjee 2013b). These differences may be present at
much finer scales but their first order manifestation is readily visible in the surface
topography as seen in various modeling approaches (Koyi 1988; Cotton and
384
Fig. 2 Generalized surface topography of the Kangra Reentrant showing NW-SE elongated
ridges separated by synclinal valleys. The main structures viz. HFT: Himalayan Frontal Thrust,
JbT: Januari back Thrust, ST: Soan Thrust, BbT: Barsar back Thrust, JMT: Jwalamukhi Thrust,
MBT: Main Boundary Thrust and MCT: Main Central Thrust are shown. The representative
anticlinal and synclinal topography are marked. The star in the top middle panel represents the last
major earthquake of 1905 around Kangra. The line of structural cross-section A–B shows location
of Fig. 3
Koyi 2000; Costa and Vendeville 2002; Nilforoushan et al. 2008; Nilfouroushan
et al. 2012; Ruh et al. 2012). In all these models, structural evolution readily
compare with the large scale topographic features. Of course, surface processes and
erosion play an equally important role in shaping the topography. Their role is to
enhance the surface manifestations of geological structure, often modulated by
lithology and climate (Singh 2008a, b; Singh and Jain 2009). Therefore surface
topography forms a primary guide to the mechanical characters leading to the
observed structural geometries and surface manifestations.
The present day surface topography of the Kangra Reentrant is characterized by
a series of NW trending (sub)parallel ridges separated by wide and relatively flat
valleys trending parallel to the ridges in between them (Fig. 2). This topographic
arrangement is more apparent towards the southern part of the Fold-Thrust Belt
(FTB) between the Himalayan Frontal Thrust (HFT) and the Jwalamukhi Thrust
(JMT; Fig. 2). Here the ridges are more simply NW-SE elongated almost parallel to
the HFT. North of JMT, although the same arrangement of ridges exists there is a
curvature which most likely replicates the strike of the MBT. These form a different
386 T. Singh and A. K. Awasthi
surface expression between the JMT and MBT that is more subdued by the cover of
Quaternary alluvial fills and fans descending from the adjoining mountain ranges.
Also, the surface expression in this part is complicated by the interaction of tectonic
and surface processes strongly modulated by the climatic signal (Dey et al. 2016).
The JMT forms an important tectonic boundary which internally constrains the
Sub-Himalayan wedge, although its activity has been identified only through
neotectonic studies and terrace dating (Singh et al. 2012; Mukherjee 2015; Dey
et al. 2016).
The most important surface topographic parameter that contributes to the taper
of the Sub-Himalayan wedge is the forelandward slope of the modern day topo-
graphic envelope (a). For the Kangra reentrant, the present day topographic slope a
has been estimated to be 0.9° ± 0.2° based on analysis of topographic maps and
Digital Elevation Models (Singh et al. 2012). This too small alpha is comparable to
many other wedges developed in the paleo-forelands of collisional settings over
weak detachments in the Salt Range of Pakistan (Davis and Lillie 1994) and the
Zagros FTB of Iran (McQuarrie 2004).
4 Detachment Geometry
Although uncertainties exist about the geometry and rheology of the detachment
(Main Himalaya Thrust: MHT) in the NW Himalaya, it is generally considered to
be a gently dipping single continuous sliding plane over which the cover sequences
deform (Fig. 3). The dip of the detachment has been estimated *3°in the external
sector using well logs and seismic profiling carried out by the ONGC (Powers
et al. 1998). The dip is more or less uniform under the entire Sub-Himalayan belt
of Kangra reentrant. This is particularly true because any undulations in the
geometry of the detachment would locally induce stress concentrations causing
clustering of seismicity, which contradicts the seismic findings (Arora et al. 2009).
However, in the down dip direction further north of the MBT, the dip of the
detachment increases to about 10°–15°. This corresponds with the zone of intense
microseismicity that defines the Himalayan mid-crustal ramp just north of the Main
Central Thrust (MCT) (Figs. 1 and 2). Mechanically the basement is a rigid mass
made-up of the Vindhayan rocks equivalent and followed by the metamorphic/
granitic basement at deeper levels (Powers et al. 1998; Rajendra Prasad et al.
2011). The internal thrusts formed within the Sub-Himalayan belt are the result of
continued contraction tectonics of the Himalaya (Singh et al. 2014). They emerge
from the basement and deform the cover sequences into broad open folds (Powers
et al. 1998). The anticlinal folds manifest in the surface topography as ridges
followed by trailing synclines manifested as wide flat intermontane depressions
(Fig. 2).
Deformation in the Kangra Reentrant, Himachal Pradesh … 387
5 Cover Sequences
The overall cross-sectional taper of the wedge is intricately related to the compe-
tency contrast of the rocks in the wedge and its basal detachment (Davis et al.
1983). In the Kangra reentrant the basal detachment is resolved to be a rigid surface
made of competent rocks. However, there are outstanding issues regarding the
extent of the resolution and role of the cover rocks involved in deformation. The
frictional strengths of a wide variety of rocks in the upper crust are similar (Davis
and Engelder 1985). However, there are special considerations for clay and shales,
which are relatively weak and their presence can significantly modify the defor-
mation style if they are predominant in the cover sequence (Davis and Lillie 1994).
They tend to behave similar to weak horizons as in the case of eastern Potwar
plateau (Pakistan), which exhibits increased sedimentation rates of 0.46–0.53 m/kyr
during the Middle Siwalik sedimentation towards the top of the sequence. The
sequence is predominantly sandy with interbedded clays of the Middle Siwaliks
(Pennock et al. 1989).
The cover sequences of the Sub-Himalayan FTB deform over a rigid basement
forming foreland verging thrusts, fault-related folds and back-thrusts. The structural
geometries are constrained through data from surface geology, 2D-seismic profiles,
well logs and other ancillary data from geophysical techniques (Figs. 2 and 3;
Powers et al. 1998; Mukhopadhyay and Mishra 1999; Jayangondaperumal et al.
2017). It is important to note the major thrusts towards from S to N are Himalayan
Fig. 3 Top panel shows a generalized structural section across the Sub-Himalaya of Kangra
Reentrant (modified from Powers et al. 1998). For section line, refer to Fig. 2. Vertical and
horizontal scales are the same. The bold line forming the base of the section represents the
basement. The internal thrusts shown in the section emerge from the basement and deform the
cover sequences into broad open folds (bottom panel). The bottom panel details the triangle zone
structure formed between the Soan Thrust (ST) and the Barsar back Thrust (BbT). Regional drag
folds (Mukherjee 2014) are also exemplified in this cross-section
388 T. Singh and A. K. Awasthi
Frontal Thrust (HFT), Soan Thrust (ST), Jwalamukhi Thrust (JWT), Palampur
Thrust (PT) and the Main Boundary Thrust (MBT). The geometry of the folds i.e.
Janauri Anticline (JA), Dera-Gopipur Anticline (DGA), Bahl Anticline (BA) and
Paror Anticline (PA), is clearly attributed to the movement on associated thrust
faults (Fig. 3). The spacing between the thrusts has controlled the development and
preservation of fault-related folds in their hanging walls. The main back thrusts are
Janauri back Thrust (JbT) and Barsar back Thrust (BbT).
In the tectonic wedge of the Kangra reentrant, the sedimentary cover rocks have
been thrust southward over a series of thrust faults (Powers et al. 1998;
Mukhopadhayay and Mishra 1999; Singh et al. 2012). There is almost no evidence
of basement rocks involved in thrusting (Dubey 2004). The vergence of main
structural elements i.e. HFT, ST, JT, PT and MBT is southward, towards the
foreland. However, structures with opposite vergence are also common towards the
frontal region e.g. JbT, BbT. The thrusts and back thrusts show active deformation
consistent with the model of distributed deformation (Singh et al. 2012; Thakur
et al. 2014; Dey et al. 2016; Jayangondaperumal et al. 2017; also Bose and
Mukherjee, submitted-1, 2). The Tertiary formations thicken towards the northeast
and the structural style becomes increasingly more complex towards the hinterland,
as compared over the *80 km length of the Sub-Himalayan wedge from SW
towards NE (Powers et al. 1998; Mukhopadhyay and Mishra 1999).
There are a series of research contributions on the active tectonic deformation along
thrust faults in the Kangra reentrant (Malik and Mohanty 2007; Singh et al. 2012;
Thakur et al. 2014; Dey et al. 2016). The last major earthquake in this area occurred
in 1905 (M = 7.8) near Kangra (Fig. 2). As this happened before the beginning of
instrumental monitoring, there are no good records on the precise location, nature
and extent of the rupture zone. However, there have been outstanding debates with
few equivocal surface manifestations found (Szeliga and Bilham 2017). The 1905
Kangra earthquake is generally attributed to the recent slip along the Main
Boundary Thrust (MBT). In the instrumental era since late 1950s when seismic
monitoring through instruments began, there have been no major earthquakes in
this area and most of the recorded smaller events are shallow (<10 km) with 3–5
magnitudes. Such seismic behaviour would also not suggest any involvement of the
basement (Mukhopadhyay and Dasgupta 2015). Geologically this would mean that
the cover rocks are poorly coupled with the basement which has a low estimated
range of basal friction (µb) i.e. 0.26–0.36 in the Kangra Reentrant (Singh et al.
2012). Also range-normal convergence velocities across the Sub-Himalayan belt
demonstrate reduced intensities connoting poor tectonic coupling between the cover
and the basement (Schiffman et al. 2013). However, in the Kangra Reentrant it can
Deformation in the Kangra Reentrant, Himachal Pradesh … 389
7 Discussions
The Kangra Reenetrant exhibits a topographic slope (a) of *1° towards SW and a
relatively high basement dip (b) of 3° towards NE (Fig. 4; in comparison to FTBs
in Pakistan and Iran). The two parameters (a and b) contribute to the taper of the
wedge, which is *4°. This wedge taper is slightly higher than the Salt
Range-Potwar Plateau (Pakistan) where it ranges 1°–3° (Davis and Lillie 1994) and
Fars and Lorestan segments of the Zagros FTB (Iran) where it is 1.5° (McQuarrie
2004). The FTB’s in Pakistan and Iran are comparable in age to the Sub-Himalayan
belt but it is notable that both the FTBs in Pakistan and Iran have evolved over a
weak detachment in salt. In fact, the wedge taper in the Kangra Reentrant is
comparable to the Dezful embayment of the Zagros FTB of Iran which is around 4°
(McQuarrie 2004). The most likely reason for the increased wedge taper of the
Dezful embayment is its limited width of the FTB, which is controlled by the spatial
distribution of the Hormuz salt (Mukherjee et al. 2010). Overall, for the Kangra
Reentrant the high basement dip b is supporting a gentler topographic slope a
(Davis et al. 1983; Dahlen 1990; Fig. 4). What is intriguing is that the topographic
slope is similar to that observed over the weak basement in Pakistan and elsewhere
(Fig. 4), whereas there is no report of such a basement rheology in the Kangra
Reentrant. The topographic slope is extremely low (*1°, almost similar to the
FTBs in Pakistan and Iran) and forms a relatively narrow wedge taper over a width
of about 80 km (largest in the Indian part). Such characters of a wedge which slides
over a rigid basement give rise to a paradoxical situation that has been dealt in detail
based on a synthesis of data presented in the earlier Sects. 3, 4, 5 and 6.
In general, assuming a particular rheology of the basement, the other important
factors controlling the growth of the FTB and surface topographic slope (a) could
be the mechanical stratigraphy, pore-fluid pressure of the cover rocks and climatic
effects. Also in a particular area, under similar climatic conditions competent rocks
would be able to retain a higher surface topography whereas incompetent rocks
would not. Therefore in the Kangra Reentrant the basement is assumed to be a rigid
competent mass over which the total thickness of incompetent clay-rich Paleogene–
Neogene section is 4–8 km. This thickness is higher than similar sediments in the
Potwar plateau (Davis and Lillie 1994). So, if a similar but thinner sediment layer
could contribute to the weak coupling between the basement and cover sequence
(Davis and Lillie 1994), it would be reasonable to believe that these horizons would
definitely provide a reduced resistance to slippage along faults. Also these sedi-
ments would effectively couple weakly between the basement and cover rocks in
the Kangra reentrant as well (Davis and Engelder 1985). In principle, this situation
390 T. Singh and A. K. Awasthi
Fig. 4 Surface slope a versus dip of detachment b for natural FTBs and Delta-Deepwater FTBs
(modified from Ford 2004; Tuitt et al. 2012)
is also analogous to many deep water FTBs, which contain detachments in shale
(e.g. Niger delta; Fig. 4). In addition, the cover strata of the Kangra Reentrant are
relatively overpressured (up to 1.5–2 times more than the hydrostatic pressure cf.
Powers et al. 1998). This is also supported by high sedimentation rates in this
oversized basin (Rajendra Prasad et al. 2011). Therefore the high pore-fluid pres-
sures would efficiently act to reduce the effective normal stress in turn reducing the
shear traction and contribute to the reduction in coupling between the cover strata
and the basement. This is especially true and important because the presence of
high pore-fluid pressure shales would act as an efficient detachment, although not as
efficient as halite (Kreuger and Gilbert, 2009; also see Dasgupta and Mukherjee
submitted). This would facilitate in generating a wider FTB such as in the Kangra
Reentrant, similar to those that are formed also see over weak detachments in salt
and over pressured shales elsewhere (Fig. 4).
The most interesting observation is that the strength of the wedge (its frictional
character), which could be either due to the rheology or other characters such as
pore pressure would strongly control the orientation of the principal stress axes.
Accordingly, an increased pore-fluid pressure would reduce the internal friction of
the wedge (µ). This in-turn would decrease Wb as demonstrated by the
Mohr-Coulomb criterion (Fig. 5). Wb strongly controls the orientation of the
principal stress axes and the style of deformation as illustrated in Fig. 6.
Deformation in the Kangra Reentrant, Himachal Pradesh … 391
Fig. 5 Mohr-Coulomb
failure criterion showing how
different values of Wb are
related to the internal friction
(µ) of the wedge. With the
reduction of internal friction,
the Wb angle also reduces. In
particular for wedges with
basement in salt Wb would be
approximately 1° (modified
after Davis and Engelder,
1985). The s and r represent
shear and normal stress
respectively
The normal (r) and shear stresses (s) acting on a plane are evaluated using the
Mohr-Coulomb failure criterion (Fig. 5). The normal to the plane is oriented at an
angle Ɵ with the maximum principal stress, r1 direction. Accordingly, there are two
planes oriented at an angle inclined on either side of the r1 axis. For frictional wedges
392 T. Singh and A. K. Awasthi
the angle Wb offsets the orientation of r1 in such a way that the r1 axis is highly
inclined with respect to the detachment (Fig. 6a). In this case, the two slip planes
formed have different orientations (vergence) with reference to the slope of the
topographic wedge. The foreland verging plane, the fore-thrust, usually dips gently
towards the hinterland whereas the hinterland verging plane, the back-thrust, dips
steeply towards the foreland (Fig. 6a). Geometrically, the fore-thrust would be the
preferred slip plane. Thus, these FTBs should have a preferred structural vergence.
On the contrary, for weaker wedges, Wb tends to be quite small (Fig. 5) leading to a
near horizontal r1. In such a situation there could be two slip planes on either side of the
r1, similar to the frictional wedges. But these planes orient in such a way that they make
almost equal angle with the surface topography. The two planes in this case also are
foreland-verging fore-thrust and hinterland verging back-thrust but unlike the fric-
tional wedges they have comparable dips. Therefore, both these planes i.e. the
fore-thrust and the back-thrust would be equally preferred slip planes but with opposite
vergence (Fig. 6b). The FTBs on such wedges have no preferred structural vergence as
is the case with the Kangra Reentrant (Figs. 2 and 3). This is very well illustrated along
the Una–Talmehra section exhibiting a triangle zone between the ST (fore-thrust) and
the BbT (back-thrust) where the ST and BbT are separated by *10 km on the surface
(Fig. 3). Therefore, wedges with incompetent rocks in the cover strata and poor cou-
pling between the basement and cover strata would allow the wedge to translate over a
wide belt usually project outward (Figs. 2 and 3). Similar inferences have been drawn
from the comparisons between the detachment strength to wedge strength (Stockmal
1983; Davis and Engelder 1985). As the wedge in the Kangra Reentrant would rapidly
propagate forward it would incorporate the undeformed foreland into the wedge,
towards the frontal part and spatially widen the wedge over the basement. This should
significantly lead to the decrease in topographic slope (a) over time and reduce the taper
of the wedge (Singh et al. 2012), as seen in the case of Kangra Reentrant.
In summary, the surface morphology of the Kangra Reentrant resembles
deformation that is distributed in nature and exhibits poor coupling between the
basement and cover units. Other attributes include: (i) presence of wide-spaced
folds, (ii) abrupt changes in deformation style across the basin that usually corre-
spond spatially to the weak layer distribution, (iii) narrow cross-sectional taper, and
(iv) bimodal vergence of structures within the FTB. The structural evolution of the
Kangra Reentrant is mainly by development of fault-related folds in the cover strata
over a rigid basement. It is well understood that the basement structure and asso-
ciated variations in stratigraphic thickness of cover rocks are potential controls for
the localization and growth of frontal anticlines (McQuarrie 2004).
8 Conclusions
The structure of the Kangra reentrant is controlled by the geometry of the sedi-
mentary basin (depocentre) that was present in front of the MBT, which largely
follows the distribution of the thin Vindhyan basement as limited by the recognition
Deformation in the Kangra Reentrant, Himachal Pradesh … 393
of an older fabric beneath by Rajendra Prasad et al. (2011). This provides a larger
accommodation space occupied by thick (4–8 km) incompetent clay rich
Paleogene-Neogene section. Therefore the distribution of the excess pore fluid
pressures at the depth appears to be stratigraphy-controlled.
The poor coupling between the cover strata together with high pore pressure and
presence of thick incompetent rocks will not allow potential accumulation of shear
stresses within the rocks of the Kangra area, which is characterized by the shallow
and low seismicity (3–5 M) in the region since the last major event of 1905. In most
cases there in no involvement of the basement in the India-Eurasia collision induced
deformation process of the Sub-Himalayas in the Kangra Reentrant. This indicates
that most of these earthquakes would occur by a dynamic weakening mechanism
due to overpressured shales. However, further north of the JMT there are potential
seismogenic zones, which are actively accumulating shear stresses over narrow
localized zones. This may lead to a high magnitude seismicity, similar to the 1905
event. However, the chances of the rupture propagating towards the HFT seem to
be limited.
References
Arora BR, Gahalaut VK, Kumar N (2009) Structural control on along-strike variation in the
seismicity of the northwest Himalaya. Journal of Asian Earth Sciences 57, 15–24
Bahroudi A, Koyi HA (2003) Effect of spatial distribution of Hormuz salt on deformation style in
the Zagros fold and thrust belt: an analogue modelling approach. Journal of the Geological
Society 160, 719–733
Bilham R, Gaur V, Molnar P (2001) Himalayan seismic hazard. Science 293, 1442–1444
Borderie S, Graveleau F, Witt C, Vendeville BC (2018) Impact of an interbedded viscous
décollement on the structural and kinematic coupling in fold-and-thrust belts: insights from
analogue modeling. Tectonophysics 722, 118–137
Bose N, Mukherjee S (Submitted-1) Field documentation and genesis of the back-structures from a
part of the Garhwal Lesser Himalaya. Uttarakhand, India. Tectonic implications. In: Sharma,
Villa IM, Kumar S (eds) Crustal architecture and evolution of the Himalaya-Karakoram-Tibet
Orogen. Geological Society of London Special Publications
Bose N, Mukherjee S (Submitted-2) Field documentation and genesis of back-structures from the
foreland part of a collisional orogen: Examples from the Darjeeling-Sikkim Lesser Himalaya,
Sikkim, India. Gondwana Research
Carrillo E, Koyi HA, Nilfouroushan H (2017) Structural significance of an evaporite formation
with lateral stratigraphic heterogeneities (Southeastern Pyrenean Basin, NE Spain). Marine and
Petroleum Geology 86, 1310–1326
Costa E, Vendeville BC (2002) Experimental insights on the geometry and kinematics of
fold-and-thrust belts above weak, viscous evaporitic decollement. Journal of Structural
Geology 24, 1729–1739
394 T. Singh and A. K. Awasthi
Cotton JT, Koyi HA (2000) Modeling of thrust fronts above ductile and frictional detachments:
application to structures in the Salt Range and Potwar Plateau, Pakistan. Geological Society of
America Bulletin 112, 351–363
Dahlen FA (1990) Critical taper model of fold-and-thrust belts and accretionary wedges. Annual
Reviews of Earth and Planetary Sciences 18, 55–99
Dasgupta T, Mukherjee S (submitted) Sediment compaction and applications in petroleum
geoscience. In: Swenner R (ed) Springer series: advances in oil and gas exploration &
production. ISSN: 2509-372X
Davis D, Engelder T (1985) The role of salt in fold-and-thrust belts. Tectonophysics 119, 67–88
Davis D, Lillie RJ (1994) Changing mechanical response during continental collision: active
examples from the foreland thrust belts of Pakistan. Journal of Structural Geology 16, 21–34
Davis D, Suppe J, Dahlen FA (1983) Mechanics of fold-and-thrust belts and accretionary wedges.
Journal Geophysical Research 88, 1153–1172
Delcaillau B, Carozza J-M, Laville E (2006) Recent fold growth and drainage development: the
Janauri and Chandigarh anticlines in the Siwalik foothills, northwest India. Geomorphology
76, 241–256
Dey et al (2016) Climate-driven sediment aggradation and incision since the late Pleistocene in the
NW Himalaya, India. Earth and Planetary Science Letters 449, 321–331
Dubey AK (2004) Formation of decollement upwarps during thrusting. Terra Nova 16, 91–94
Ford M (2004) Depositional wedge tops: interaction between low basal friction external orogenic
wedges and flexural foreland basins. Basin Research 16, 361–375
Jayangondaperumal R, Kumahara Y, Thakur VC, Kumar A, Srivastava P, Dubey S, Joevivek V,
Dubey AK (2017) Great earthquake surface ruptures along backthrust of the Janauri anticline,
NW Himalaya. Journal of Asian Earth Sciences 133, 89–101
Koyi H (1988) Experimental modeling of role of gravity and lateral shortening in Zagros mountain
belt. AAPG Bulletin 72, 1381–1394
Krueger A, Gilbert E (2009) Deepwater fold-thrust belts: not all the beasts are equal. In: American
association of petroleum geologists international conference and exhibition. Cape Town, South
Africa. AAPG Search and Discovery. Article 30085: http://www.searchanddiscovery.com/
documents/2009/30085krueger
Kumar S, Wesnousky SG, Rockwell TK, Briggs RW, Thakur VC, Jayangondaperumal R (2006)
Paleoseismic evidence of great surface rupture earthquakes along the Indian Himalaya. Journal
Geophysical Research 111, B03304
Larson K, Burgmann R, Bilham R, Freymueller JT (1999) Kinematics of the India-Eurasia
collision zone from GPS measurements. Journal Geophysical Research 104, 1077–1093
Lavé J, Avouac JP (2000) Active folding of fluvial terraces across the Siwalik Hills, Himalayas of
central Nepal. Journal Geophysical Research 105, 5735–5770
Malik J, Mohanty C (2007) Active tectonic influence on the evolution of drainage and landscape:
geomorphic signatures from frontal and hinterland areas along the Northwestern Himalaya,
India. Journal of Asian Earth Sciences 29, 604–618
McQuarrie N (2004) Crustal scale geometry of the Zagros fold–thrust belt, Iran. Journal of
Structural Geology 26, 519–535
Mukherjee S (2013a) Channel flow extrusion model to constrain dynamic viscosity and Prandtl
number of the Higher Himalayan Shear Zone. International Journal of Earth Sciences 102,
1811–1835
Mukherjee S (2013b) Higher Himalaya in the Bhagirathi section (NW Himalaya, India): its
structures, backthrusts and extrusion mechanism by both channel flow and critical taper
mechanisms. International Journal of Earth Sciences 102, 1851–1870
Mukherjee S (2014) Review of flanking structures in meso- and micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2015) A review on out-of-sequence deformation in the Himalaya. In: Mukherjee S,
Carosi R, van der Beek P, Mukherjee BK, Robinson D (eds) Tectonics of the Himalaya,
vol 412. Geological Society, London, Special Publications, pp 67–109
Deformation in the Kangra Reentrant, Himachal Pradesh … 395
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Talbot CJ, Koyi HA (2010) Viscosity estimates of salt in the Hormuz and Namakdan
salt diapirs, Persian Gulf. Geological Magazine 147, 497–507
Mukhopadhayay DK, Mishra P (1999) A balanced cross section across the Himalayan foreland
belt, the Punjab and Himachal foothills: a reinterpretation of structural styles and evolution.
Proceedings of Indian Academy of Science 108, 189–205
Mukhopadhyay B, Dasgupta S (2015) Seismic hazard assessment of Kashmir and Kangra valley
region, Western Himalaya, India. Geomatics, Natural Hazards and Risk 6, 149–183
Nilforoushan F, Koyi H, Swantesson J, Talbot CJ (2008) Effect of basal friction on surface and
volumetric strain in models of convergent settings measured by laser scanner. Journal of
Structural Geology 30, 366–379
Nilfouroushan F, Pysklywec R, Cruden A (2012) Sensitivity analysis of numerical scaled models
of fold-and-thrust belts to granular material cohesion variation and comparison with analog
experiments. Tectonophysics 526–529, 196–206
Paul D, Mitra S (2015) Role of salt in decoupling deformation and fault evolution in the Smørbukk
area, Halten Terrace, offshore Mid-Norway. In: Annual convention and exhibition. Denver,
Colorado, 31 May–3 June 2015
Pennock ES, Lillie RJ, Hamid Zaman AS, Yousaf M (1989) Structural interpretation of seismic
reflection data from eastern salt range and Potwar Plateau, Pakistan. AAPG Bulletin 73, 841–
857
Powers PM, Lillie RJ, Yeats RS (1998) Structure and shortening of the Kangra and Dehra Dun
Reentrants, sub-Himalaya, India. Geological Society of America Bulletin 110, 1010–1027
Raiverman V (2002) Foreland sedimentation in Himalayan TectonicRegime. A relook at the
orogenic process, Bishen Singh Mahendra Pal Singh Publisher, Dehra Dun, India, p 371
Rajendra Prasad B, Klemperer SL, VijayaRao V, Tewari HC, Khare P (2011) Crustal structure
beneath the sub-Himalayan fold-thrustbelt, Kangra recess, north west India, from seismic
reflection profiling: implications for Late Paleoproterozoicorogenesis and modern Earth hazard.
Earth and Planetary Science Letters 308, 218–228
Ruh JB, Kaus BJP, Burg J-P (2012) Numerical investigation of deformation mechanics in
fold-and-thrust belts: influence of rheology of single and multiple décollements. Tectonics 31,
TC3005
Schiffman C, Bali BS, Szeliga W, Bilham R (2013) Seismic slip deficit in the Kashmir Himalaya
from GPS observations. Geophysical Research Letters 40, 5642–5645
Singh T (2008a) Hypsometric analysis of watersheds developed on actively deforming Mohand
Anticlinal ridge, NW Himalaya. Geocarto International 23, 417–427
Singh T (2008b) Tectonic implications of geomorphometric characterization of watersheds using
spatial correlation: Mohand Ridge, NW Himalaya, India. Zeitschrift fur Geomorphologie 52,
489–501
Singh T, Jain V (2009) Tectonic constraints on watershed development on frontal ridges: Mohand
Ridge, NW Himalaya, India. Geomorphology 106, 231–241
Singh T, Sharma U, Awasthi AK, Virdi NS, Kumar R (2011) Geomorphic and structural evidences
of neotectonic activity in the Sub-Himalayan belt of Nahan Salient, India. Journal of
Geological Society of India 77, 175–182
Singh T, Awasthi AK, Caputo R (2012) The Sub-Himalayan fold-thrust belt in the 1905-Kangra
earthquake zone: a critical taper model perspective for Seismic Hazard Assessment. Tectonics
31, TC6002
Singh T, Awasthi AK, Paul D, Caputo R (2014) Role of internal thrusts in NW Sub-Himalaya,
India, for SHA. Journal of Himalayan Earth Sciences 153
Stevens VL, Avouac J-P (2015) Interseismic coupling on the main Himalayan thrust. Geophysical
Reseach Letters 42, 5828–5837
Stockmal GS (1983) Modeling of large-scale accretionary wedge deformation. Journal of
Geophysical Research, 88, 8271–8287
396 T. Singh and A. K. Awasthi
1 Introduction
M. Kumar (&)
National Geotechnical Facility, Survey of India, Dehradun, India
e-mail: puniyamohit@gmail.com
R. C. Joshi
Department of Geography, Kumaun University, DSB Campus, Nainital, India
P. D. Pant
Department of Geology, Kumaun University, Nainital, India
Himalaya (see Mukherjee 2013a, 2015a; Mukherjee et al. 2012 etc.). MCT con-
stitutes a wide zone near Kund (approximately 12 km downstream from
Guptakashi), and marks the boundary between Munsiari Formation at north and
Bhatwari Formation at south (Figs. 1 and 3). Tilwara Thrust (TT) separates the
Ramgarh Group and Lesser Himalayan Metasedimentary Rocks (Fig. 1). Different
shear zones rocks viz. mylonitic augen-gneiss, calc-silicates, biotite-gneiss and
fine-grained gneiss (Fig. 2) are identified in the field during traverse mapping. The
Bhatwari Formation consists of flaggy quartzite and basic intrusion in the form of
amphibolites (Fig. 3).
3 Methodology
Detailed structural mapping has been done with the help of handheld GPS and
available data have been collected such as foliation, different joint sets, folds, faults,
shear sense indicators, slope angle and height. 1:10,000 scale map has been
Impact of Structural Damage Zones on Slope Stability: A Case … 399
Fig. 2 a mylonitic augen gneiss with symmetric clasts (Mukherjee 2017) at location-24 (30° 32′
17.05′N, 79° 04′ 49.15″E); b a contact between clac-silicate and biotite gneiss dipping towards
NNW (322°) with 48° dip amount (dash line represents the contact) at location-13 (30° 35′ 46.38′
N, 79° 01′ 20.14″E); c a xenoliths in fine grained gneiss at location-19 (30°34′41.99′N, 79° 02′
38.18″E)
prepared along the National highway-109 parallel to Mandakini river valley from
Sonparyag to Kund. Study area falls in the Survey of India toposheet numbers
53 N/3 and 53 N/4.
This study focuses on the structural damage zones and their relationship to the
different rock mass classification methods, viz., rock mass rating (RMR), kine-
matics analysis of the slopes, slope mass rating (SMR). RMR has application for
tunnels, slopes (for SMR calculation) and foundations (dam, bridges and other civil
structures; Bieniawski 1989). Field data collection such as, the orientations of
different discontinuities (joints, shear zones, faults etc.) and uniaxial compressive
strength (UCS) were calculated using Schmidt hammer method as per the
International Society for Rock Mechanics (ISRM) recommended methods (ISRM
1978, 1981, 2007; Bieniawski 1989; Brencich et al. 2013). Spacing of disconti-
nuities, conditions of discontinuities (aperture, roughness, type of filling and
weathering conditions), rock quality designation (RQD), groundwater condition,
and orientations of slopes with dip angles have been collected (Appendix-I). All
aforesaid parameters were used to calculate the RMR in this study. Kinematics
analysis has been performed for the slopes to delineate the kind of failure. Internal
400 M. Kumar et al.
friction angle has been taken as 25°–35° for the slope according to Hoek and Bray
(1981). The RMR is calculated as per Bieniawski (1989).
The slope mass rating (SMR) proposed by Romana (1985) and is obtained from
RMR by:
Impact of Structural Damage Zones on Slope Stability: A Case … 401
Here
F1 depends on parallelism between joints and slope face. Its range is from 1.00
to 0.15.
F1 ¼ ð1 SinAÞ2 ð2Þ
A is angle between the dip directions of slope face and the joint for the planar
and toppling failure. In wedge failure, A is the angle between trend and slope dip
direction.
F2 refers to joint dip angle in the planar mode of failure.
F2 ¼ tan2 B ð3Þ
B is dip angle of joint for planer failure and for wedge failure, B is plunge
amount.
F3 represents the angle between the slope and joint dip in the planar failure. In
wedge failure, F3 indicates the relationship between the slope face and plunge
amount two joints.
F4 factor represents the excavation method and the rating depends on the method
as:
and
The study area is mapped along the Mandakini River for *31 km length. The main
lithology is mylonite, chloritic-schist, calc-silicates, schistose-gneiss, mylonitic
augen gneiss, gneiss of the Munsiari Formation, and flaggy quartzite and
402 M. Kumar et al.
Fig. 4 Field photographs show: a contact between mylonite and protomylonite at location-5; b a
ductile damage zone show shearing and folding at location-10; c sulphur leaching on the surface of
sheared mylonitic gneiss at location-20; d ductile-brittle shearing show top to SW movement at
location-22
amphibolites of the Bhatwari Formation (Fig. 3). The shear zone classification is
performed in the field based on the proportion of matrix and grains. Based on the
shear zone rocks and structure present at the field, we have identified ductile-shear
zone, ductile-brittle shear zone and brittle shear zones (Ghosh 1993). Damage zones
are deformed wall rock spatially associated with faults that form during the initi-
ation and propagation of a fault, during the interaction of slip in fault linkages or
jogs, or during the flexure of beds around a fault (Jamison and Stearns 1982;
Chester and Logan 1986; Shipton and Cowie 2001, 2003; Flodin and Aydin 2004;
Kim et al. 2004b). To delineate these damage zones in the field, we have collected
several structures like rocks with mylonitic foliations (Figs. 3 and 4a) and asym-
metric folding (Figs. 4b and 5b),
S-C structures (Fig. 4b; Mukherjee 2010a, b, 2011), phyllonites and mylonites
(Figs. 4c and 5a; Bose et al. 2018), brittle-ductile S-C structures (Fig. 4d), fault
stariations (Fig. 5c; Dasgupta and Mukherjee 2017; Vanik et al. 2018) and asym-
metric boudinage (Fig. 5d; few examples in Mukherjee and Koyi 2010a, b;
Mukherjee 2013b, 2015b).
Impact of Structural Damage Zones on Slope Stability: A Case … 403
Fig. 5 Field photographs: a a contact between phyllonite and mylonitic gneiss at Main Central
Thrust near Kund; b asymmetric folding, typically intrafolial: Mukherjee et al. (2015) shows top to
NE shearing at location-34; c fault stariation show up dip movement at location-32; d a
ductile-brittle shearing show top to SW movement at location-32
L-10, L-21, L-22, L-32 and L-33 show wedge failure. Two locations L-1 and L-12
are identified as anthropogenic induced landslides (Fig. 6; Table 1). Damage zones
weaken the rocks so the slopes become more prone for failure (Fig. 7). Location-19
near the town Phata show planar failure. Here ductile shear zones are present along
the foliation with 35° dip amount and dipping towards NE (055°). Rocks tend to fail
due to these shear zones (Fig. 7; Table 1). At this location, the SMR values range
31.9–39.9 (Table 1). SMR values vary from 0 to 55.25 at L-3 and L-33, respec-
tively (Table-1). SMR values also confirm the potential damage zone for landslides.
Similarly we have investigated all the locations and find that damage zone
assessment and SMR confirms our observations.
Table 1 Data of geological investigation at different locations
Sr. Location Rock type RMR F1 F2 F3 F4 SMR Structural condition
No. (basic)
Latitude Longitude Min Max Min Max
1 1 30° 37′ 79° 00′ Mylonitic augen 54 55 0.15 1 −60 0 45 46 Due to anthropogenic
05.08″ 35.48″ gneiss activity
2 3 30° 37′ 79° 00′ Mylonitic augen 53 55 0.85 1 −60 0 2 4 Wedge Failure/
15.49″ 32.12″ gneiss ductile-brittle shear
3 4 30° 37′ 79° 00′ Mylonitic augen 54 55 0.15 1 −50 0 46.5 47.5 Wedge failure/
25.28″ 28.63″ gneiss ductile-brittle shear
4 4 (A) 30° 37′ 79° 00′ Mylonitic gneiss 54 55 0.4 1 −25 0 44 45 Wedge failure/
25.28″ 25.28″ ductile-brittle shear
5 5 30° 37′ 79° 00′ Mylonitic augen 49 53 0.85 1 −50 0 6.5 10.5 Wedge Failure/
31.49″ 19.91″ gneiss Ductile-brittle shear
6 7 30° 37′ 78° 59′ Mylonitic gneiss 53 57 1 1 −50 0 3 7 Planar failure/mica bands
33.90″ 54.69″
7 9 30° 36′ 79° 00′ Mylonitic gneiss 47 56 1 1 −25 0 22 31 Planar failure/brittle shear
55.11″ 55.52″
8 10 30° 36′ 79° 00′ Biotite gneiss 52 57 0.85 1 −60 0 1 6 Wedge failure/
43.82″ 45.01″ ductile-brittle shear
Impact of Structural Damage Zones on Slope Stability: A Case …
9 11 30° 36′ 79° 01′ Biotite gneiss 51 51 0.85 1 −60 0 0 0 Planar failure/ductile shear
45.23″ 05.63″
10 12 30° 36′ 79° 01′ Augen gneiss 45 48 0.7 1 0 0 45 48 Lithological contact/Joint
32.11″ 05.98″ control
11 13 30° 36′ 79° 01′ Biotite gneiss 54 55 0.4 1 −60 0 30 31 Planar failure/joint control
14.22″ 18.34″
12 15 30° 35′ 79° 01′ Granitic gneiss 30 35 0.15 1 0 0 30 35 Planar failure/
34.07″ 35.81″ buctile-brittle shear
(continued)
405
Table 1 (continued)
406
Fig. 7 Location-19 near Phata town show planar failure along the foliation plane (J0), along J0
ductile shear zone inducing the landslide
5 Conclusions
23 structural damage zones are identified in the study area. 6 joint sets are identified
that reduce the rock strength. Kinematics analyses confirm that 12 slopes show
planar failure and 9 are wedge failure. SMR for different modes of failure ranges 0–
55.25. This study confirms that structural damage zones are also responsible for the
slope failure and are important to consider for the road construction in the hilly area.
Structural damage zone assessment mapping also show relation with the slope
kinematics analyses and the SMR.
Acknowledgements Authors are very thankful to the Department of Geography, DSB Campus
Kumaun University Nainital to provide the necessary facility for the research. Authors are also very
thanking to the NRDMS, Department of Science and Technology to the financial assistance in the
project no. NRDMS/06/11/015 (G) Dt. 11.08.2015 “Large Scale Geological-Geomorphological
Mapping of Rudraprayag-Sonprayag Area, Mandakini valley National highway No.109. Reviewing
and editorial handling by Soumyajit Mukherjee. This work is summarized in Mukherjee (2019).
The Springer team is thanked for assistance in proofreading.
408 M. Kumar et al.
References
Ambrosi C, Crosta GB (2011) Valley slope influence on deformation mechanisms of rock slopes.
In: Jaboyedoff M (ed) Slope tectonics, vol 351. Geological Society, London, Special
Publications, pp 215–233
Badger TC (2002) Fracturing within anticlines and its kinematic control on slope stability.
Environmental & Engineering Geoscience 8, 19–33
Bieniawski ZT (1989) Engineering rock mass classification. Wiley, New York
Bose N, Dutta D, Mukherjee S (2018) Role of grain-size in phyllonitisation: Insights from
mineralogy, microstructures, strain analyses and numerical modeling. Journal of Structural
Geology 112, 39–52
Brencich A, Cassini G, Pera D, Riotto G (2013) Calibration and reliability of the rebound
(Schmidt) hammer test. Civil Engineering and Architecture 1, 66–78
Brideau M-A, Stead D, Couture R (2006) Structural and engineering geology of the east gate
landslide, Purcell mountains, British Columbia, Canada. Engineering Geology 84, 183–206
Caine JS, Evans JP, Forster CB (1996) Fault zone architecture and permeability structure. Geology
24, 1025–1028
Chester FM, Logan JM (1986) Implications for mechanical properties of brittle faults from
observations of the Punchbowl fault zone, California. Pure and Applied Geophysics 124,
79–106
Childs C, Manzocchi T, Walsh JJ, Bonson CG, Nicol A, Schöpfer MPJ (2009) A geometric model
of fault zone and fault rock thickness variations. Journal of Structural Geology 31, 117–127
Choi JH, Edwards P, Ko K, Kim YS (2016) Definition and classification of fault damage zones: a
review and a new methodological approach. Earth-Science Reviews 152, 70–87
Cowie PA, Scholz CH (1992) Growth of faults by accumulation of seismic slip. Journal of
Geophysical Research 97, 11085–11095
Dasgupta S, Mukherjee S (2017) Brittle shear tectonics in a narrow continental rift: asymmetric
non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology 125, 561–591
Flodin EA, Aydin A (2004) Evolution of a strike-slip fault network, Valley of Fire State Park,
southern Nevada. Geological Society of America Bulletin 116, 42–59
Ghosh SK (1993) Structural geology, fundamentals and modern developments, vol 1. Pergamon
Press, p 598
Gupta V, Bist KS (2004) The 23 September 2003 Varunavat Parvat landslide in Uttarkashi city,
Uttaranchal. Current Science 87, 1600–1605
Hoek E, Bray J (1981) Rock slope engineering. The Institution of Mining and Metallurgy, London
International Society for Rock Mechanics (ISRM) (1978) Suggested methods for the quantitative
description of discontinuities in a rock mass. International Journal of Rock Mechanics and
Mining Sciences and Geomechanics Abstracts 6, 319–368
ISRM (1981) Suggested methods for determining hardness and abrasiveness of rocks. Rock
characterization, testing and monitoring: ISRM suggested methods. Pergamon, Oxford,
pp 95–96
ISRM (2007) The complete ISRM suggested methods for rock characterization, testing and
monitoring: 1974–2006. Springer, Berlin
Jamison WR, Stearns DW (1982) Tectonic deformation of Wingate Sandstone, Colorado National
Monument. American Association of Petroleum Geologists Bulletin 66, 2584–2608
Kim YS, Peacock DCP, Sanderson DJ (2004a) Fault damage zones. Structural Geology 26, 503–
517
Kim YS, Peacock DCP, Sanderson DJ (2004b) Fault damage zones. Structural Geology 26, 503–
517
Kumar M, Rana S, Pant PD, Patel RC (2017) Slope stability analysis of Balia Nala landslide,
Kumaun Lesser Himalaya, Nainital, Uttarakhand, India. Journal of Rock Mechanics and
Geotechnical Engineering 9, 150–158
Impact of Structural Damage Zones on Slope Stability: A Case … 409
McGrath AG, Davison I (1995) Damage zone geometry around fault tips. Structural Geology 17,
1011–1024
Mukherjee S (2010a) Structures in Meso- and Micro-scales in the Sutlej section of the Higher
Himalayan Shear Zone, Indian Himalaya. e-Terra 7, 1–27
Mukherjee S (2010b) Microstructures of the Zanskar Shear Zone. E-journal: Earth Science India 3,
9–27
Mukherjee S (2011) Mineral fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth Sciences 100, 1303–1314
Mukherjee S (2013a) Channel flow extrusion model to constrain dynamic viscosity and Prandtl
number of the Higher Himalayan Shear Zone. International Journal of Earth Sciences 102,
1811–1835
Mukherjee S (2013) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy,
Berlin, pp 1–111. ISBN 978-3-642-25608-0
Mukherjee S (2014) Atlas of shear zone structures in Meso-scale. Springer, Berlin, pp 1–128.
https://doi.org/10.1007/978-3-319-00089-3
Mukherjee S (2015a) A review on out-of-sequence deformation in the Himalaya. In: Mukherjee S,
Carosi R, van der Beek P, Mukherjee BK, Robinson D (eds) Tectonics of the Himalaya, vol
412. Geological Society, London, Special Publications, pp 67–109
Mukherjee S (2015) Atlas of structural geology. Elsevier. Amsterdam. ISBN: 978-0-12-420152-1
Mukherjee S (2017) Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences 106, 1453–1468
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Zanskar Section- microstructural
studies & extrusion mechanism by a combination of simple shear & channel flow. International
Journal of Earth Sciences 99, 1083–1110
Mukherjee S, Koyi HA (2010b) Higher Himalayan Shear Zone, Sutlej section- structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Mukherjee S, Koyi HA, Talbot CJ (2012) Implications of channel flow analogue models in
extrusion of the Higher Himalayan Shear Zone with special reference to the out-of-sequence
thrusting. International Journal of Earth Sciences 101, 253–272
Mukherjee S, Punekar J, Mahadani T, Mukherjee R (2015) A review on intrafolial folds and their
morphologies from the detachments of the western Indian Higher Himalaya. In: Mukherjee S,
Mulchrone KF (eds) Ductile Shear Zones: From Micro- to Macro-scales. Wiley Blackwell,
pp 182–205
Panikkar SV, Subramanyan V (1997) Landslide hazard analysis of the area around DehraDun and
Mussoorie, Uttar Pradesh. Current Science 73, 1117–1123
Pant PD, Luirei K (1999) Malpa rockfalls of 18 August 1998 in the Northeastern Kumaun
Himalaya. Journal of the Geological Society of India 54, 415–420
Paul SK, Mahajan AK (1999) Malpa rockfall disaster, Kali valley, Kumaun Himalaya. Current
Science 76, 485–487
Peacock DCP, Nixon CW, Rotevatn A, Sanderson DJ, Zuluaga LF (2017) Interacting faults.
Journal of Structural Geology 97, 1–22
Puniya MK, Joshi P, Pant PD (2013) Geological investigation of Nainital Bypass: a special
emphasis on slope stability analysis, Kumaun Lesser Himalaya. In: Himalayan vulnerability,
Uttarakhand, 2013: learning for planning and action. Xpressions Print & Graphics Pvt. Ltd.
pp 65–72
Romana M (1985) New adjustment ratings for application of Bieniawski classification to slopes.
In: Proceedings of the international symposium on the role of rock mechanics in excavations
for mining and civil works. International Society of Rock Mechanics, Zacatecas, pp 49–53
Sah N, Kumar M, Upadhyay R, Dutt S (2018) Hill slope instability of Nainital City. Kumaun
Lesser Himalaya, Uttarakhand, India. https://doi.org/10.1016/j.jrmge.2017.09.011
410 M. Kumar et al.
Shipton ZK, Cowie PA (2001) Damage zone and slip-surface evolution over mm to km scales in
high-porosity Navajo sandstone, Utah. Journal of Structural Geology 23, 1823–1844
Shipton ZK, Cowie PA (2003) A conceptual model for the origin of fault damage zone structures
in high-porosity sandstone. Journal of Structural Geology 25, 333–344
Stead D, Eberhardt E (2013a) Understanding the mechanics of large landslides. Italian Journal of
Engineering Geology and Environment Book Series (6), 85–108
Stead D, Eberhardt E (2013b) Understanding the mechanics of large landslides. Italian Journal of
Engineering Geology and Environment, 85–109
Stead D, Wolter A (2015) A critical review of rock slope failure mechanism: the importance of
structural geology. Journal of structural Geology 74, 1–23
Valdiya KS (1980) Geology of Kumaun Lesser Himalaya. Wadia Institute of Himalayan Geology,
Dehradun
Vanik N, Shaikh H, Mukherjee S, Maurya DM, Chamyal LS (2018) Post-Deccan trap stress
reorientation under transpression: evidence from fault slip analyses from SW Saurashtra,
western India. Journal of Geodynamics, 121, 9–19
Documentation of Brittle Structures
(Back Shear and Arc-Parallel Shear)
from Sategal and Dhanaulti Regions
of the Garhwal Lesser Himalaya
(Uttarakhand, India)
1 Introduction
We study the brittle shear tectonics from the Lesser Himalayan terrain in the
Mussoorie, Dhanaulti and Satengal regions, Uttarakhand, India. The Lesser
Himalaya is bound by the Main Boundary Thrust (MBT) in the south and the Main
Central Thrust Zone (MCTZ) in the north. The Lesser Himalayan sequence is
divided in this region as the Inner Lesser Himalaya and the Outer Lesser Himalaya
(Célérier et al. 2009a, b). The Mussoorie- and the Garhwal synclines are the two
major structures of the Outer Lesser Himalaya. The area around Dhanaulti (Fig. 1)
constitutes a part of the northern limb of the Mussoorie syncline. The Satengal
region (Fig. 1) lies at the core of the Mussoorie syncline. Several workers (e.g.,
Srivastava et al. 2011; Bose et al. submitted-1, 2; Bose et al. 2018) discuss major
structural features and the deformation phases from the Garhwal Himalaya.
However, a detail study remained due on its brittle deformations. This work
describes brittle shear sense indicators from the terrain.
2 Previous Works
The first-hand rock description and regional mapping in this area have been done by
Middlemiss (1887), Auden (1937), Jain (1972), Shankar and Ganesan (1973),
Valdiya (1975), Saklani (1978). The Mussoorie Syncline mainly consists of sedi-
mentary rocks and low-grade metamorphic rocks of Tal-, Krol-, Blaini-, Nagthat-,
Chandpur-, and Mandhali Formation. Out of all these, the Mandhani, the Chandpur
Fig. 1 Structural map of the central portion of Mussoorie syncline is showing different structural
data. Structural data in red colour are taken from previous works, and those in black colour
indicates those collected in the present study. The map is reproduced from Dubey and
Jayangondaperumal (2005)
and the Nagthat Formation belong to the Jaunsar Group and the remainder to the
Mussoorie Group (Fig. 1).
Raman Spectroscopic study of carbonaceous material of samples collected
across the axis of Mussoorie Syncline yields a deformation temperature <300 °C
(Célérier et al. 2009a, b). This is overlain by less metamorphosed rocks of the
Simur Group and constitutes partly the Garhwal nappe. The southern limb of the
Mussoorie Syncline overrides the Siwalik zone along the MBT (Fuchs and Sinha
1978). Intense-sheared black Proterozoic Chandpur phyllites along the MBT near
the Sahenshahi temple has been studied in detail by Bose and Mukherjee
(submitted-a). The authors use Raman Spectroscopy on carbonaceous materials and
obtain a deformation temperature <550 °C (Bose and Mukherjee, submitted-a).
Further, *6–49 MPa of flow stress and *10−15–10−16 s−1 of strain rate has been
obtained from the MBT zone rocks from this region that reveals a top-to-S/SW
fore-shear in field and under an optical microscope (Bose et al. submitted-a).
According to Srivastava et al. (2011), the Sategal area deformed four times. The
first three deformation phases are attributed to ductile deformations and the last one
a brittle deformation. The Mussoorie Syncline, a structural part of the Krol Nappe,
consists of Tal-, Krol-, Blaini-, Nagthat-, Chandpur-, and Subathu Formation. In
Dhanaulti and nearby areas, mainly Blaini-, Tal-, and Krol Formation crop out. The
main meso-scale structures in this area are primary bedding planes, slaty cleavages,
Documentation of Brittle Structures … 413
axial planes of different generations of folds (Srivastava et al. 2011), and thrusts/
faults (Dubey 2005). The pop-up klippen structure is indicated by the rocks of
Chandpur, Mandhali and Subathu Formation at the Satengal area (Dubey and
Jayangondaperumal 2005).
The Baliana Formation (/Chandpur Formation) and the Mandhani Formation (/
Saklana Formation) are separated by the Ringalgarh Thrust. The pop-up klippen
developed since the MBT bifurcates at south into an oblique ramp and into a blind
thrust that generated two low-angle backthrusts (Dubey and Jayangondaperumal
2005).
At Satengal and in the adjoining areas, folding happened thrice. The main
schistosity trends NW and NE. The NW trending earlier tight to isoclinal F1 folds
are characterized by deformed quartz veins and quartzo-feldspathic layers. The
usually NE trending F2 folds developed on the limbs of the F1 folds with open to
isoclinal geometries. F3 chevron folds occur in phyllites of Chandpur Formation
near Satengal and trends E-W (Srivastava et al. 2011).
3 Present Work
This section addresses details of S0 and S1 planes. Brittle shear detail is also presented
from the following three traverses: (i) Mussoorie up to Kathu Khal (Surkunda
Devi temple) along the Mussourie-Chamba road; (ii) Kathu-Khal to Satengal; and
(iii) *NE-SW traverse was taken from Raipur to Silla.
Primary sedimentary bedding planes (S0) are commonly observed in the study area.
These S0 planes can be identified based on lithology change and color variation.
From L-1 to L-6, dip of S0 ranges 40°–42° with *SW dip direction. The dip
direction changes from SW to NW in and around L-7. Dip ranges 25°–65° with
a *SW dip direction between L-8 and L-11. S0 dips oppositely towards NE and N
with 30° dip at L-12 and L-13, south of Dhanaulti.
The S0-planes near the Satengal region between L-15 to 18 show 40°–45° dip,
and dip direction towards SW. At L-24 grey coloured quartzites dip N, from L-28 to
31 the dip of the S0 varies 35°–60° with dip direction NW to E, in between L-32
and 33 dip direction varies from SE to S. S0 dips toward SW direction in locations
33–37.
Slaty Cleavage (S1) (Fig. 2):
The exposed rocks in and around Dhanaulti are slaty. Phyllites are observed in the
Satengal region. A pervasive, parallel foliation (S1) of fine-grained platy minerals
occurs in these rocks, which generally develop perpendicular to the direction of
414 S. Mahato et al.
Fig. 2 Stereo-plots of planer features S0 and S1 on equal area net. L-21 to L-24 shows S1-plots,
and the rest represents S0-plots
maximum stress. Usually this S1 plane parallels the S0 planes if not folded or
sheared. In Dhanaulti region, the S1 planes dip 30°–40° towards SW. In Satengal
region, S1 dips 25°–50° and the dip direction varies widely from SW to S to SE to
N.
Shear fractures (Fig. 3) are very common structural features observed in the study
area. The P-planes at L-28 and 36 generally dip in opposite direction and with the
M/Y shear fracture this makes a low-angle. Kinematically, such shear planes can
usually be correlated with low-angle thrusts. Within the same setting, the Riedel
fractures can be represented as low-angle normal faults. The R/-planes/antithetic
faults make a high-angle with the Y plane. The angles between the P, R and R/with
the Y-plane can depend on the rock’s physical properties and the stress regime.
Documentation of Brittle Structures … 415
Fig. 3 Nomenclatures for shear fractures developed in a top-to-right shear zone. M/Y, Main shear
plane; P: shear fractures; R: Riedel shears; R/: fracture. From Petit (1987)
The first *E-W traverse was taken along the Mussoorie-Chamba road from
Mussoorie up to Kathu Khal (Surkunda Devi temple). A vertical shear zone (Fig. 4)
exists at L-2 on a N140°S striking road cut section. Around 13 km east of L-3,
top-to-SE brittle shear (Fig. 5a) was observed at L-7. Top-to-NW down shear
(Fig. 6a) was observed *650 m SE from L-6. At L-7, top-to-NE up brittle shear
(Fig. 6c), i.e., back deformation, occurs in slates. About 600 m east from L-7, the
same top-to-NE up brittle shear (Fig. 6d) occurs in slates at L-8. A top-to-NE
sub-horizontal brittle shear (Fig. 6b) exists in slate close to L-9. Around 300 m south
of L-9, a top-to-SW up shear, also can be called a fore shear that is common in
Himalayan orogen in its various units (e.g., Mukherjee 2010a, b, 2012, 2013, 2014;
Mukherjee and Koyi 2010a, b) exists in slate. Top-to-NW shear (Fig. 5b) also exists.
Fig. 5 a Field photograph of vertical exposure of quartzite showing top-to-SE shear. Y- and
P-planes are stereo-plotted in inset. b Top-to-SW up brittle shear. Pole of P-plane (25° dip towards
10°) is plotted in inset
Fig. 6 Brittle shear Y- and P-planes observed in a vertical rock section. Poles of brittle planes are
plotted inside insets. a Note a sharp contact between slate at top and quartzite at bottom.
Top-to-NW brittle shear localized in slate due to high accumulation of shear strain. P-plane
attitude: 10° dip towards 140°. b Shear fractures developed on slate. Top-to-NE brittle shear sense.
The P-plane dips 14° and towards 225°. c Y- (dip 34° towards 200°) and P-planes (dip 65° towards
240°) developed on slate. Top-to-NE up shear. d Top-to-NE up shear. Y-plane: dip 35° towards
190°, P-plane: dip 70° towards 230°
Documentation of Brittle Structures … 417
Fig. 7 a, b Sub-vertical sections at L-12 showing S- and C-fabrics developed in slates. top-to-NE
up ductile shear. Stereo-plot of S planes in inset. Fractures developed along the S-planes in slates.
In particular, in c the C-planes are occupied by secondary quartz mineralization. Here the pen
length is 13 cm
Fig. 8 Brittle shear planes in vertical sections close to L-12. a P-plane dips 30° towards 248°,
top-to-NNE shear. b P-plane dips 40° towards 238°, top-to-NE shear. c P-plane dips 70° towards
190°, top-to-NNE shear. d P plane dips 65° towards 255°, at L-13. Top-to-NNW horizontal shear
Fig. 9 a Clear-cut top-to-NE brittle shear observed in vertical section of slates. P-plane dips 46°
towards 230°. b Top-to-NEE shear developed in white Ringalgarh quartzite. P-plane dips 58°
towards 252°. c Top-to-NE up shear prominent in white quartzite. P-plane dips 65° towards 232°.
d Sharp contact between quartzite and phyllite noted at L-19
Documentation of Brittle Structures … 419
Fig. 10 a Top-to-NE down shear in quartzite. Y-plane dips 70° towards 78°, P-planes dips 15°
towards 80°. b Top-to-SE down shear in shale. Y-plane dips 22° towards 38°, P-plane dips 28°
towards 320°. c Field photograph showing vertical sensed of shear Y-, P-, R-, and R/-planes are
marked. d Top-to-SE down shear in bedded quartzite. Y-plane dips 22° towards 180°; P-plane dips
36° towards 286°. a, b: sub-vertical sections; c, d: inclined sections
The second traverse was taken from Kathu-Khal to Satengal. At L-14, top-to-NE
horizontal shear (Fig. 9a) was observed in slates. Near Satengal, at L-17 top-to-NE
shear exists in quartzites. Top-to-NE up shear (Fig. 9b, c) occur at L-17 and 18 in
white quartzites.
The third *NE-SW traverse was taken from Raipur to Silla. At L-28 shear frac-
tures are quite prominent in the quartzites showing top-to-NE down slip (Fig. 10a).
About 1.5 km NE of L-28, top-to-SE down shear (Fig. 10b) occurs in black shale.
Vertical slip (Fig. 10c) exists at L-31. Top-to-SE down shear sense (Fig. 10d) was
observed at L-32. At L-33, top-to-SW down shear sense (Fig. 11a) was observed on
quartzite. At L-34, top-to-NW up shear (Fig. 11b) was observed in quartzites, Near
L-34, a top-to-NE up shear (Fig. 11c) was observed in white quartzite. Top-to-SW
down shear (Fig. 11d) was observed at L-36 in black shale. All the observed shear
senses are shown along the field traverses presented in Fig. 12.
Cross-sections (Figs. 13 and 14) were prepared along lines a–b close to
Dhanaulti and along the line X–Y. The lines a–b and X–Y can be found in Fig. 12.
The conventional method is used to draw the cross section. As we do not know the
actual curvature at the limbs therefore straight lines are drawn to construct the fold.
As sharp hinge bearing chevron folds were noted in the smaller scale (Figure with
authors), a large fold of such geometry looks plausible (Fig. 14).
420 S. Mahato et al.
Fig. 11 a Shear fractures developed on quartzite showing top-to-SW down brittle shear sense.
Y plane attitude dips 22° towards 223°; b top-to-NW up shear. P-plane dips 23° towards 218°. The
rock type is quartzite. c Top-to-NE up shear. Y-plane dips 33° towards 226°, P-plane dips 62°
towards 226°. Shear fractures developed quartzite. d Shear fractures developed on shale showing
top- to-SW down shear sense. Y-plane dips 31° towards 222°, P-plane dips 15° towards 18°
Fig. 12 Distributions of different brittle shear senses are shown in the map. The map is
reproduced from Dubey and Jayangondaperumal (2005)
Documentation of Brittle Structures … 421
Fig. 13 Cross-section of the Mussoorie syncline is drawn along the X–Y line of Fig. 12. The X–
Y line trends N36°E. B: Chandpur Fm, C: Nagthat Fm, D: Blaini Fm, E: Krol Fm, F: Tal Fm
Acknowledgements This work was a part of the M.Sc. thesis of S. Mahato. S. Mukherjee thanks
IIT Bombay for providing a sabbatical in 2017. N. Bose acknowledges Junior Research
Fellowship of UGC, grant number: F.2-2/98(SA-1). The Springer proofreading team is thanked for
assistance. Consult Banerjee et al. (2019) for tectonic updates on arc-parallel deformation from
Sikkim Lesser Himalaya, and Dutta et al. (submitted) from that in Mohand region, Siwalik. This
will establish link between chapters in the book. Mukherjee (2019) summarizes this work.
References
Auden JB (1937) The structure of the Himalaya in Garhwal. Records of the Geological Survey of
India 71, 407–433
Banerjee S, Bose N, Mukherjee S (2019) Field structural geological studies around Kurseong,
Darjeeling-Sikkim Himalaya, India. In: Mukherjee S (ed) Tectonics and structural geology:
Indian context. Springer International Publishing AG, Cham, pp 425–440. ISBN
978-3-319-99340-9
422 S. Mahato et al.
Bell TH, Rubenach MJ, Fleming PD (1986) Porphyroblast nucleation, growth and dissolution in
regional metamorphic rocks as a function of deformation partitioning during foliation
development. Journal of Metamorphic Geology 4, 37–67
Bose N, Dutta D, Mukherjee S (2018) Role of grain-size in phyllonitisation: insights from
mineralogy, microstructures, strain analyses and numerical modeling. Journal of Structural
Geology 112, 39–52
Bose N, Mukherjee S (Submitted-1) Estimation of deformation temperature, flow stress and strain
rate from the Main Boundary Thrust rocks, Garhwal Lesser Himalaya, India. Journal of Earth
System Science
Bose N, Mukherjee S (Submitted-2) Field documentation and genesis of the back-structures from
a part of the Garhwal Lesser Himalaya. Uttarakhand, India. Tectonic implications. In: Sharma,
Villa IM, Kumar S (eds) Crustal Architecture and Evolution of the Himalaya-Karakoram-Tibet
Orogen. Geological Society of London Special Publications
Célérier J, Harrison TM, Beyssac O, Herman F, Dunlap WJ, Webb AAG (2009a) The Kumaun
and Garwhal Lesser Himalaya, India: Part 2. Thermal and deformation histories. Geological
Society of America Bulletin 121, 1281–1297
Célérier J, Harrison TM, Webb AAG, Yin A (2009b) The Kumaun and Garhwal Lesser Himalaya,
India: Part 1. Structure and stratigraphy. Geological Society of America Bulletin 121, 1262–
1280
Dubey AK (2014) Understanding an orogenic belt. Springer, Cham. ISBN 978-3-319-05587-9
Dubey AK, Jayangondaperumal R (2005) Pop-up klippen in the Mussoorie Syncline Lesser
Himalaya: evidence from field and model deformation studies. In: Saklani RS (ed) Himalaya
(Geological Aspects), Satish Serial Publishing House, Delhi 3, pp 203–222
Dutta D, Biswas T, Mukherjee S (Submitted) Orogen-parallel compression in NW Himalaya:
Evidence from structural and paleostress studies of brittle deformation from the pebbles of
Upper Siwalik conglomerates, Uttarakhand, India. Journal of Earth System Science
Fuchs G, Sinha AK (1978) The tectonics of the Garhwal-Kumaun Lesser Himalaya. Jahrbuch der
Geologischen Bundesanstalt 121, 219–241
Jain AK (1972) Overthrusting and emplacement of basic rocks in Lesser Himalaya, Garhwal,
UP. Journal of Geological Society of India 13, 226–237
Jayangondaperumal R, Dubey AK, Sangode SJ, Sathyanarayana KV (2001) Superimposed
folding, finite strain and magnetic lineation in the Mussoorie Syncline, Lesser Himalaya:
implications for regional thrusting and the Indian Plate motion. Himalayan Geology 22, 207–
216
Long S, McQuarrie N (2010) Placing limits on channel flow: insights from the Bhutan Himalaya.
Earth and Planetary Science Letters 290, 375–390
Mandl G (1999) Faulting in brittle rocks: an introduction to the mechanics of tectonic faults.
Springer Science and Business Media
Marshak S (1988) Kinematics of orocline and arc formation in thin-skinned orogens. Tectonics 7,
73–86
Marshak S (2009) Essentials of geology, 4th edn. WW Norton and Company, pp 18. ISBN:
9780471405214
Middlemiss CS (1887) Crystalline and metamorphic rocks of the Lower Himalaya, Garhwal and
Kumaun. Records of the Geological Survey of India 20, 134–143
Mukherjee S (2010a) Structures in meso- and micro-scales in the Sutlej section of the Higher
Himalayan Shear Zone, Indian Himalaya. e-Terra 7, 1–27
Mukherjee S (2010b) Microstructures of the Zanskar Shear Zone. e-Journal: Earth Science India 3,
9–27
Mukherjee S (2012) Tectonic implications and morphology of trapezoidal mica grains from the
Sutlej section of the Higher Himalayan Shear Zone, Indian Himalaya. The Journal of Geology
120, 575–590
Mukherjee S (2013) Higher Himalaya in the Bhagirathi section (NW Himalaya, India): its
structures, backthrusts and extrusion mechanism by both channel flow and critical taper
mechanisms. International Journal of Earth Sciences 102, 1851–1870
Documentation of Brittle Structures … 423
Mukherjee S (2014) Mica inclusions inside Host Mica Grains from the Sutlej section of the Higher
Himalayan Crystallines, India-Morphology and Constrains in Genesis. Acta Geologica Sinica
88, 1729–1741
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Zanskar section-microstructural
studies and extrusion mechanism by a combination of simple shear and channel flow.
International Journal of Earth Sciences 99, 1083–1110
Mukherjee S, Koyi HA (2010b) Higher Himalayan Shear Zone, Sutlej section-structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Passchier CW, Trouw RAJ (2005) Microtectonics. Springer, Berlin, pp 1–366. ISBN10
3-540-64003-7
Petit JP (1987) Criteria for the sense of movement on fault surfaces in brittle rocks. Journal of
Structural Geology 9, 597–608
Powell CM, Conaghan PJ (1973) Plate tectonics and the Himalayas. Earth and Planetary Science
Letters 20, 1–12
Saklani PS (1978) Deformation and tectonism of Mukhem area, Lesser Himalaya. Tectonic
Geology of the Himalaya 15, 42
Shanker R, Ganesan TM (1973) A note on the Garhwal Nappe. Himalayan Geology 3, 72–82
Srivastava HB, Sinha LK, Katiyar V (2011) Mesoscopic structures from the area around Satengal,
Lesser Garhwal Himalaya. Journal of Scientific Research 55, 25–34
Valdiya KS (1975) Lithology and age of the Tal Formation in Garhwal, and implication on
stratigraphic scheme of Krol Belt in Kumaun Himalaya. Geological Society of India 16, 119–
134
Valdiya KS (1980) Geology of Kumaun Lesser Himalaya. Wadia Institute of Himalayan Geology
Field Structural Geological Studies
Around Kurseong, Darjeeling-Sikkim
Himalaya, India
1 Introduction
Fig. 1 Lithological map reproduced from Lahiri and Gangopadhway (1974). The garnet isograd
is shown
is feldspar, biotite and occasional muscovite (Sen 1971). The Lesser Himalayan
phyllites connote the sediments of the northern boundary of the Indian plate
(Acharyya et al. 2017). However, different authors provided different names for the
similar litho-/structural-units. For example, Bhattacharyya and Mitra (2009), and
their subsequent works report that the following lithounits are present in the current
study area (from S to N): Siwalik Formation, Gondwana/Buxa Formation, Daling
Group with Lingtse Gneiss intrusions, Paro Gneiss, Darjeeling/Kanchenjunga
Gneiss. This has been followed in Figs. 2, 3 and 4.
2 Structures
3 Metamorphism
analyses that both the Daling and Darjeeling rocks are of same origin but due to
heterogeneity in bulk composition the metamorphic products are different.
4 Lithologies
In the field area (Fig. 2), the lithology varies along the N-S transect. A change in
metamorphic grade was also observed. Starting with phyllites near Tindharia in the
southern part of the field area, up to gneissic grade of rock was encountered near
Kurseong (Fig. 3). According to the classification of Mallet (1874), phyllites belong
to the Daling Group of the (Inner-) Lesser Himalaya. The other rock-types,
quartz-mica-schist, garnet bearing mica-schist, mica-gneiss and quartzite, belong to
the Darjeeling Formation.
Field Structural Geological Studies Around Kurseong … 431
5 Structures
Foliation planes in the Tindharia–Kurseong section dip towards N, NE, and SW. In
Kurseong–Rohini section, they dip towards E and NW, and in the Kurseong–
Pankhabari section towards NW, N and NNE (Figs. 4 and 5).
432 S. Banerjee et al.
Fig. 5 Location-wise stereo-plots of planar features (joints, schistosity and other foliation planes).
Poles of planar data shown
Field Structural Geological Studies Around Kurseong … 433
5.1 Folds
Tight, isoclinal and open folds are noted in the field area. Tight isoclinal asym-
metric folds are found in light-coloured quartz rich gneissic bands in L-11 to -14 in
Darjeeling Formation (Figs. 6 and 7). Open folds are observed in quartz mica schist
(L-9) before reaching Mahanadi. Fold axes plunge in variable amount: 8°–38°
broadly in three directions: NE, SW and SE.
Three shear senses are noted in mica schist and garnet bearing quartz mica schist of
Darjeeling Formation: top-to-S, top-to-E and top-to-W. The inclination of the
P-planes with respect to the Y-planes are used to deduce the slip sense conclusively,
even in absence of any slipped marker layer (e.g., Passchier and Trouw 2005;
Mukherjee 2014a, b, 2015). Out of these, only the former shear sense has been
reported by the previous authors from this and other Himalayan segments, which is
rather common in the collisional mountains as the “fore-shear” (Mukherjee 2007,
Fig. 6 Folds in gneiss in sub-vertical sections. a Folded light quartz rich garnetiferous mica-
gneissic and dark micaceous layer in L-13. b Tight fold in lighter coloured gneissic layer.
A rootless fold is noted at L-15
434 S. Banerjee et al.
Fig. 7 Tight folds in sub vertical sections of light colouredgarnetiferous quartz mica gneiss.
a Isoclinal fold in L-10, b secondary Z-fold in L-10
Fig. 8 Top-to-west ductile sheared quartz body within non-foliated micaceous rock in
sub-vertical sections. a Two shear senses observed in quartz vein within the non-foliated rock
at L-10. b Sheared quartz vein within the gneissic rock at L-18. Intrafolial folding in garnet-rich
light colored band within a non-foliated micaceous rock
Field Structural Geological Studies Around Kurseong … 435
Fig. 9 Top-to-E sheared quartz vein in garnet-bearing mica schist in a sub-vertical section.
a Lens-shaped quartz vein (L-14). b Rootless fold (L-14). c Quartz lens (L-10)
Fig. 10 Top-to-S shear sense observed in sub-vertical sections. a Quartz fish at L-6. b Brittle Y-
and P-planes in low grade schist in L-6
436 S. Banerjee et al.
Fig. 11 a Small scale normal fault observed in L-6 marked by displacement of quartz vein seen in
a sub vertical section, b fault plane containing slicken slide within it, c sketch of the slicken slide
lineation on the fault plane. plunge 34° towards 62°E, d stereo-plot of the fault plane and the
lineation on it
Field Structural Geological Studies Around Kurseong … 437
2012, 2013a, b, c; Mukherjee and Koyi 2010a, b). The rest of the shear senses are the
new findings through this work. Intrafolially folded quartz veins show a top-to-W
shear around L-10 (Mukherjee et al. 2015; Fig. 8). Top-to-E shear in L-14 is marked
by rootless intrafolial folds, lens-shape mineral fish of quartz (e.g., Mukherjee 2011),
fracture planes in quartz vein (L-19), around Rohiniroad (Fig. 9). Asymmetric eye/
augen-shaped clasts (Fig. 3 and 11a) and brittle Y- and P-planes in L-6 (Fig. 10)
characterize the top-to-S shear. Orogen-parallel shear (top-to-E and -W) are also new
findings in this work. Mahato et al. (2019) also report the same from the Lesser
Himalaya from Garhwal region, Uttarakhand. Orogen parallel deformation is also
recently further reported from Mohand area, Siwalik (Dutta et al. submitted).
In L-6 (*950 m before the Gayabari railway station), a single small-scale NNE
dipping normal fault (Fig. 11) was observed within the high-grade phyllitic rock
unit marked by the slip of a narrow quartz vein. The vein also shows a small-scale
pinch-and-swell geometry.
Fig. 12 Joint planes in phyllitic rock cross-cutting quartz vein in L-5, observed in a sub-vertical
section
438 S. Banerjee et al.
Acknowledgements This work was a part of SB’s M.Sc. thesis. SM acknowledges the research
sabbatical for the year 2017 and the CPDA support from IIT Bombay. Annett Buettener, Helen
Ranchner and the Springer proof-reading team are thanked for assistance. Mukherjee (2019)
summarizes this work.
References
Acharyya SK (1989) The Daling Group, its nomenclature, tectonostratigraphy and structural grain:
with notes on their possible equivalents. Geological Survey of India 5–13 (Special Publication
22)
Acharyya SK, Ghosh S, Mandal N, Bose S, Pande K (2017) Pre-Himalayan tectono-magmatic
imprints in the Darjeeling-Sikkim Himalaya (DSH) constrained by 40Ar/39Ar dating of
muscovite. Journal of Asian Earth Sciences 146, 211–220
Acharyya SK, Ray KK (1977) Geology of the Darjeeling–Sikkim Himalaya. Guide to Excursion
No. 3. In: Fourth international Gondwana symposium (Calcutta). Geological Survey of India,
pp 1–25
Auden JB (1935) Traverses in the Himalaya. Records of Geological survey of India 69, 123–167
Bhattacharyya K, Mitra G (2009) A new kinematic evolutionary model for the growth of a duplex
—an example from the Rangit duplex, Sikkim Himalaya, India. Gondwana Research 16,
697–715
Bose S, Mandal N, Acharyya SK, Ghosh S, Saha P (2014) Orogen-transverse tectonic window in
the Eastern Himalayan fold belt: a superposed buckling model. Journal of Structural Geology
66, 24–41
Dutta D, Biswas T, Mukherjee S (Submitted) Orogen-parallel compression in NW Himalaya:
evidence from structural and paleostress studies of brittle deformation from the pebbles of
Upper Siwalik conglomerates, Uttarakhand, India. Journal of Earth System Science
Jangpangi BS (1972) Some observations on the stratigraphy and reverse metamorphism in
Darjeeling Hills. Himalayan Geology 2, 356–370
Lahiri S (1973) Some observations on structure and metamorphism of the rocks of Kurseong–
Tindharia region, Darjeeling district, West Bengal. Himalayan Geology 3, 365–371
Lahiri S, Gangopadhway PK (1974) Structure pattern in rocks in Pankhabari–Tindharia region
Darjeeling district WB with special reference its bearing on stratigraphy. Himalayan Geology
4, 151–170
Mahato S, Mukherjee S, Bose N (2019) Documentation of brittle structures (back shear and
arc-parallel shear) from Sategal and Dhanaulti regions of the Garhwal Lesser Himalaya
(Uttarakhand, India). In: Mukherjee S (ed) Tectonics and structural geology: Indian context.
Springer International Publishing AG, Cham, pp 411–423. ISBN 978-3-319-99340-9
Mallet FR (1874) On the geology and mineral resources of the Darjiling District and the western
Duars. Memoirs of the Geological Survey of India 11, 1–50
Mukherjee S (2007) Geodynamics, deformation and mathematical analysis of metamorphic belts,
NW Himalaya. Ph.D. thesis, Indian Institute of Technology Roorkee, pp 1–267
Field Structural Geological Studies Around Kurseong … 439
Mukherjee S (2011) Mineral fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth Sciences 100, 1303–1314
Mukherjee S (2012) Tectonic implications and morphology of trapezoidal mica grains from the
Sutlej section of the Higher Himalayan Shear Zone, Indian Himalaya. The Journal of Geology
120, 575–590
Mukherjee S (2013a) Higher Himalaya in the Bhagirathi section (NW Himalaya, India): its
structures, backthrusts and extrusion mechanism by both channel flow and critical taper
mechanisms. International Journal of Earth Sciences 102, 1851–1870
Mukherjee S (2013b) Deformation microstructures in rocks. Springer Geochemistry/Mineralogy,
Berlin. pp 1–111. ISBN 978-3-642-25608-0
Mukherjee S (2013c) Channel flow extrusion model to constrain dynamic viscosity and Prandtl
number of the Higher Himalayan Shear Zone. International Journal of Earth Sciences 102,
1811–1835
Mukherjee S (2014a) Review of flanking structures in meso- and micro-scales. Geological
Magazine 151, 957–974
Mukherjee S (2014b) Atlas of shear zone structures in meso-scale. Springer Geology, Cham, pp 1–
124. ISBN 978-3-319-0088-6
Mukherjee S (2015) Atlas of structural geology. Elsevier, Amsterdam. ISBN: 978-0-12-420152-1
Mukherjee S (2017a) Shear heating by translational brittle reverse faulting along a single, sharp
and straight fault plane. Journal of Earth System Science 126(1)
Mukherjee S (2017b) Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences 106, 1453–1468
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Mukherjee S, Khonsari MM (2017) Brittle rotational faults and the associated shear heating.
Marine and Petroleum Geology 88, 551–554
Mukherjee S, Khonsari MM (2018) Inter-book normal fault-related shear heating in brittle
bookshelf faults. Marine and Petroleum Geology 97,45–48
Mukherjee S, Koyi HA (2010a) Higher Himalayan Shear Zone, Sutlej Section—structural geology
& extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences 99, 1267–1303
Mukherjee S, Koyi HA (2010) Higher Himalayan Shear Zone, Zanskar Section—microstructural
studies & extrusion mechanism by a combination of simple shear & channel flow. International
Journal of Earth Sciences 99, 1083–1110
Mukherjee S, Mulchrone KF (2013) Viscous dissipation pattern in incompressible Newtonian
simple shear zones: an analytical model. International Journal of Earth Sciences 102, 1165–
1170
Mukherjee S, Punekar JN, Mahadani T, Mukherjee R (2015) A review on intrafolial folds and their
morphologies from the detachments of the western Indian Higher Himalaya. In: Mukherjee S,
Mulchrone KF (eds) Ductile shear zones: from micro- to macro-scales. Wiley Blackwell,
pp 182–205
Mukhopadhyay MK, Gangopadhyay PK (1971) Structural characteristics of rocks around
Kalimpong, West Bengal. Himalayan Geology 1, 213–230
Mukhopadhyay MK, Gangopadhyay PK (1973) Rotational Garnets And Quartz Micro-Fabric in Si
and Se: An Example from Metamorphic Rocks around Takdah, Darjeeling District, West
Bengal. Himalayan Geology 3, 162–175
Mulchrone KF, Mukherjee S (2015) Shear senses and viscous dissipation of layered ductile simple
shear zones. Pure and Applied Geophysics 172, 2635–2642
Mulchrone KF, Mukherjee S (2016) Kinematics and shear heat pattern of ductile simple shear
zones with ‘slip boundary condition’. International Journal of Earth Sciences 105, 1015–1020
Passchier CW, Trouw, RAJ (2005) Microtectonics, 2nd edn. Springer, Berlin. pp 1–366. ISBN-10
3-540-64003-7
440 S. Banerjee et al.
Ray S (1947) Zonal metamorphism in the eastern Himalaya and some aspects of local geology.
Quarterly Journal of the Geological, Mining and Metallurgical Society of India 19, 117–140
Sen A (1971) Preliminary observations on the Banded Darjeeling gneiss occurring around
Sukhiapokhri, Darjeeling district, West Bengal. Himalayan Geology 1, 276–278
Sen A (1973) Structural features of the rocks in Sukhiapokhri-Bijanbari region, Darjeeling District,
West Bengal. Himalayan Geology 3, 357
Singh NK (1972) Petrochemistry of the rocks around Kurseong. Darjeeling District,
WBHimalayan Geology 2, 502–514
Sinha Roy S (1973) Gondwana pebble-slate in the Rangit Valley tectonic window, Darjeeling
Himalayas and its significance. Bulletin-Geological Society of India 14, 31–39
Wager LR (1939) The Lachi Series of Northern Sikkim. Records of the Geological Survey of India
74, 171–188
Pb—Isotopic Characterization of Major
Indian Gondwana Coalfields:
Implications for Environmental
Fingerprinting and Gondwana
Reconstruction
1 Introduction
Lead (Pb) isotopic ratios (IR) can be exploited as a mighty tracer for the global
atmospheric Pb dispersal derived from natural as well as anthropogenic sources.
Geological origin controls the characteristic isotopic composition of Pb, which is a
unique property. This property is used to trace Pb sources in the environment
(Kylander et al. 2010). Since the phasing out of leaded gasoline from India in the
year 2000, coal—combustion remains the principal source of Pb and mercury
pollution in the environment, which in turn is responsible for severe health prob-
lems in the Indian scenario.
Pb as a trace element can be present in coal in the form of its own minerals such
as galena (PbS) or clausthalite (PbSe) or as an admixture of Ba minerals, in the
pyrite crystalline structure substituting for Fe, or can be bonded to organic matter in
low rank coals (Swaine 1990; Finkelman 1994). The geological changes and the
atmospheric inputs during coalification also control the origin of trace elements in
the coal. The interactions in the milieu of geological, biochemical and physio—
chemical factors are controlling factors for concentration and distribution of trace
elements in coals. Hence, it is evident that the trace elements in coal may have
diverse origins (Swaine 1990). Knowledge of the mode of occurrence of a haz-
ardous element, such as lead, is of great interest because it may help us to
understand its behaviour during coal conversion processes and its potentially
hazardous activity. Moreover, interest in lead is enhanced by the fact that by
comparing LIRs in ores, contaminant sources and contaminated sites, it can be used
as a tracer for environmental management (Díaz-Somoano et al. 2007, 2009).
Apart from environmental application, Pb isotopes have the potential for
stratigraphic correlation also. The high precision radiometric dates constrain the
absolute age of Formation. Likewise, Sr—isotopes, d13C and d18O measurements
provide meticulous correlation constrains (Veizer et al. 1999; Melchin and
Holmden 2006; Jones et al. 2010). It is well established that geological origin
controls the characteristic isotopic composition of Pb, which is a unique property.
In the ambit of this, the LIRs can be used for stratigraphic correlation and recon-
struction also. The occurrence of coal globally dates during the Gondwana period.
The coal deposits from India and the deposits from Antarctica can be similarly
correlated in global context of Gondwana reconstruction on the basis of LIRs.
The present work aims at the measurement of high precision LIRs of Indian
Gondwana coals by Multi-Collector ICPMS housed at the Geological survey of
India, Kolkata. This dataset will help in establishing a background for identifying
relative contribution of coal, especially those used in coal-based thermal plants for
the lead pollution in India. Not only this, these data set can be extended to have
wider geological implications such as Gondwana reconstruction.
2 Regional Geology
3 Methodology
Coal samples have been collected from actively mined Gondwana coal deposits
spread all over India. The coalfields covered in the purview of this study are Talcher
and Ib Valley Coalfield (40 Samples), Rajmahal Coalfield (20 samples), Wardha
444 R. Kumar et al.
Valley Coalfield (18 samples), Godavari Valley Coalfield (18 samples) and Korba
Coalfield (14 samples). Over all 110 coal samples have been analysed in this study
covering the E, S and central Indian Coalfields with the objective to delineate inter
as well as intra—seam variations, if at all present which can be traced after anal-
yses. All the samples are bituminous to sub—bituminous coal and feed the major
thermal power plants (TPPs) for power generation. In addition to the coal samples,
fly ash samples of major TTPs in the corresponding area have also been analysed
for Pb-isotopic composition. Figure 1 shows the disposition of Indian Gondwana
Coal deposits and location of the samples collected.
The physical and chemical processing methodology adopted for the determination
of Pb—isotopic composition for coal samples was dry ashing and mineralization
process as per Díaz-Somoano et al. (2007). The coal samples were pulverised in
agate motor and dried to constant weight at room temperature. After this step the
samples were finally dried in a muffle furnace (*90 °C; 2 h) to remove the
moisture. One gm of coal sample was taken and then mineralized under controlled
conditions using a programmable furnace by the dry ashing with a temperature
increase of 1 °C per minute to a maximum temperature of 500 °C. The total
mineralization time was 10 h after attaining the required temperature. The ash—
content was determined gravimetrically. The ash was then dampened with one drop
of Milli-Q deionized water before treating with the 5 mL of concentrated HF and
0.5 mL of concentrated HClO4 in savillex vials and left overnight. The mixture of
acids was evaporated to near dryness and this step was repeated. After attaining the
incipient moisture—content, the residue was dissolved in 2 mL of concentrated
HNO3, transferred to a 100 mL volumetric flask, filled to the mark with Milli-Q
deionised water and transferred to a 100 mL PP Nalgene bottle. This stock solution
was later diluted to 2% (v/v) HNO3 for determination of the LIRs. All the exper-
iments were performed with considerable number of duplicate and procedural
blanks to determine possible contamination from reagents and general handling.
The fly ash samples were dried to remove moisture content and then 0.2 g of
sample has been digested using an acid solution of HCl:HNO3:HF::3:3:2 using
Anton Paar Multiwave 3000 Microwave digestion system. The mixture of acids and
fly ash was evaporated to near dryness and this step was repeated again (Zhou et al.
2014). The residue was dissolved in 2 ml of concentrated HNO3, transferred to a
100 ml volumetric flask, filled to the mark with MilliQ water and transferred to a
100 ml HDPE bottle.
These chemically processed samples of coal and fly ash have been analysed for
LIRs using Nu Plasma II MC ICPMS housed at LAMCI facility of G & IG Division
of the Geological Survey of India, Kolkata. The instrument is equipped with 22
collectors (17 F Collectors and 5 Ion Counters). Routine measurements made in
static mode using the Faraday detectors with 1011 X resistors and Ion counting
Pb—Isotopic Characterization of Major Indian Gondwana … 445
channels with ion yield *1.0 mv per ppm. The typical operating parameters of
measurement are provided in Table 1.
Samples and standards were adjusted to a Pb/Tl ratio of 9:1 (after Weiss et al.
2004).The concentration of Tl solution was maintained at 8 ppb. The analytical
protocol followed during the course of analyses was standard-sample-standard
bracketing method using two Standards (SRM 981 & 983) after every five to seven
unknown solutions. NIST SRM-981 are 206Pb/204Pb = 16.934 ± 0.007,
207
Pb/204Pb = 15.486 ± 0.012, and 208Pb/204Pb = 36.673 ± 0.033 (n = 15),
respectively, which matches well with the recalibrated magnitudes: 16.9322,
15.4855, and 36.6856, respectively. Similarly, for NIST SRM 983 ratios are
204
Pb/206Pb = 0.000298, 207Pb/206Pb = 0.070906, and 208Pb/206Pb = 0.013095
(n = 20) respectively, which matches well with the recalibrated magnitudes:
0.000371, 0.071201, and 0.013619, respectively. (Todt et al. 1996; Belinda et al.
2000; Tanimizu and Ishikawa 2006 references rearranged with oldest at first and
youngest at last). The contribution of 204Hg in the measurement of 204Pb was
periodically monitored and mathematical correction was performed in the data
acquisition program (called as the NICE: Nu Instruments Calculation Editor: File in
Nu Plasma II) to account for any small contribution from Hg on the signal at 204Pb
(typically <300 counts s−1, or below 0.4% of the total 204Pb signal). The precision
of the measurements by MC-ICP-MS was below 0.07% for all measured isotope
ratios. The corresponding blank Pb-content value obtained during the analyses span
has been in the range of 10–11 ng ml−1 (White et al. 2000; Das et al. 2016 this
reference attached as protocols followed are similar to this for our Laboratory).
4 Results
207 208
Samples Pb/206Pb Pb/206Pb Pb—content (mg kg−1)
Mean Max Min RSD Mean Max Min RSD Mean Max Min RSD
Talcher & Ib valley 0.8220 0.8531 0.7215 0.0266 2.0863 2.1655 1.9484 0.0192 40.94 72.73 15.40 0.34
coalfields (n = 40)
Rajmahal coalfield 0.8022 0.8467 0.7748 0.0233 2.0104 2.0560 1.9746 0.0140 74.60 160.89 23.48 0.44
(n = 20)
Godavari valley 0.7761 0.8219 0.7150 0.0449 2.0382 2.0767 1.9858 0.0130 237.74 565.91 54.67 0.79
coalfield (n = 18)
Wardha valley 0.8303 0.8731 0.8050 0.0170 2.0735 2.0998 2.0373 0.0069 180.98 437.92 39.13 0.59
coalfield (n = 18)
Korba coalfield 0.8514 0.8845 0.8050 0.0270 2.1442 2.2231 2.0725 0.0209 18.43 42.10 3.20 0.68
(n = 14)
East Indian coals 0.8263 0.8007 0.8518 2.0837 2.1606 2.0067 238 440 37
(Kumar et al. 2016b,
35th IGC)
Kolaghat fly ash 0.8726 ± 0.0007 – – – 2.1542 ± 0.0016 – – – – – – –
Bandel fly ash 0.8608 ± 0.0005 – – – 2.1382 ± 0.0008 – – – – – – –
Korba fly ash 0.8478 ± 0.0010 – – – 2.1070 ± 0.0002 – – – – – – –
Talcher fly ash 0.8668 ± 0.0001 – – – 2.1534 ± 0.0007 – – – – – – –
Khaiperkheda fly 0.8757 ± 0.0001 – – – 2.1706 ± 0.0003 – – – – – – –
ash
Raniganj coalfield 0.8162 0.8275 0.8007 0.0097 2.0602 2.0875 2.0175 0.230 – – – –
(Das et al. 2016) (Std (Std
Dev) Dev)
Jharia coalfield (Das 0.8396 0.8573 0.8191 0.0148 2.0946 2.1308 2.0543 0.0280 – – – –
et al. 2016) (Std (Std
Dev) Dev)
(continued)
R. Kumar et al.
Table 2 (continued)
207 208
Samples Pb/206Pb Pb/206Pb Pb—content (mg kg−1)
Mean Max Min RSD Mean Max Min RSD Mean Max Min RSD
Chinese Coal 0.8479 0.8661 0.8326 – 2.0892 2.141 1.99 – 40 400 – –
(Zhang et al. 2007;
Cheng and Hu 2010)
World Coal 0.8322 0.8546 0.8036 – 2.0567 2.1013 1.9822 – – – – –
(Díaz-Somoano
et al. 2009)
Indonesian Coal 0.8449 0.8474 0.8417 – 2.0929 2.1013 2.0878 – – – – –
(Díaz-Somoano
et al. 2009)
Chinese fly ash 0.868 0.867 0.855 – 2.13 2.157 2.103 – – – – –
(Chen et al. 2005)
Pb—Isotopic Characterization of Major Indian Gondwana …
447
448 R. Kumar et al.
Fig. 2 Compiled three ratio plot of Indian Coal and fly ash in comparison to the published values
of coal and fly ash elsewhere
materials analysed in the present work are coal and fly ash, hence, it is apposite to
represent the results in terms of 207Pb/206Pb (or 206Pb/207Pb) and 208Pb/206Pb ratios.
The results obtained in the present wok as well as dataset available from published
literature are summarised in Table 2. The compiled three ratio plot of Indian Coal
and Fly ash in comparison to the published values of coal and fly ash elsewhere are
shown in Fig. 2.
The coal deposits of Talcher and Ib valley are situated in the eastern part of India
and follow the Mahanadi graben along which the coal has developed. Total 40
samples have been analysed from these coal—deposits collected from working
seams and act as feed to the National Thermal Power Corporation (NTPC) Super
Thermal Power plant at Talcher. These coal also act as feed in the Bandel and
Kolaghat Thermal Power Plant. The 207Pb/206Pb ratio range between 0.7215 and
0.8531 with a mean value of 0.8220 whereas the 208Pb/206Pb ratio ranges between
1.9484 and 2.1655 with a mean value of 2.0863. The Pb-concentration measured of
Pb—Isotopic Characterization of Major Indian Gondwana … 449
these coal—deposits lie between 15.40 and 72.73 mg kg−1 with a mean value of
40.94 mg kg−1.
The coal-deposits of Rajmahal are situated in the eastern part of India where the
coal development has taken place along the main Rajmahal—Purnea Basin. The
analysed coal samples (n = 20) from Rajmahal Coalfield have isotopic signatures in
range of 0.7748–0.8467 with mean value of 0.8022 and 1.9746–2.0560 with mean
value of 2.0104 for the 207Pb/206Pb and 208Pb/206Pb ratios respectively. This
coalfield is also the supplier of coal to Bandel and Kolagahat Thermal Power Plant
for power generation.
The coal deposits of Godavari Valley have been developed along the course of
river Godavari. On the Gondwana Basin’s map of India (Fig. 1); in fact two
coalfields have a linear extension from NW to SE. The NE extension of the coal-
field is along Wardha River and hence, known as Wardha Valley Coalfield and the
SE extension is known as the East Godavari Valley Coalfield. The 18 coal samples
analysed from Godavari Valley Coalfield have 207Pb/206Pb ratios in range of
0.7150–0.8219 with mean value of 0.7761 and 208Pb/206Pb ratios in range of
1.9858–2.0767 with mean value of 2.0382. The concentration of Pb in this coalfield
obtained in range 238–566 mg kg−1 with a mean value of 238 mg kg−1.
The 18 coal samples analysed from Wardha Valley Coalfield have 207Pb/206Pb
ratios in range of 0.8050–0.8731 with mean value of 0.8303 and 208Pb/206Pb ratios
in range of 2.0373–2.0998 with mean value of 2.0735. The concentration of Pb in
this coalfield ranges 39–438 mg kg−1 with a mean value of 181 mg kg−1.
The Korba Coalfield forms the central Indian Gondwana Coal deposits.
Quarriable potentiality of the seam of Barakar Formation and its occurrence at
shallow depth (8.80 m) further enhances its importance. It is 16.82–18.74 m thick
and occurs in the western part of the area only in between two NNE-SSW faults. The
fourteen samples analysed from this coalfield has isotopic composition in range
0.8050–0.8845 with mean value of 0.8514 and 2.0725–2.2231 with mean value of
2.1442 for the 207Pb/206Pb and 208Pb/206Pb ratios respectively. The Pb—concen-
tration of these coal samples are in range of 3.2–42 mg kg−1 with mean value of
18.5 mg kg−1.
During the course of this work fly ash samples have also been analysed from the
power plants using the coal deposits as feed. The Kolaghat, Bandel and NTPC
Talcher Termal Power Plants have their source in coalfields of Raniganj, Jharia,
Talcher, Ib and Rajmhal. The NTPC Korba has its source for power generation in
Korba Coalfield. The coals of Wardha valley coalfields act as feed to the
Khaiperkheda Thermal Power Plant. The results obtained are presented in Table 2.
450 R. Kumar et al.
5 Discussion
As is evident from the previous section that the LIRs of Indian Gondwana coal—
deposits ranges 0.7150–0.8845 for 207/206 Pb and 1.9484–2.2231 for 208/206 Pb
with Pb concentration ranges 3.2–566 mg kg−1. The variability of the Pb—content
and its isotopic composition can be attributed to high and variable ash—content of
the coal seams including its pyrite—content and/or more radiogenic behaviour. It
might be worth noting that the data obtained in the present work compares well
with the records from the Chinese coals (Shanghai, Nanjing and Huinan Coalfield)
that have Pb—content varying 40–400 ppm with a 206Pb/207Pb IR variation of
0.8326–0.8661 and 208Pb/206Pb IR values of 1.99–2.141 (Zhang et al. 2007; Cheng
and Hu 2010).
Similarly the LIRs obtained in the fly ash analyses ranges 0.8478–0.8757 for
207/206 Pb and 2.1070–2.1706 for 208/206 Pb. These values have an increasing
trend with respect to the feed coal Pb isotopic compositions when compared to the
respective coal—deposits. Also the dataset obtained complies well with the pub-
lished literature which reports the Pb isotopic composition for Chinese fly ash,
which ranges 0.855–0.867 for 207/206 and 2.103–2.157 for 208/206, respectively
(Chen et al. 2005).
Uranium can be bonded to the organic material of the coal and also to the inorganic
components (Swaine 1990). In this case, the isotopic composition i.e. LIRs of the
coal is a result of mixing of the isotope ratios of common Pb in organic material at
the time of its sedimentation and the radioactive products of the subsequently
transported U and Th. Because of the predominance of 238U in natural isotopic
mixtures, the fingerprints of 206Pb are substantially affected by systems highly
enriched in U. This anomalous isotopic composition of Pb and the influence of U on
the Pb IR are the properties used to characterize and ultimately “fingerprinting/
tracing” the source.
The present work reports and reviews the isotopic composition i.e., the LIRs of
the Indian Gondwana basin Coal deposits. These dataset aim at establishing a
background for identifying relative contribution of coal, especially those used in
coal—fired thermal plants for the Pb pollution in India. To test this hypothesis, the
LIRs data generated by Das et al. (2016) for the different anthropogenic contrib-
utors of Pb in Kolkata City and LIRs reported by Kumar et al. (2016c) in the dust
storm aerosols of Indian subcontinent has been used for tracing/fingerprinting the
pollutant in the township of Kolkata. The dataset obtained by Das et al. (2016) are
on a gamut of food items (such as Rice, Dal, Spices, Vegetables, Fish, Meat,
Chicken and Herbs), street dust, diesel and rainwater of Kolkata city. By applying
binary mixing model on these dataset along with data reported by Kumar et al.
Pb—Isotopic Characterization of Major Indian Gondwana … 451
(2016a, b) has shown that coal—combustion is not the major contributor in the
atmosphere of Kolkata City instead the diesel is the principal contributor in the
form of vehicular exhaust. Thus using the LIR in coal samples from East Indian
Gondwana coal deposits, a logical clue for “tracing/fingerprinting” the Pb pollution
source in the Indian scenario particularly for the Kolkata City has been achieved
(Das et al. 2016; Kumar et al. 2017).
It is further recommended that LIRs of environmental milieu/matrices should be
taken up in different power—hub Indian cities and the inventory on LIRs of Indian
coal should be compared for getting a novel qualitative as well as quantitative
“fingerprinting” of environmental pollution and contribution from anthropogenic
sources.
Apart from these environmental aspect, the LIR data set generated in the present
work particularly of the Mahanadi and Godavari valley coal—deposits can have
implications on the Gondwana reconstruction.
The Gondwana sequence in the East Antarctica is restricted to the area around
Beaver Lake in the Lambert graben. The sequence is known as the Amery Group
(Mond 1972) and represents >4000 m thick sequence of coal—bearing conglom-
erate—sandstone—minor shale association. Broadly, the sequence is >400 m thick
Upper Permian Radok Conglomerate in the basal part, followed by *3000 m thick
coals bearing Bainmedart Coal Measures which is overlain by around 750 m thick
Triassic Flagstone Bench Formation. The lithofacies and palynology of the rocks of
the Amrey Group has been studied (Playford 1990; Webb and Fielding 1993;
McLoughlin and Drinnan 1997), but the information on lead- isotopic composition
of this Group of rocks is not yet reported.
In the Gondwana reconstruction, there is a consensus that the *1.0 Ga mobile
belts of the Eastern Ghats in India and the Eastern Reyner Complex in Antarctica
were parts of a suture zone associated with Rhodinia Formation. However, these
belts occur parallel to the boundaries of the continents and hence using them in
Gondwana reconstruction could be superfluous. The Gondwana basins of the
Amery Group of the Lambert Graben and those of the Mahanadi and Godavari
basins in India that are aligned transverse to the continental margins are favoured
for a tighter Gondwana reconstruction. Available literature postulates correlation of
Lambert Graben with Godavari basin on the basis of regional geology and coal
petrology (Holdgate et al. 2005) while the correlation model with Mahanadi basin is
favoured based on geophysics and bathymetry (Stagg 1985; Mishra et al. 1999)
(Fig. 3). Since geophysics and bathymetry gets much influenced due to
post-breakup events in the seafloor, a stronger correlation of the Lambert Graben
with either the Mahanadi or the Godavari basin would require further constraints.
The distinctive Pb—isotopic compositional variation between the coals from the
Mahanadi basin and the coals from the Godavari—Wardha basin as obtained in the
452 R. Kumar et al.
present work and similar data set from Amrey Group and its coal deposit could offer
a good constrain to achieve a better fit between India and Antarctica in a Gondwana
framework. Moreover, the possible reconstruction—hypotheses proposed between
Indian and East Antarctica can also be tested with the aid of LIRs in coal of these
two presently widely separated terrains (Stagg 1985; Boger and Wilson 2003;
Harrowfield et al. 2005).
6 Conclusions
References
Acharyya SK (2000) Tectonic setting and nature of the Gondwanic Indian crust. Proc. Vol., Int.
Seminar, Precambrian Crust in Eastern and Central India. Geological Survey of India, Special
Publication 57, 1–8
Biswas SK (1999) A review on the evolution of rift basins in India during Gondwana with special
reference to western Indian basins and their hydrocarbon prospects. In: Sahni A, Loyal RS
(eds) Gondwana assembly: new issues and perspectives. proceedings of indian national science
academy special issue 65, pp 261–283
Boger SD, Wilson CJL (2003) Brittle faulting in the Prince Charles Mountains, East Antarctica:
cretaceous transtensional tectonics related to the break-up of Gondwana. Tectonophysics 367,
173–186
Chen JM, Tan MG, Li YL, Zhang YM, Lu WW, Tong YP, Zhang GL, Li Y (2005) A lead isotope
record of Shanghai atmospheric lead emissions in total suspended particles during the period of
phasing out of leaded gasoline. Atmospheric Environment 39, 1245–1253
Cheng H, Hu Y (2010) Lead (Pb) isotopic fingerprinting and its applications in lead pollution
studies in China: a review. Environmental Pollution 158, 1134–1146
Das A, Krishna KVSS, Kumar R, Das A, Sengupta S, Ghosh JG (2016) Tracing lead
contamination in foods in the city of Kolkata, India. Environmental Science and Pollution
Research 23, 22454–22466
Díaz-Somoano M, Suárez-Ruiz I, Alonso JIG, Ruiz Encinar J, López-Anton MA,
Martínez-Tarazona MR (2007) Lead isotope ratios in Spanish coals of different characteristics
and origin. International Journal of Coal Geology 71, 28–36
Díaz-Somoano M, Kylander ME, Lopez-Anton MA, Suarez-Ruiz I, Martínez Tarazona MR,
Ferrat M, Kober B, Weiss DJ (2009) Stable lead isotope compositions in selected coals from
around the world and implications for present day aerosol source tracing. Environmental
Science and Technology 43, 1078–1085
Finkelman RB (1994) Modes of occurrence of potentially hazardous elements in coal: levels of
confidence. Fuel Processing Technology 39, 21–34
Flem B, Grimstvedt A, Cook N (2000) Lead isotope determinations by inductively-coupled plasma
mass spectrometry (ICP-MS): potential of sector field instruments. NGU-BULL 436, 203–207
Harrowfield M, Holdgate RG, Wilson CJL (2005) Tectonic significance of the Lambert graben,
East Antarctica: Reconstructing the Gondwanan rift. Geology 33, 197–200
454 R. Kumar et al.
Holdgate GR, McLoughlin S, Drinnan AN, Finkelman RB, Willett JC (2005) Inorganic chemistry,
petrography and palaeobotany of Permiancoals in the Prince Charles Mountains, East
Antarctica. International Journal of Coal Geology 63, 156–177
Jones DS, Maloof AC, Hurtgen MT, Rainbird RH, Schrag DP (2010) Regional and global
chemostratigraphic correlation of the early Neoproterozoic Shaler Supergroup, Victoria Island,
Northwestern Canada. Precambrian Research 181, 43–63
Kumar R, Patel SS, Manjhi JK (2016a) Pb-isotopic characterization of major Indian Coalfields and
their potential as heavy metal pollutants. GSI Unpublished Report
Kumar R, Patel SS, Ghosh JG (2016b) Variability of Pb-isotopes in the East Indian Gondwana
Coal deposits: its influence on Kolkata street dust. 35TH IGC ABSTRACT
Kumar S, Aggarwal SG, Malherbe J, Barre JPG, Berial S, Gupta PK, Donard (2016c) Tracing dust
transport from Middle-East over Delhi in March 2012 using metal and lead isotope
composition. Atmospheric Environment 132, 179–187
Kumar R, Patel SS, Das A, Ghosh JG, Sengupta S, Krishna KVSS, Guha D (2017) Variability of
Pb-isotopes in the East Indian Gondwana Coal deposits: It’s influence on Kolkata street dust.
Indian Journal of Geoscience 71(4), 575–588
Kylander ME, Klaminder J, Bindler R (2010) Natural Lead Isotope Variations in the Atmosphere.
Earth and Planetary Science Letters 290, 44–53
McLoughlin S, Drinnan AN (1997) Revised stratigraphy of the Permian Bainmedart coal
measures, northern Prince Charles Mountains, East Antarctica. Geological Magazine 134, 335–
353
Melchin MJ, Holmden CE (2006) Carbon isotope chemostratigraphy in Arctic Canada: sea-level
forcing of carbonate platform weathering and implications for Hirnantian global correlation.
Palaeogeography, Palaeoclimatology, Palaeoecology 234, 186–200
Mishra DC, Chandra Sekhar DV, Venkata Raju DCh, Vijay Kumar V (1999) Crustal structure
based on gravity-magnetic modelling constrained from seismic studies under Lambert Rift,
Antarctica and Godavari and Mahanadi rifts, India and their interrelationship. Earth and
Planetary Science Letters 172, 287–300
Mond A (1972) Permian sediments of the Beaver Lake area, Prince Charles Mountains. In:
Adie RJ (ed) Antarctic geology and geophysics. Universitetsforlaget, Oslo, pp 585–589
Mukherjee S (2019) Introduction to “Tectonics and Structural Geology: Indian Context”. In:
Mukherjee S (ed) Tectonics and structural geology: Indian context. Springer International
Publishing AG, Cham, pp 1–5. ISBN: 978-3-319-99340-9
Playford G (1990) Proterozoic and Palaeozoic palynology of Antarctica: a review. In: Taylor TN,
Taylor EL (eds) Antarctic paleobiology—its role in the reconstruction of Gondwana. Springer
Verlag, New York. Plumstead, pp 50–70
Raja Rao CS (1982) Coalfields of India. Vol-II, coal resources of Tamil Nadu, Andhra Pradesh,
Orissa and Maharashtra. Geological Survey of India Bulletin Series A, 45
Raja Rao CS (1983) Coalfields of India. Vol‐III, Coal resources of Madhya Pradesh and Jammu
and Kashmir. Geological Survey of India Bulletin Series A, 45
Raja Rao CS (1987) Coalfields of India. Vol-IV, Part I. Coal resources of Bihar. Geological
Survey of India Bulletin Series A, 45
Stagg HMJ (1985) The structure and origin of Prydz Bay and MacRoberstonShelf. East Antarctica:
Tectonophysics 114, 315–340
Swaine DJ (1990) Trace elements in coal. Butterworths and Co. Publishers
Tanimizu M, Ishikawa T (2006) Development of rapid and precise Pb isotope analytical
techniques using MC-ICP-MS and new results for GSJ rock reference samples. Geochemical
Journal 40, 121–133
Todt W, Cliff RA, Hanser A, Hofmann AW (1996) Evaluation of a 202Pb/205Pb double spike for
high precision lead isotope analysis in earth processes: reading the isotopic code. American
Geophysical Union, pp 429–437
Veevers JJ, Tewari RC (1995) Gondwana master basin of peninsular India between Tethys and the
interior of the Gondwanaland province of Pangea. Memoir of Geological Society of America
187, 1–73
Pb—Isotopic Characterization of Major Indian Gondwana … 455
Veizer J, Ala D, Azmy K, Bruckschen P, Buhl D, Bruhn F, Carden GAF, Diener A, Ebneth S,
Godderis Y, Jasper T, Korte C, Pawellek F, Podlaha O, Strauss H (1999) 87Sr/86Sr, 13C and
18 O evolution of Phanerozoic seawater. Chemical Geology 161, 59–88
Webb A, Fielding R (1993) Permo-Triassic sedimentation within the Lambert Graben, northern
Prince Charles Mountains, East Antarctica. In: Findlay RH, Uruug R, Banks MR, Veevers JJ
(eds) Gondwana eight: assembly, evolution and dispersal. A. A. Balkema, Rotterdam, pp 357–
369
Weiss D, Kober B, Dolgopolova A, Gallagher K, Spiro B, Le Roux G, Mason TFD, Kylander M,
Coles BJ (2004) Accurate and precise Pb isotope ratio measurements in environmental samples
by MCICP- MS. International Journal of Mass Spectrometry 232, 205–215
White WM, Albaredeb F, Teloukb P (2000) High-precision analysis of Pb-isotope by
multicollector ICP-MS. Chemical Geology 167, 257–270
Zhang GL, Yang FG, Zhao WJ, Zhao YG, Yang JL, Gong ZT (2007) Historical change of soil Pb
content and Pb isotope signatures of the cultural layers in urban Nanjing. CATENA 69, 51–56
Zhou C, Liu G, Cheng S, Fang T, Lam PKS (2014) The environmental geochemistry of trace
elements and naturally radionuclides in a coal gangue brick-making plant. Scientific Reports 4,
6221
Correction to: Geomorphic
Characteristics and Morphologic Dating
of the Allah Bund Fault Scarp, Great
Rann of Kachchh, Western India
Correction to:
Chapter “Geomorphic Characteristics and Morphologic
Dating of the Allah Bund Fault Scarp, Great Rann
of Kachchh, Western India” in: S. Mukherjee (ed.),
Tectonics and Structural Geology: Indian Context, Springer
Geology, https://doi.org/10.1007/978-3-319-99341-6_3
The original version of the book was inadvertently published without incorporating
the author corrections in figure captions 5 & 6 in chapter “Geomorphic
Characteristics and Morphologic Dating of the Allah Bund Fault Scarp, Great Rann
of Kachchh, Western India”. The correction chapter and the book have been now
updated with the changes.