You are on page 1of 10

Journal of Volcanology and Geothermal Research 424 (2022) 107496

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


journal homepage: www.journals.elsevier.com/journal-of-volcanology-and-geothermal-research

Tectonics, cooling rates and temperatures during emplacement of the


Rajmahal traps, India
Amar Agarwal a, *, L.M. Alva-Valdivia b, Arnaldo Hernández-Cardona b, Roshan A. Shukla a,
Gaurav Joshi a, K.K. Agarwal c
a
Department of Earth Sciences, Indian Institute of Technology-Kanpur, Kanpur 208016, India
b
Universidad Nacional Autónoma de México, Instituto de Geofísica, Laboratorio de Paleomagnetismo, Ciudad Universitaria, 04510 Ciudad de México, Mexico
c
Department of Geology, University of Lucknow, Lucknow 226007, India

A R T I C L E I N F O A B S T R A C T

Keywords: Paleomagnetism is used to decipher the emplacement setting of the Rajmahal traps. Sampling was restricted to
Rajmahal traps, India flows with known age. Three magnetic minerals are identified, titanomaghemite, Ti-poor magnetite, and possibly
Paleomagnetism Mg–Al-rich titanomagnetite. Detailed analysis of thermomagnetic curves revealed that the emplacement tem­
Low emplacement temperature
perature for Rajmahal traps was lower than 600 ◦ C. Two magnetization components are distinguished. The low
Fast cooling rate
coercivity component, probably chemical remanent magnetization, is erased in fields up to 3.1 mT. In contrast,
Penecontemporaneous tectonics
the high coercivity thermal remanent magnetization is more stable, sometimes throughout the demagnetization.
The paleomagnetic results show that the sites have suffered up to 11◦ of clockwise and 17◦ of counter-clockwise
rotation. As there is no field evidence of faulting and pervasive deformation, we propose that this rotation is
probably due to penecontemporaneous joints or faults. These structures could have acted as conduits for fluids
that increased the cooling rate.

1. Introduction reconstruction of the Indian subcontinent. The present study builds upon
the established paleomagnetic aspects of the Rajmahal Traps. This study
Regional-scale normal faults run along the western and eastern sampled flow with known age, geochemistry, and mineralogy. The re­
boundaries of the Rajmahal traps, which downthrow the region in the sults of this study quantify the magnetic mineralogy, domain size and
middle, forming a graben (Singh et al., 2004; Tiwari and Jassal, 2001). coercivity distribution. Then, the emplacement conditions, such as the
Together with the Damodar Gondwana graben, these faults represent a emplacement temperatures (Te), cooling rates and the role of tectonics
sort of triple junction at the mouth of the Bengal basin (Burke and in the emplacement, are determined.
Dewey, 1973). This geometry led to the argument that the Rajmahal
volcanism is related to the lithospheric thinning and continental rifting 2. Geological setup of the Rajmahal Traps
(Mukhopadhyay, 2000; Mukhopadhyay et al., 1986). However, other
workers have suggested that the traps originated from the Kerguelen The Rajmahal Traps (24◦ 15′ to 25◦ 15′ N, 87◦ 20′ to 87◦ 45′ E) extend
hotspot (Curray and Munasinghe, 1991; Kapawar and Mamilla, 2020; over an area of ~4100 km2 in the Jharkhand and West Bengal states of
Müller et al., 1993; Subrahmanyam et al., 1999). India (Fig. 1). High precision whole-rock 40Ar/39Ar dating reveals that
The paleomagnetic characteristics of the Rajmahal traps have been the eruption occurred between 113 and 118 Ma (Kent et al., 2002; Ray
reported in detail by a series of reports (Kapawar and Mamilla, 2020; et al., 2008). On the western side, they rest over the Precambrian
Kent et al., 1997; Klootwijk, 1971; McDougall and McElhinny, 1970; basement of the Chota Nagpur terrain with a sharp igneous contact
McElhinny and Luck, 1970; Sherwood and Malik, 1996). These studies marked by silicic tuffs and bentonite deposits (Das Gupta, 1996). At
have used the standard sampling strategy and have covered the entire places, sediments of the Gondwana Supergroup are present between the
traps, sampling various flows. This strategy has proven helpful in the volcanic and the Precambrian basement along the western margin of the
paleomagnetic dating of the traps and for the paleogeographic Rajmahal Traps. On the eastern side, they are covered by Quaternary

* Corresponding author.
E-mail address: amar@daad-alumni.de (A. Agarwal).

https://doi.org/10.1016/j.jvolgeores.2022.107496
Received 21 September 2021; Received in revised form 21 January 2022; Accepted 30 January 2022
Available online 4 February 2022
0377-0273/© 2022 Elsevier B.V. All rights reserved.
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

and the recent alluvium of the Ganga–Brahmaputra delta (Kale, 2020; least five demagnetization steps and a maximum angular deviation
Kent et al., 1997). (MAD) below 5◦ . Individual magnetization directions were averaged per
The Rajmahal Traps are primarily, 75%, tholeiitic basalts (Baksi site, and statistical parameters were calculated assuming a Fisher dis­
et al., 1987; Kale, 2020; Kent et al., 1997). Rhyolitic lava flows, tribution (Fisher, 1953). The ChRM directions were used to calculate the
outcropping at places like Taljhari and Berhait, contribute a smaller net rotation of the sites with respect to each other, using the methods
portion of ~24–28 km2 (Deshmukh et al., 1964; Kale, 2020; Rao and detailed in the appendix of Butler (1992). Determining the vertical axis
Purushottam, 1963). The traps comprise 28 lava flows having a rotation has proven to be effective in revealing and confirming the ef­
maximum thickness of 332 m, seen in the Kansat borehole, West Bengal fects of faulting (e.g., Hornafius, 1985; Kamerling and Luyendyk, 1979;
(Das Gupta and Mukherjee, 2006; Mukhopadhyay et al., 1986). The Luyendyk et al., 1985; Wilson and Cox, 1980). The demagnetization
flows are sometimes separated by intertrappean sedimentary beds. intensity plots were used to calculate the median destructive field
These centimetre to meter thick beds are made up of sandstone, silt­ (MDF), which allows identifying the relative magnetic hardness of the
stone, arenaceous clay bed, carbonaceous and siliceous shale and tuffs, samples (e.g., Agarwal and Alva-Valdivia, 2019).
and are characterized by a large variety of fossils but are mainly known The variation in low-field magnetic susceptibility at high-
for their paleofloral assemblage like Ptilophyllum (Kale, 2020; Sahni, temperature (ᴋ-T) was measured by equipping the Kappabridge
1932; Tripathi et al., 2013). The flows are massive, compact and are (KLY–4S) with a CS-3 furnace. The sample was heated from room
devoid of vesicles (Fig. 2). The southern limit of the traps is unknown temperature up to 700 ◦ C at a rate of 10 ◦ C/min and then cooled back.
due to the burial beneath the Bengal Fan (Kent et al., 1997). The bulk magnetic susceptibility was measured during the entire cycle of
heating and cooling. Possible oxidation while heating was minimized by
3. Sampling and methods flushing out air from the sample holder using argon flowing at 50 ml/
min. Curie temperature (TC) was marked at the maximum negative
We collected 26 oriented blocks from 12 paleomagnetic sites, mostly slope. We also calculated ΔTc, which equals Tc from the heating curve
in the northern part of the Rajmahal Traps. Each site represents an ‘in minus Tc from the cooling curve. Tc, ΔTc and paleomagnetic measure­
situ’ volcanic unit, avoiding sites with possible fractures and/or high ments are used to calculate the cooling rate and Te of the rocks (e.g.,
alteration. Cylindrical cores were drilled from the blocks to recover six Agarwal et al., 2017; Alva-Valdivia et al., 2019c; Hernández-Cardona
to ten paleomagnetic specimens (2.54 cm in diameter and 2.2 cm long) et al., 2019; Lied et al., 2020).
from each site. The drilling was done using a laboratory setup with a Hysteresis curves were determined using Alternating Gradient Field
nonmagnetic diamond bit at the Magnetic Observatory, Prayagraj, Magnetometer Micromag apparatus in fields up to 1.4 T. The remanent-
India. Some specimens broke during drilling, and a total of 68 oriented coercivity to coercive-force (Hcr/Hc) and the remanent-magnetization
paleomagnetic specimens were recovered for further investigations to saturation-magnetization (Mrs/Ms) ratios were determined from the
(Table 1). curves. These ratios are used to estimate the relative proportions of the
The specimens were demagnetized in an AF Molspin demagnetizer single domain (SD), multi-domain (MD), and pseudo-single domain
(Molspin Limited, England) and non-inductive Schonstedt furnace and (PSD) grains (Dunlop, 2002a, 2002b). In general, the ratio of Mrs/Ms is
measured using a JR-6 spinner magnetometer (AGICO, Czech Republic) higher for more SD materials and lower for more MD-like materials. The
in the magnetically shielded room. Alternating field (AF) demagneti­ stepwise increasing uniaxial field is applied to the specimens to calculate
zation was performed in peak fields up to 96.1 mT. Prior to demagnet­ the saturated and non-saturated isothermal remanent magnetization
izing all the specimens, initially, two pilot specimens from each site were (IRM) acquisition curves.
demagnetized by AF treatment. 10–12 steps were used to determine the UNMIX analysis (Egli, 2005) was done on IRM curves to calculate the
best laboratory strategy to obtain the characteristic remanent magneti­ coercivity distribution of the ferromagnetic minerals in the samples. The
zation (ChRM) directions. The ChRM directions are determined using at investigation was done using the CODICA software, which is part of the

Fig. 1. Geological map showing the distribution of Rajmahal Traps and the adjacent lithologies along with the sample locations (after Singh et al., 2004). The map is
prepared using QGIS software and Mapbox satellite streets v9 is used as the base map.

2
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

Fig. 2. Field photographs of outcrops of Rajmahal traps from which the oriented block samples were collected.

MAG - MIX routine to process the IRM data (Egli, 2005). While the
Table 1
hysteresis and Day plot reveal the bulk behaviour, i.e., the dominant
Compilation of paleomagnetic sites, specimens, nearby village (location), age
domain size, the UNMIX processing allows determining the individual
and geographical location. The ages are taken from Kent et al. (2002).
coercivity component. Combined with the thermomagnetic curves, they
Site Specimen Location Age (my) Latitude Longitude are a powerful tool for a more comprehensive understanding of mag­
113.6 ± N 25◦ 11′ E 087◦ 30′ netic behaviour (Alva-Valdivia et al., 2019a).
1 01–08 Ayodhya Pahar
0.4 52.5" 30.3" The pieces of the oriented block that were left after drilling the cy­
113.6 ± N 25◦ 16′ E 087◦ 28′
2 11–18 Dubauli lindrical cores, were used to prepare thin sections for transmitted light
0.4 08.3" 53.9"
Vindari 113.6 ± N 25◦ 02′ E 087◦ 37′ microscopy and polished sections for reflected light microscopy using a
3 19–30
bandarcola 0.4 35.1" 13.7" Leica DM2700P microscope. JEOL JSM-6010LA, W-Scanning electron
4 31–38 Chakjo Pahar
113.6 ± N 24◦ 53′ E 087◦ 44′ microscopy (SEM) and energy dispersive X-ray spectroscopy (EDS) were
0.4 36.5" 09.1" used to map the relative amounts of Si, Fe, Ti, Al and Mg present in and
Aamdanda 113.6 ± N 24◦ 53′ E 087◦ 44′
5 39–45
Pahar 0.4 43.9" 05.0"
around titanomagnetites.
Aamdanda 113.6 ± N 24◦ 53′ E 087◦ 44′
6 46–53
Pahar 0.4 44.5" 07.9" 4. Results and interpretations
113.6 ± N 24◦ 55′ E 087◦ 47′
7 54–62 Kelabaadi
0.4 30.7" 44.9"
113.6 ± N 24◦ 55′ E 087◦ 47′ 4.1. Mineralogy and texture
8 63–70 Rakso Pahar
0.4 42.6" 47.2"

9 71–78 Rakso Pahar


113.6 ± N 24◦ 55′ E 087◦ 47′ In the basalts of the Rajmahal traps, the phenocrysts are mostly
0.4 30.6" 44.9" labradorite feldspars and pyroxenes, which are clustered together,
98.5 ± N 25◦ 12 ′
E 087◦ 39 ′
10 79–86 Lohanda Vedo imparting glomeroporphyritic texture (Figs. 3a and b). Plagioclase
0.5 45.5" 38.6"
98.5 ± N 25◦ 12′ E 087◦ 39′ exhibit characteristic Carlsbad and Albite twinning, and zoning with a
11 87–94 Lohanda Vedo
0.5 47.5" 36.4" homogeneous core surrounded by a zoned rim. Most clinopyroxene
12 95–102 –
118.1 ± N 24◦ 23′ E 087◦ 28′ phenocrysts are subhedral to euhedral augite. The groundmass is made
0.3 22.5" 56.6"
up of feldspars (andesine to labradorite) and pyroxenes, opaques, glass,
and secondary palagonite. Some plagioclase and pyroxenes are altered
to saussurite and biotite, respectively. The glass is more or less devitri­
fied and exhibits a spherulitic texture.
Oxide minerals from the sampled rock units of Rajmahal Traps

3
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

Fig. 3. (a and b) Photomicrographs presenting large


plagioclase and smaller pyroxene phenocrysts embedded
in a fine matrix. (c) Long strands of ilmenite and subhedral
titanomagnetites are common at sites 1 to 3. (d) Sites 4 to
6 present subhedral and skeletal titanomagnetite without
ilmenite. (e) Sites 7 to 11 present mostly skeletal and
cruciform titanomagnetite. (f) At site 12, the subhedral
titanomagnetite grain presents partial replacement with
titanohematite and titanomaghemite along the fractures.

belong to the spinel–magnetite series. Generally, subhedral titano­ magma. Smaller grain sizes ranging from 5 to 40 μm indicate low to
magnetite and magnetite minerals are found as fine- to coarse-grained medium oxidation state. The most frequently observed textures of tita­
dispersed crystals in the matrix of almost all the sites, hematite and nomagnetite are skeletal and cruciform homogeneous to subhedral
maghemite are scarce (Figs. 3a to 3f). We followed Buddington and particles of varying grain sizes indicating low oxidation state (C1-C3 and
Lindsley (1964) and Haggerty (1976) classification for textures and R1 to R3). The titanomagnetite is altered to titanohematite and titano­
oxidation states, respectively. maghemite along the fractures.
The most common oxide mineral textures are: homogeneous C1-C3
and R1-R3 subhedral, skeletal to anhedral magnetite and ilmenite par­
ticles that likely formed from deuteric oxidation during cooling of

Table 2
Paleomagnetic site, specimens (Sp.), number of specimens collected (N), Curie temperature (Tc), median destructive field (MDF). The number of specimens used for
calculating mean paleo directions is represented by ‘n’. Dec and Inc. represent the mean declination and inclination at each site; ‘α95’and ‘k’ are the 95% of confidence
limit, dispersion parameter, respectively. ‘Lat’ and ‘Long’ show the calculated virtual geomagnetic pole (VGP) coordinates for each site.
Site Sp. N Tc Heating [◦ ] Tc Cooling [◦ ] MDF Paleo direction [◦ ] VGP [◦ ]
[mT]
Tc(1) Tc(2) Tc(3) Tc(1) Tc(2) Tc(3) n Dec Inc α95 κ Lat Long

1 01–08 8 239 304 543 – – 543 225.7 8 302.2 − 69.2 1.2 265.62 2.7 118
2 11–18 0 233 – 560 – – 560 Specimen broken
3 19–30 8 – 302 490 – – 490 169.6 8 312 − 61.3 1.6 96.01 − 9.3 121
4 31–38 0 196 450 517 286 450 517 Unstable directions
5 39–45 7 – 338 539 – – 539 64.7 4 312.9 − 70.5 3.5 49.74 − 0.7 112
6 46–53 8 – – 494 – – 494 233.0 8 320.6 − 59.6 0.8 503.77 − 15 117
7 54–62 0 263 – 509 – – 509 Unstable directions
8 63–70 8 – 318 500 – – 500 115.5 8 327.2 − 69.7 2.9 57.39 − 6.6 106
9 71–78 8 240 – 487 – – 487 120.8 8 298.3 − 54.1 3.1 34.18 − 6.6 134
10 79–86 8 – 333 518 – – 518 97.0 8 302.4 − 54.9 2 80.55 − 8.5 131
11 87–94 8 274 402 – – 472 – 230.9 8 303.8 − 56.7 1 392.85 − 8.2 129
12 95–102 5 – – 495 – – 495 34.6 Unstable directions

4
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

4.2. Magnetic transition temperatures Some loops present higher saturation remanence than others, perhaps
due to relatively higher SD components (Tauxe, 2003). In most cases, the
Rajmahal traps present complex temperature-dependent magnetic branch of the loop with an increasing applied field overlaps with the
susceptibility curves (ᴋ-T). These curves reveal up to three magnetic branch representing the decreasing or negative applied field. This usu­
transition temperatures. The first transition temperature, Tc1, ranges ally indicates that the magnetization is carried by low coercivity min­
between 196 ◦ C and 286 ◦ C and is never reversible (Table 2, Fig. 4). This erals (Tauxe et al., 2002), such as titanomaghemite (Lowrie, 1990). The
transition temperature may be the Tc of the Mg–Al-rich titanomagnetite dominance of PSD behaviour at all sites is also evident from the Day plot
(Lied et al., 2020). SEM-EDS analysis indicates that there might be (Fig. 5). Site 8 and 9 have the highest SD component (c. 80%), while site
heterogeneous distribution of Mg and Al within the titanomagnetite 12 has the highest MD component (c. 90%).
grains (Supplementary Fig. 1). Tc1 is not observed in any cooling curves, Each peak in the UNMIX curves represents an individual coercivity
except for a unique instance (Site 4). The cooling curves without Tc1 component (Fig. 6, supplementary fig. 3). The Rajmahal traps have
indicate the transformation/destruction of the Mg–Al-rich titano­ single to multiple coercivity components. For example, site 2 has a single
magnetite at temperatures over 274 ◦ C into a lower susceptibility phase, peak revealing a unique component. Two components are conspicuous
thus causing lower susceptibility in the cooling curves (Fig. 4, Supple­ in Site 5 and subtle in Site 3 and 9. The lower coercivity component may
mentary Fig. 2). represent titanomaghemite, while the higher coercivity component may
The second transition temperature, Tc2, varies from 302 to 450 ◦ C in be owed to titanomagnetite.
the heating curve (Table 2). The cooling curve has lower susceptibility in
this temperature range, and Tc2 is absent in all but two cooling curves 4.4. Intensities and directions of remanent magnetization
(Sites 4 and 11). This indicates the transformation of this phase to a
paramagnetic mineral, like ilmenite. This second phase could be tita­ The magnetization intensities vary from 16.9 A/m for specimen 82 of
nomaghemite with a transition temperature between 350 and 550 ◦ C. site 10 to 0.5 A/m for specimen 26 of Site 3. The MDF varies from 3.46
Similar behaviour of titanomaghemite is reported from the lava flows of mT for site 12 to 23.3 mT for site 6 (Fig. 7). In the present case, low MDF
the El Jorullo volcano, Mexico (Alva-Valdivia et al., 2019b) and the can be attributed to larger magnetic domain size, as Site 12, with the
ilmenite bearing Olmec artefacts (Alva-Valdivia et al., 2017b). lowest MDF has the largest MD population (Fig. 5). Low MDF indicates
The third transition temperature, Tc3, is stable and occurs at similar magnetically soft behaviour, which usually results in unstable paleo­
temperatures in all sites, except Site 11. Tc3 ranges from 487 to 560 ◦ C. magnetic directions (e.g., Agarwal et al., 2016). In agreement site 12
The curves are reversible beyond Tc3. Tc3 is, therefore, related with a presents unstable directions, one of the lowest Tc3 (Table 2,.
relatively stoichiometric titanomagnetite with a composition close to Some specimens have unstable paleomagnetic directions (type-I).
the end-member magnetite with ulvöspinel content ranges around 0.1 to These specimens were excluded from the calculation of the ChRM. Most
0.0 (e.g., Alva-Valdivia et al., 2017a; Lattard et al., 2006). other sites present two magnetization directions (e.g., Site 3 in Fig. 8).
The lower coercivity component (type-II) accounts for about 10% of the
4.3. Magnetic coercivity and domain size total remanent magnetization. It demagnetizes at relatively lower fields,
up to 3.1 mT. Note that 3.1 mT was the first step of demagnetization
The general shape of the hysteresis loops is wasp-waisted, and the used. The type-III magnetization component with higher coercivity re­
magnetization in all loops is saturated in the applied field up to 1.4 T mains stable throughout the demagnetization process, i.e., applied fields
(Fig. 5). The loops, therefore, indicate dominance of PSD component. up to 96.1 mT (e.g., sites 3 and 9, Fig. 8). This component represents the

Fig. 4. Thermomagnetic curves present the variation in


magnetic susceptibility with temperature in four repre­
sentative specimens. The heating and cooling curves are
represented in red and blue, respectively. The Curie tem­
peratures, Tc1 and Tc2, are observed in just the heating
curve, while Tc3 is observed in both heating and cooling
curves. Curves from other samples are given in the sup­
plementary material. (For interpretation of the references
to colour in this figure legend, the reader is referred to the
web version of this article.)

5
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

Fig. 5. Hysteresis of one specimen from each paleomagnetic site. Note the different vertical scales. The specimens are also plotted in the Day Plot (Day et al., 1977).

Fig. 6. Characteristic UNMIX curves presenting different coercivity components (see supplementary fig. 3 for other paleomagnetic sites). Shaded area represents the
error in the coercivity estimation.

ChRM acquired as the thermal remanent magnetization during the 5. Discussion


cooling of the basalts.
The declination of the ChRM varies from 302.2◦ to 327.2◦ and the 5.1. Cooling rate and emplacement temperature
inclination from − 54.1◦ to − 69.7◦ (Figs. 8, 9, Supplementary Fig. 4).
These values are comparable to previously published directions for Our ᴋ-T measurements reveal a complex behaviour with up to three
Rajmahal traps (Kapawar and Mamilla, 2020; Kent et al., 1997, 2002; magnetic transition temperatures and irreversible heating–cooling
Klootwijk, 1971; McDougall and McElhinny, 1970; McElhinny and Luck, curves. Based on the transition temperatures, we can distinguish two
1970; Poornachandra Rao et al., 1996; Sherwood and Malik, 1996). types of ᴋ-T curves. Type-I curves are characterized by lower Tc1 in the
cooling curve compared to the heating. They have a ΔTc1 of − 90 ◦ C.
Previous studies have demonstrated a correlation of ΔTc with cooling
rates or Te, and interpreted those results in terms of cation ordering and

6
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

5.2. Characteristic remanent directions

Three types of remanent directions are identified in specimens from


the Rajmahal Traps. Type-I is unstable and excluded from the calcula­
tion of the ChRM. Type-II demagnetizes at applied alternating-field
below 3.1 mT, indicating a lower coercivity magnetic carrier. Based
on the Tc2 in the ᴋ-T curves, we attribute the type-II remanent magne­
tization to titanomaghemite. Type-II magnetization may represent
either viscous or chemical remanent magnetization.
The third type-III remanent directions are stable, sometimes
throughout the AF demagnetization, i.e., in applied fields up to 96.1 mT.
They represent the ChRM acquired during cooling and emplacement.
This component is due to SD magnetite, based on the Tc3 in the ᴋ-T and
high Mrs/Ms ratios (sites 3 and 9 in Figs. 4 and 5).
The ᴋ-T curves reveal that site 9 has two ferrimagnetic minerals,
Fig. 7. The removal of magnetization due to stepwise alternating field
demagnetization. The field at which 50% of magnetization remains is the me­ which may be Mg–Al-rich titanomagnetite and relatively stoichiometric
dian destructive field. Sites 2, 4 and 7 are not included. titanomagnetite (Fig. 4). Site 9 has a single remanent component, type-
III, which is stable (Fig. 8). The magnetization is therefore carried only
by the stoichiometric titanomagnetite.
disordering (Jackson and Bowles, 2014; Lied et al., 2020; Redfern et al.,
1996; Stimpfl et al., 2005). For example, a ΔTc of − 60 ◦ C is known to
correspond with a cooling rate faster than 11 ◦ C / min (Lied et al., 2020). 5.3. Emplacement setting
Therefore, a lower ΔTc (− 90 ◦ C), as in the present case, may indicate a
cooling rate faster than 11 ◦ C / min. There is a certain degree of un­ Discordant paleomagnetic directions may reflect the vertical axis
certainty in these estimates due to compositional sensitivity. rotation of the sites (Butler, 1992). The net rotation suffered by the
Type-II ᴋ-T curves are owed to titanomaghemite and are character­ basalts is calculated using the ChRM directions and the apparent polar
ized by irreversible behaviour. Previous studies have shown that tita­ wander path (Torsvik et al., 2012). See appendix of Butler (1992) for
nomaghemite either transforms into Ti-poor titanomagnetite and
ilmenite (in the ambient atmosphere) or into titanomagnetite (in argon
atmosphere) during heating at about 425 ◦ C (Lied et al., 2020). In type-II
curves, maghemitization of titanomagnetite is completely masked by the
effects of cation ordering.
Te of ignimbrites from the 1980 Mt. St Helens eruption was directly
measured by Banks and Hoblitt (1981, 1996). At location where Te was
below 600 ◦ C the ignimbrites contained mainly Al- and Mg-bearing
titanomagnetites (Tc < 450 ◦ C); but as Te exceeded 600◦ the ignim­
brites contained oxyexsolved composites of nearly pure magnetite and
Ti-, Al- and Mg-bearing phases (Tc ≥ 525 ◦ C). To summarize, the Mt. St.
Helens and Novarupta deposits with Te > 600 ◦ C have Tc ≥ 525 ◦ C, and
those with Te < 600 ◦ C have Tc < 450 ◦ C (Bowles et al., 2013). Using the
same analogy and considering the presence of Al- and Mg-bearing tita­
nomagnetites in our samples which causes low Tc, we propose that the
Te for the studied flows of the Rajmahal traps was lower than 600 ◦ C,
which was the Te of Mt. St. Helens and Novarupta deposits. Our
emplacement temperature estimates are similar to those by Lied et al.
(2020) for their samples with similar curves and maghemitization. Fig. 9. 3D orthographic projection of the paleomagnetic poles. The shapefile
However, different redox conditions in basalt flows as compared to ig­ for the Asia map is taken from https://tapiquen-sig.jimdofree.com/, and the
nimbrites may render some uncertainty to these estimates. VGPs are visualized using GPlates software (Müller et al., 2018).

Fig. 8. (Left) The mean characteristic remanent magnetization directions (squares) and the α95 confidence ellipse (black circles) are presented in the lower
hemisphere stereographic projection. (Right) Orthogonal projections of the paleomagnetic directions were obtained during alternating field demagnetization of two
representative specimens.

7
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

details. The sites have suffered up to 11◦ of clockwise, and 17◦ of Traps, varies from 15◦ S to 3◦ N and 106◦ E to 134◦ E (Table 2, Fig. 9).
counter-clockwise rotation, which indicates both right and left lateral However, individual sites have suffered up to 11◦ of clockwise, and 17◦
movement. Similar vertical axis rotation in Xalapa basalts, Mexico, was of counter-clockwise rotation, which indicates both right and left lateral
attributed to extensional faulting and Riedel shear in the region (Alva- movement. Owing to the absence of pervasive deformation and faulting
Valdivia et al., 2019a). Therefore, we argue that the rotation of the sites in the traps, it is safe to conclude that the faulting was
is due to extensional faulting and Riedel shear. The reports of exten­ penecontemporaneous.
sional faulting from the margins of the Rajmahal traps support this Our results showed that the emplacement temperature for Rajmahal
argument (Singh et al., 2004; Tiwari and Jassal, 2001). There is no field traps was lower than 600 ◦ C and the cooling rate were faster than 11 ◦ C /
evidence of pervasive deformation in the traps. The rotation may, thus, min. These temperatures and cooling rates may be due to mixing of
be attributed to faulting during the emplacement of the traps when the basalts with the continental crust (e.g., Kent et al., 1997), and interac­
magnetic minerals had already solidified. Recent reports have shown tion with fluids traversing through the faults at the time of emplace­
that the early Rajmahal eruptions took place in an extensional tectonic ment. However, sensitivity towards composition of titanomagnetites
setting accompanied by rifting (Ghose et al., 2017). It is likely that the and different redox conditions in basalt flows as compared to ignim­
extensional tectonics continued and affected the Rajmahal flows. brites may render some degree of uncertainty to these estimates.
Both earlier proposed theories of the origin of the traps imply either a
direct or indirect role of faulting. First theory relates the emplacement of CRediT authorship contribution statement
the Rajmahal traps to lithospheric thinning and continental rifting (e.g.,
Mukhopadhyay, 2000; Mukhopadhyay et al., 1986). This theory directly Amar Agarwal: Conceptualization, Methodology, Formal analysis.
implies the significance of extensional faulting in the emplacement of L.M. Alva-Valdivia: Conceptualization, Methodology, Funding acqui­
the traps. The second theory is that the traps originated from the Ker­ sition. Arnaldo Hernández-Cardona: Formal analysis, Investigation.
guelen hotspot (e.g., Curray and Munasinghe, 1991; Müller et al., 1993; Roshan A. Shukla: Formal analysis, Investigation. Gaurav Joshi:
Subrahmanyam et al., 1999). Notably, the rise of the magma plume is Formal analysis, Investigation. K.K. Agarwal: Funding acquisition.
also usually accompanied by extensional faulting (Maruyama, 1994; Su
et al., 2011). The second theory of the origin of the Rajmahal Traps,
thus, also indicates the role of extensional faulting in their emplacement. Declaration of Competing Interest
The present results, for the first time, reveal the influence of the
extensional faulting on the emplacement of Rajmahal Traps. Faulting The authors declare that they have no known competing financial
not only affected the paleomagnetic directions, but most likely, also the interests or personal relationships that could have appeared to influence
cooling rates and the emplacement temperatures. We argue that, pene­ the work reported in this paper.
contemporaneous faults may have provided conduits for the movement
of fluids that would decrease the emplacement temperature and increase Acknowledgements
the cooling rates.
Some previous studies have attributed the lower emplacement tem­ National Autonomous University of Mexico and Department of Ge­
peratures to mixing with the continental crust (Kent et al., 1997). Which ology (CAS program phase II), Lucknow University are thanked for
is evident from the high silica content >68 wt% (Ghose et al., 2017). supporting the fieldwork. Dr. S.K. Patil of I.I.G.M., Allahabad provided
Similar crustal mixing with only heat supplied by the basaltic magma access to drilling facilities. Ms. S. Tiwari processed the UnMix curves.
has also been demonstrated in parts of the Deccan basalts (Chatterjee IITK initiation grant is thanked for financial support for the analysis. Dr.
and Bhattacharji, 2001; Sheth et al., 2011). Ranjit Singh of the Department of Geology, Sahibganj College, is
The geochemical studies of the Rajmahal Traps have revealed a po­ thanked for his guidance during the fieldwork. Dr. Jesús Roberto Vidal-
tential mantle temperature of about 1350 ◦ C (Kent et al., 1997) and two- Solano and Dr. Mike Jackson are thanked for their constructive reviews
pyroxene geothermometer temperature of about 1120 ◦ C (Sarkar et al., that improved the manuscript. Prof. Heidy M Mader provided editorial
1989). Present results show that the emplacement temperature basalts guidance.
were lower than 600 ◦ C and the cooling rate at some sites was faster than
11 ◦ C / min. These cooling rates are fast for basalts, even though natural Appendix A. Supplementary data
cooling rates of pyroclastic and hyaloclastic fragments span over several
orders of magnitude, from 106 to 10− 2 ◦ C s− 1 (Helo et al., 2013; Nichols Supplementary data to this article can be found online at https://doi.
et al., 2009; Potuzak et al., 2008; Wilding et al., 1996). org/10.1016/j.jvolgeores.2022.107496.
Rajmahal traps contain intertrappean layers of clastic sedimentary
and volcaniclastic rocks, pockets and lenses of bentonite, and thin black References
shale/mudstone and oolitic beds, which were attributed to subaqueous
eruptive phases in parts of the Rajmahal traps (e.g., Ghose et al., 2017; Agarwal, A., Alva-Valdivia, L.M., 2019. Curie temperature of weakly shocked target
basalts at the Lonar impact crater, India. Earth Planets Sp. 71, 141. https://doi.org/
Pascoe, 1973). Interaction with water caused high temperature deuteric 10.1186/s40623-019-1120-9.
oxidation of titanomagnetite to titanohematite and pseudobrookite in Agarwal, A., Kontny, A., Srivastava, D.C., Greiling, R.O., 2016. Shock pressure estimates
some parts of the traps (Sherwood and Malik, 1996). In our samples the in target basalts of a pristine crater: a case study in the Lonar crater, India. Geol. Soc.
Am. Bull. 128 (B31172), 1. https://doi.org/10.1130/B31172.1.
deuteric oxidation is most likely evidenced by homogeneous C1-C3 and
Agarwal, A., Alva-Valdivia, L.M., Rivas-Sánchez, M.L., Herrero-Bervera, E., Urrutia-
R1-R3 subhedral, skeletal to anhedral magnetite and ilmenite particles. Fucugauchi, J., Espejel-García, V., 2017. Emplacement dynamics and hydrothermal
Interaction of water with basaltic magma may lead to magmatic alteration of the Atengo ignimbrite, southern Sierra Madre Occidental, northwestern
Mexico. J. S. Am. Earth Sci. 80, 559–568. https://doi.org/10.1016/j.
explosivity, contact-surface steam explosivity, bulk interaction steam
jsames.2017.10.017.
explosivity, or cooling-contraction granulation (Kokelaar, 1986). In fact, Alva-Valdivia, L.M., Agarwal, A., Caballero-Miranda, C., García-Amador, B.I., Morales-
the high viscosity magma of Rajmahal traps led to the explosive erup­ Barrera, W., Rodríguez-Elizarraráz, S., Rodríguez-Trejo, A., 2017a. Paleomagnetic
tions in the north-central and the northeastern margin (Ghose et al., and AMS studies of the El Castillo ignimbrite, central-east Mexico: source and rock
magnetic nature. J. Volcanol. Geotherm. Res. 336, 140–154. https://doi.org/
2017). These locations are close to the present sampling sites. 10.1016/j.jvolgeores.2017.02.014.
Alva-Valdivia, L.M., Cyphers, A., De La Luz Rivas-Sánchez, M., Agarwal, A., Zurita-
6. Conclusions Noguera, J., Urrutia-Fucugauchi, J., 2017b. Mineralogical and magnetic
characterization of Olmec ilmenite multi-perforated artifacts and inferences on
source provenance. Eur. J. Mineral. 29, 851–860. https://doi.org/10.1127/ejm/
The paleomagnetic poles of 113.6 ± 0.4 My flow of the Rajmahal 2017/0029-2654.

8
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

Alva-Valdivia, L.M., Agarwal, A., García-Amador, B., Morales-Barrera, W., Agarwal, K.K., Kamerling, M.J., Luyendyk, B.P., 1979. Tectonic rotations of the Santa Monica
Rodríguez, S., Gonzalez-Rangel, J.A., 2019a. Paleomagnetism and tectonics from the Mountains region, western Transverse Ranges, California, suggested by
late Pliocene to late Pleistocene in the Xalapa monogenetic volcanic field, Veracruz, paleomagnetic vectors. Geol. Soc. Am. Bull. 90, 331–337.
Mexico. GSA Bull. 131, 1581–1590. https://doi.org/10.1130/B32006.1. Kapawar, M.R., Mamilla, V., 2020. Paleomagnetism and rock magnetism of early
Alva-Valdivia, L.M., Rodríguez-Trejo, A., Morales, J., González-Rangel, J.A., Agarwal, A., Cretaceous Rajmahal basalts, NE India: implications for paleogeography of the
2019b. Paleomagnetism and age constraints of historical lava flows from the El Indian subcontinent and migration of the Kerguelen hotspot. J. Asian Earth Sci. 201,
Jorullo volcano, Michoacán, Mexico. J. S. Am. Earth Sci. 93, 439–448. https://doi. 104517 https://doi.org/10.1016/j.jseaes.2020.104517.
org/10.1016/j.jsames.2019.05.016. Kent, W., Saunders, A.D., Kempton, P.D., Ghose, N.C., 1997. Rajmahal Basalts, Eastern
Alva-Valdivia, L.M., Rodríguez-Trejo, A., Vidal-Solano, J.R., Paz-Moreno, F., Agarwal, A., India: Mantle sources and melt distribution at a volcanic rifted margin. In:
2019c. Emplacement temperature resolution and age determination of Cerro Mahoney, J.J., Coffin, M.F. (Eds.), Large Igneous Provinces: Continental, Oceanic,
Colorado tuff ring by paleomagnetic analysis, El Pinacate Volcanic Field, Sonora, and Planetary Flood Volcanism. Washington, D. C, pp. 145–182. https://doi.org/
Mexico. J. Volcanol. Geotherm. Res. 369, 145–154. https://doi.org/10.1016/j. 10.1029/GM100p0145.
jvolgeores.2018.11.012. Kent, R.W., Pringle, M.S., Müller, R.D., Saunders, A.D., Ghose, N.C., 2002. 40Ar/39Ar
Baksi, A.K., Barman, T.R., Paul, D.K., Farrar, E., 1987. Widespread early cretaceous flood Geochronology of the Rajmahal Basalts, India, and their Relationship to the
basalt volcanism in eastern India: Geochemical data from the Rajmahal-Bengal- Kerguelen Plateau. J. Petrol. 43, 1141–1153. https://doi.org/10.1093/petrology/
Sylhet Traps. Chem. Geol. 63, 133–141. https://doi.org/10.1016/0009-2541(87) 43.7.1141.
90080-5. Klootwijk, C.T., 1971. Palaeomagnetism of the-Upper Gondwana-Rajmahal traps,
Banks, N.G., Hoblitt, R.P., 1981. Summary of temperature studies of 1980 deposits. The Northeast India. Tectonophysics 12, 449–467. https://doi.org/10.1016/0040-1951
1980 Eruptions of Mount St. Helens 295–313. (71)90045-X.
Banks, N.G., Hoblitt, R.P., 1996. Direct Temperature Measurements of Deposits, Mount Kokelaar, P., 1986. Magma-water interactions in subaqueous and emergent basaltic. Bull.
St. Helens, Washington, 1980–1981. US Geological Survey. Volcanol. 48, 275–289. https://doi.org/10.1007/BF01081756.
Bowles, J.A., Jackson, M.J., Berquó, T.S., Sølheid, P.A., Gee, J.S., 2013. Inferred time- Lattard, D., Engelmann, R., Kontny, A., Sauerzapf, U., 2006. Curie temperatures of
and temperature-dependent cation ordering in natural titanomagnetites. Nat. synthetic titanomagnetites in the Fe-Ti-O system: effects of composition, crystal
Commun. 4, 1916. https://doi.org/10.1038/ncomms2938. chemistry, and thermomagnetic methods. J. Geophys. Res. 111, B12S28. https://doi.
Buddington, A.F., Lindsley, D.H., 1964. Iron-titanium oxide minerals and synthetic org/10.1029/2006JB004591.
equivalents. J. Petrol. 5, 310–357. Lied, P., Kontny, A., Nowaczyk, N., Mrlina, J., Kämpf, H., 2020. Cooling rates of
Burke, K., Dewey, J.F., 1973. Plume-generated triple junctions: key indicators in pyroclastic deposits inferred from mineral magnetic investigations: a case study from
applying plate tectonics to old rocks. J. Geol. 81, 406–433. the Pleistocene Mýtina Maar (Czech Republic). Int. J. Earth Sci. https://doi.org/
Butler, R.F., 1992. Paleomagnetism: Magnetic Domains to Geologic Terranes. Blackwell 10.1007/s00531-020-01865-1.
Scientific Publications, Boston. Lowrie, W., 1990. Identification of ferromagnetic minerals in a rock by coercivity and
Chatterjee, N., Bhattacharji, S., 2001. Origin of the Felsic and Basaltic Dikes and Flows in unblocking temperature properties. Geophysical Research Letters 17 (2), 159–162.
the Rajula-Palitana-Sihor Area of the Deccan Traps, Saurashtra, India: a Geochemical https://doi.org/10.1029/GL017i002p00159.
and Geochronological Study. Int. Geol. Rev. 43, 1094–1116. https://doi.org/ Luyendyk, B.P., Kamerling, M.J., Terres, R.R., Hornafius, J.S., 1985. Simple shear of
10.1080/00206810109465063. southern California during Neogene time suggested by paleomagnetic declinations.
Curray, J.R., Munasinghe, T., 1991. Origin of the Rajmahal Traps and the 85◦ E Ridge: J. Geophys. Res. Solid Earth 90, 12454–12466.
preliminary reconstructions of the trace of the Crozet hotspot. Geology 19, 1237. Maruyama, S., 1994. Plume tectonics. J. Geol. Soc. Japan 100, 24–49. https://doi.org/
https://doi.org/10.1130/0091-7613(1991)019<1237:OOTRTA>2.3.CO;2. 10.5575/geosoc.100.24.
Das Gupta, S., 1996. Bentonite deposits intercalated with the Rajmahal volcanic rocks of McDougall, I., McElhinny, M.W., 1970. The Rajmahal Traps of India - KAr ages and
eastern India. J. SE Asian Earth Sci. 13, 133–137. https://doi.org/10.1016/0743- palaeomagnetism. Earth Planet. Sci. Lett. 9, 371–378. https://doi.org/10.1016/
9547(96)00014-1. 0012-821X(70)90138-X.
Das Gupta, A.B., Mukherjee, B., 2006. Geology of NW Bengal basin. Geol. Soc. India, McElhinny, M.W., Luck, G.R., 1970. Paleomagnetism and Gondwanaland. Science (80-.)
Bangalore 1–154. 168, 830–832. https://doi.org/10.1126/science.168.3933.830.
Day, R., Fuller, M., Schmidt, V.A., 1977. Hysteresis properties of titanomagnetites: grain- Mukhopadhyay, M., 2000. Deep crustal structure of the West Bengal basin deduced from
size and compositional dependence. Phys. Earth Planet. Inter. 13, 260–267. https:// gravity and DSS data. J. Geol. Soc. India 56, 351–364.
doi.org/10.1016/0031-9201(77)90108-X. Mukhopadhyay, M., Verma, R.K., Ashraf, M.H., 1986. Gravity field and structures of the
Deshmukh, S.S., Chakraborty, S.C., Murthy, M.V.N., 1964. The origin of the shapes and Rajmahal Hills: example of the Paleo-Mesozoic continental margin in eastern India.
sizes of the amygdules in the pitchstone and basalt flows from Taljhari and Berhait. Tectonophysics 131, 353–367. https://doi.org/10.1016/0040-1951(86)90182-4.
Santhal Parganas, Bihar. Rec. Geol. Surv. India 93, 45–50. Müller, R.D., Royer, J.-Y., Lawver, L.A., 1993. Revised plate motions relative to the
Dunlop, D.J., 2002a. Theory and application of the Day plot (Mrs / Ms versus Hcr / Hc) hotspots from combined Atlantic and Indian Ocean hotspot tracks. Geology 21,
1. Theoretical curves and tests using titanomagnetite data. J. Geophys. Res. 107, 275–278.
2056. https://doi.org/10.1029/2001JB000486. Müller, R.D., Cannon, J., Qin, X., Watson, R.J., Gurnis, M., Williams, S., Pfaffelmoser, T.,
Dunlop, D.J., 2002b. Theory and application of the Day plot (Mrs / Ms versus Hcr / Hc) Seton, M., Russell, S.H.J., Zahirovic, S., 2018. GPlates: building a virtual earth
2. Application to data for rocks, sediments, and soils. J. Geophys. Res. 107, 2057. through deep time. Geochem. Geophys. Geosyst. 19, 2243–2261. https://doi.org/
https://doi.org/10.1029/2001JB000487. 10.1029/2018GC007584.
Egli, R., 2005. User's Guide to the MAG-MIX, Magnetic Unmixing Software Packet, Nichols, A.R.L., Potuzak, M., Dingwell, D.B., 2009. Cooling rates of basaltic hyaloclastites
Release 1, April 2005 [WWW Document]. Free. Distrib. http://dourbes.meteo.be and pillow lava glasses from the HSDP2 drill core. Geochim. Cosmochim. Acta 73,
/aarch.net/magmix.man.pdf. 1052–1066. https://doi.org/10.1016/j.gca.2008.11.023.
Fisher, R., 1953. Dispersion on a Sphere. Proc. R. Soc. A Math. Phys. Eng. Sci. 217, Pascoe, E.H., 1973. A Manual of the Geology of India and Burma. Controller of
295–305. https://doi.org/10.1098/rspa.1953.0064. Publications.
Ghose, N.C., Chatterjee, N., Windley, B.F., 2017. Subaqueous early eruptive phase of the Poornachandra Rao, G.V.S., Mallikharjuna Rao, J., Subba Rao, M.V., 1996.
late Aptian Rajmahal volcanism, India: evidence from volcaniclastic rocks, Palaeomagnetic and geochemical characteristics of the Rajmahal Traps, eastern
bentonite, black shales, and oolite. Geosci. Front. 8, 809–822. https://doi.org/ India. J. SE Asian Earth Sci. 13, 113–122. https://doi.org/10.1016/0743-9547(96)
10.1016/j.gsf.2016.06.007. 00012-8.
Haggerty, S.E., 1976. Oxidation of opaque mineral oxides in basalts. Oxide Miner. Potuzak, M., Nichols, A.R.L., Dingwell, D.B., Clague, D.A., 2008. Hyperquenched
Hg1–Hg100. volcanic glass from Loihi Seamount, Hawaii. Earth Planet. Sci. Lett. 270, 54–62.
Helo, C., Clague, D.A., Dingwell, D.B., Stix, J., 2013. High and highly variable cooling https://doi.org/10.1016/j.epsl.2008.03.018.
rates during pyroclastic eruptions on Axial Seamount, Juan de Fuca Ridge. Rao, C.S.R., Purushottam, A., 1963. Pitchstone flows in the Rajmahal Hills, Santhal
J. Volcanol. Geotherm. Res. 253, 54–64. https://doi.org/10.1016/j. Parganas, Bihar. Rec. Geol. Surv. India 91, 341–348.
jvolgeores.2012.12.004. Ray, R., Shukla, A.D., Sheth, H.C., Ray, J.S., Duraiswami, R.A., Vanderkluysen, L.,
Hernández-Cardona, A., Alva-Valdivia, L.M., Acosta-Ochoa, G., Agarwal, A., Cruz-y- Rautela, C.S., Mallik, J., 2008. Highly heterogeneous Precambrian basement under
Cruz, T., McClung de Tapia, E., 2019. Unblocking temperature of secondary the Central Deccan Traps, India: direct evidence from xenoliths in dykes. Gondwana
magnetic component to outline anomaly thermomagnetic maps of archaeological Res. 13, 375–385. https://doi.org/10.1016/j.gr.2007.10.005.
fireplaces from the southern region of Mexico City, Mexico. Rev. Mex. Ciencias Redfern, S.A.T., Henderson, C.M.B., Wood, B.J., Harrison, R.J., Knight, K.S., 1996.
Geológicas 36, 411–418. https://doi.org/10.22201/cgeo.20072902e.2019.3.1528. Determination of olivine cooling rates from metal-cation ordering. Nature 381,
Hornafius, J.S., 1985. Neogene tectonic rotation of the Santa Ynez Range, Western 407–409. https://doi.org/10.1038/381407a0.
Transverse Ranges, California, Suggested by paleomagnetic investigation of the Sahni, B., 1932. A petrified Williamsonia (W. sewardiana, sp. nov.) from the Rajmahal
Monterey Formation. J. Geophys. Res. 90, 12503. https://doi.org/10.1029/ Hills, India. Alexander Doweld.
JB090iB14p12503. Sarkar, S.S., Nag, S.K., Basu Mallik, S., 1989. The origin of andesite from Rajmahal traps,
Jackson, M., Bowles, J.A., 2014. Curie temperatures of titanomagnetite in ignimbrites: eastern India: a quantitative evaluation of a fractional crystallization model.
Effects of emplacement temperatures, cooling rates, exsolution, and cation ordering. J. Volcanol. Geotherm. Res. 37, 365–378. https://doi.org/10.1016/0377-0273(89)
Geochem. Geophys. Geosyst. 15, 4343–4368. https://doi.org/10.1002/ 90090-5.
2014GC005527. Sherwood, G.J., Malik, S.B., 1996. A palaeomagnetic and rock magnetic study of the
Kale, V.S., 2020. Cretaceous volcanism in Peninsular India: Rajmahal–Sylhet and Deccan northern Rajmahal volcanics, Bihar, India. J. SE Asian Earth Sci. 13, 123–131.
Traps. In: Gupta, N., Tandon, S. (Eds.), Geodynamics of the Indian Plate, Sheth, H.C., Choudhary, A.K., Bhattacharyya, S., Cucciniello, C., Laishram, R., Gurav, T.,
pp. 233–289. https://doi.org/10.1007/978-3-030-15989-4_8. 2011. The Chogat-Chamardi subvolcanic complex, Saurashtra, northwestern Deccan

9
A. Agarwal et al. Journal of Volcanology and Geothermal Research 424 (2022) 107496

Traps: Geology, petrochemistry, and petrogenetic evolution. J. Asian Earth Sci. 41, Tauxe, L., Bertram, H.N., Seberino, C., 2002. Physical interpretation of hysteresis loops:
307–324. https://doi.org/10.1016/j.jseaes.2011.02.012. Micromagnetic modeling of fine particle magnetite. Geochemistry, Geophysics,
Singh, A.P., Kumar, N., Singh, B., 2004. Magmatic underplating beneath the Rajmahal Geosystems 3 (10), 1–22. https://doi.org/10.1029/2001GC000241.
Traps: Gravity signature and derived 3-D configuration. Proc. Indian Acad. Sci. Earth Tiwari, S., Jassal, G.S., 2001. Origin and evolution of the Garo-Rajmahal Gap. J. Geol.
Planet. Sci. 113, 759–769. https://doi.org/10.1007/BF02704035. Soc. India 57, 389–403.
Stimpfl, M., Ganguly, J., Molin, G., 2005. Kinetics of Fe2+-Mg order-disorder in Torsvik, T.H., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B.,
orthopyroxene: Experimental studies and applications to cooling rates of rocks. Doubrovine, P.V., van Hinsbergen, D.J.J., Domeier, M., Gaina, C., Tohver, E.,
Contrib. Mineral. Petrol. 150, 319–334. https://doi.org/10.1007/s00410-005-0016- Meert, J.G., McCausland, P.J.A., Cocks, L.R.M., 2012. Phanerozoic polar wander,
9. palaeogeography and dynamics. Earth-Sci. Rev. 114, 325–368. https://doi.org/
Su, B.-X., Qin, K.-Z., Sakyi, P.A., Li, X.-H., Yang, Y.-H., Sun, H., Tang, D.-M., Liu, P.-P., 10.1016/j.earscirev.2012.06.007.
Xiao, Q.-H., Malaviarachchi, S.P.K., 2011. U–Pb ages and Hf–O isotopes of zircons Tripathi, A., Jana, B.N., Verma, O., Singh, R.K., Singh, A.K., 2013. Early cretaceous
from Late Paleozoic mafic–ultramafic units in the southern Central Asian Orogenic palynomorphs, dinoflagellates and plant megafossils from the Rajmahal Basin,
Belt: tectonic implications and evidence for an Early-Permian mantle plume. Jharkhand, India. J. Palaeontol. Soc. India 58, 125–134.
Gondwana Res. 20, 516–531. https://doi.org/10.1016/j.gr.2010.11.015. Wilding, M., Webb, S., Dingwell, D., Ablay, G., Marti, J., 1996. Cooling rate variation in
Subrahmanyam, C., Thakur, N.K., Rao, T.G., Khanna, R., Ramana, M.V., natural volcanic glasses from Tenerife, Canary Islands. Contrib. Mineral. Petrol. 125,
Subrahmanyam, V., 1999. Tectonics of the Bay of Bengal: new insights from satellite- 151–160.
gravity and ship-borne geophysical data. Earth Planet. Sci. Lett. 171, 237–251. Wilson, D., Cox, A., 1980. Paleomagnetic evidence for tectonic rotation of Jurassic
Tauxe, L., 2003. Paleomagnetic Principles and Practice, 1st ed. Kluwer Academic plutons in Blue Mountains, eastern Oregon. J. Geophys. Res. Solid Earth 85,
Publishers. 3681–3689.

10

You might also like