Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-20T01:12:33.275Z Has data issue: false hasContentIssue false

Tracing multiple sources of sediments using trace element and Nd isotope geochemistry: provenance of the Mesozoic succession in the Kutch Basin, western India

Published online by Cambridge University Press:  10 June 2020

Angana Chaudhuri
Affiliation:
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai400076, India
Anirban Chatterjee
Affiliation:
Department of Geology, Presidency University, Kolkata700073, India
Santanu Banerjee*
Affiliation:
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai400076, India
J.S. Ray
Affiliation:
Geosciences Division, Physical Research Laboratory, Navrangpura, Ahmedabad380009, India
*
Author for correspondence: Santanu Banerjee, Email: santanu@iitb.ac.in
Rights & Permissions [Opens in a new window]

Abstract

An integrated approach involving Sr–Nd isotope, trace and rare earth element analyses tracks multiple sources of the Mesozoic sediments of the Kutch Basin at the western continental margin of India. High (87Sr/86Sr)t (ratio at time of deposition), negative εNd and high concentrations of large-ion lithophile elements (LILEs) indicate the upper continental source. Ratios of Nb/Ta and Zr/Hf suggest sedimentary and felsic igneous sources of sediments. The moderate to high concentration of La, Th and Sc, light rare earth elements (LREE-) enrichment, weak negative Eu anomalies and the relationship between εNd(0) and Th/Sc indicate the dominantly felsic composition of source rocks. However, low contents of Th, low values of (87Sr/86Sr)t and depleted mantle model age TDM < 1600 Ma indicate input from a younger mafic source. Increasing concentrations of Zr, Hf and Nd isotopes and a gradual increase in mean TDM from the older to the younger formations indicate erosional unroofing at the source terrain. The increasing (87Sr/86Sr)t through time relates to increased weathering of the source rock. The overwhelmingly southwesterly palaeocurrent direction of current-generated sedimentary structures, and the mean TDM ages trace suggest source areas of the Kutch Basin to Precambrian rocks in the north and NE of this basin. The TDM ages highlight the dominance of late Palaeoproterozoic source rocks. Nd isotope composition indicates that Proterozoic rocks of Marwar Supergroup and Erinpura Granite, in particular, served as main sediment contributors for the Mesozoic sediments in Kutch. We therefore conclude that the Mesozoic sediments in the Kutch Basin are predominantly of late Palaeoproterozoic age with lesser inputs from rocks of early Mesoproterozoic and early Palaeoproterozoic age.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2020. Published by Cambridge University Press

1. Introduction

Sedimentary successions in pericratonic rift basins preserve the evolutionary history of adjacent cratons and orogenic belts (Dickinson & Suczek, Reference Dickinson and Suczek1979; Dickinson, 1988). Provenance interpretations of these sedimentary successions contribute to the understanding of tectonic setting, climate, extent of weathering and drainage patterns during the time of deposition (Nesbitt & Young, Reference Nesbitt and Young1982; Algeo & Maynard, Reference Algeo and Maynard2004; Tribovillard et al. Reference Tribovillard, Algeo, Lyons and Riboulleau2006; Hofer et al. Reference Hofer, Wagreich and Neuhuber2013; Verma & Armstrong-Altrin, Reference Verma and Armstrong-Altrin2016; Armstrong-Altrin et al. Reference Armstrong-Altrin, Botello, Villanueva and Soto2019). However, temporal provenance shifts often remain untraceable because of their effects being masked by simultaneous sediment supply from multiple sources. This can be resolved only when mineralogical and geochemical variations in a sedimentary formation are correlated not only to probable sources but also to time. Due to their limited mobility during weathering and transportation, many trace elements, including rare earth elements (REEs), present in the clastic sedimentary rocks are useful in tracking dispersal of sediments (McLennan et al. Reference McLennan, Taylor and Eriksson1983; Bhatia & Crook, Reference Bhatia and Crook1986; Armstrong-Altrin & Verma, Reference Armstrong-Altrin and Verma2005; Armstrong-Altrin et al. Reference Armstrong-Altrin, Nagarajan, Madhavaraju, Rosalez-Hoz, Lee, Balaram, Cruz-Martínez and Avila-Ramírez2013, Reference Armstrong-Altrin, Ramos-Vázquez, Zavala-León and Montiel-García2018; George & Ray, Reference George and Ray2017; Ramos-Vázquez & Armstrong-Altrin, Reference Ramos-Vázquez and Armstrong-Altrin2019). The concentrations of trace elements and Sr and Nd isotopic composition provide information about the chemical composition and age of mantle derivation (or duration of crustal residence) of the source rocks. The integrated approach of using trace-element content, Nd isotope model ages and Sr isotope ratios is therefore likely to resolve secular shifts in the sedimentary source. The result may further enlighten us about the related plate tectonic setting.

The c. 3000 m thick Mesozoic succession in the Kutch Basin, western India is taken as a test case here. The choice is prompted by the fact that the secular shift in sediment contributing source has already been suggested for this succession based on petrographic modal analysis and major-element composition (Chaudhuri et al. 2020). However, the present approach moves forwards to the compositional make-up of the source rocks of specific time ranges. The main objective of this work is to determine the lithology and age of the source rock of Mesozoic sediments in Kutch. For this purpose, we use geochemical (trace elements) and isotopic (Sr–Nd) fingerprinting.

2. Geological setting

During the Late Jurassic Period, after the break-up of eastern Gondwana from its western counterpart, the earlier-formed Kutch Basin remained attached to the Indian subcontinent. Madagascar remained with western India until 88–90 Ma (Storey et al. Reference Storey, Mahoney, Saunders, Duncan, Kelley and Coffin1995; Torsvik et al. Reference Torsvik, Tucker, Ashwald, Eidea, Rakotosolofod and De Wite1998) while Seychelles split at c. 63 Ma (Collier et al. Reference Collier, Sansom, Ishizuka, Taylor, Minshull and Whitmarsh2008). The Kutch Basin is a proven petroliferous basin at the western continental margin of India formed by the reactivation of primordial faults along the Aravalli–Delhi fold belt during the Gondwanaland break-up (Biswas, Reference Biswas1982, Reference Biswas1987, Reference Biswas2005). During the syn-rift to early post-rift stage, the Kutch Basin accumulated Upper Triassic – Lower Cretaceous mixed siliciclastic and carbonate sediments of thickness c. 3000 m (Biswas, Reference Biswas1982, Reference Biswas1987). The post-rift Cenozoic succession consists of mixed carbonate siliciclastic intervals (Biswas, Reference Biswas1981; Banerjee et al. Reference Banerjee, Chattoraj, Saraswati, Dasgupta and Sarkar2012 a, b; Saraswati et al. Reference Saraswati, Khanolkar and Banerjee2018). The Kutch Basin remained surrounded by Nagar Parkar Igneous Suite to the north, Bhilwara, Aravalli and Delhi Supergroups to the NE, Dharwar Supergroup to the south and Central Indian Tectonic Zone (CITZ) to the east. The Aravalli highlands to the east and the Nagar Parkar Ridge to the north are traditionally considered as source areas of the Mesozoic sediments in the Kutch Basin (Ahmad & Bhat, Reference Ahmad and Bhat2006; Ramakrishnan & Vaidyanadhan, Reference Ramakrishnan and Vaidyanadhan2008; Ahmad et al. Reference Ahmad, Noufal, Masroor and Khan2014; Valdiya, Reference Valdiya2015). The E–W-trending faults result in a series of uplifts exposing the Mesozoic sections in Island Belt Uplift (comprising Patcham, Khadir, Bela and Chorad islands), Wagad Uplift and Kutch Mainland Uplift (Biswas, Reference Biswas1980, Reference Biswas2005) (Fig. 1).

Fig. 1. Geological map of Kutch mainland showing extents of Cenozoic and Mesozoic outcrops (after Biswas, Reference Biswas1977, Reference Biswas1981). Mesozoic succession of the Kutch Basin, western India (after Biswas, Reference Biswas1977, Reference Biswas1981).

The Mesozoic stratigraphy of the Kutch Mainland Uplift comprises Jhurio, Jhumara, Jhuran and Bhuj formations in ascending order of succession overlying the basement rocks, separated by a basal granite–cobble conglomerate (Biswas & Deshpande, Reference Biswas and Deshpande1968; Biswas, Reference Biswas1987) (Fig. 1). The marine Jhurio Formation (Bathonian–Callovian) rests unconformably on the Precambrian basement and mainly comprises limestone and minor shale. The overlying Jhumara Formation (Oxfordian) consists of argillaceous sediments at the base, and limestone and minor sandstone beds at the top (Biswas, Reference Biswas2005). The shallow-marine Jhuran Formation (Kimmeridgian–Tithonian), unconformably overlying the Jhumara Formation, consists primarily of sandstone–shale alternations and it is divided into three constituent members based on three coarsening-upwards cycles of depositions (Arora et al. Reference Arora, Banerjee and Dutta2015, Reference Arora, Dutta, Gogoi and Banerjee2017). The youngest Bhuj Formation (Valanginian–Albian), consisting exclusively of sandstone and shale, was deposited in fluvio-deltaic conditions (Biswas, Reference Biswas, Tandon, Pant and Casshyap1991). Several workers reported a predominantly southwesterly palaeocurrent pattern for the entire Mesozoic sequence (Biswas, Reference Biswas, Tandon, Pant and Casshyap1991, Reference Biswas1993, Reference Biswas2005; Mandal et al. Reference Mandal, Koner, Sarkar, Tawfik, Chakraborty, Bhakta and Bose2016; Arora et al. Reference Arora, Dutta, Gogoi and Banerjee2017; Desai & Biswas, Reference Desai and Biswas2018).

3. Methods

Samples for the present study were collected from different locations in the Kutch Mainland Uplift, namely Zara, Nirona, Palara, Bhuj, Yaksh, Rukmavati, Gangeshwar and Tapkeshwar. The majority of the collected samples from the oldest Jhurio Formation contain a significantly high amount of calcareous fragments and were therefore not included in the current study. Besides, pore-filling and replacive carbonate cements are abundant within sandstones in Jhumara, Jhuran and Bhuj formations, whereas they are absent from the associated shales in these formations (Chaudhuri et al. Reference Chaudhuri, Banerjee and Le Pera2018). We have therefore mainly selected shale samples from Jhumara, Jhuran and Bhuj formations along with two sandstone samples from the Bhuj Formation because of the negligible amounts of carbonate cementation in them. In total, 15 samples from Jhumara Formation, 53 samples from Jhuran Formation and 17 samples from Bhuj Formation were analysed for the present investigation. The Jhuran Formation is considerably thick and is mapped by previous workers up to the Member level in the Kutch Mainland. A large number of studied samples therefore belong to the Jhuran Formation. For trace-element and REE concentrations, c. 50 mg of each sample powder was used to prepare sample solutions with 5 mL of 1 ng mL–1 103Rh solution added as an internal standard. From this stock solution, 5 mL was diluted to 50 mL using deionized water and analysed at CSIR-NGRI, Hyderabad, using a high-resolution inductively coupled plasma mass spectrometer (HR-ICP-MS) (Nu Instruments Attom, UK). Chinese standard reference material GSR-5 was used to check the accuracy and precision of measurements. Some of the trace-element data (sample series JP, NP and BP) were generated at the Physical Research Laboratory, Ahmedabad, India using a Thermo X-Series Q-ICP-MS. For this, carbonate and organic matter were removed from powdered samples before being dissolved in HF-HNO3. Analyses were carried out in 2% HNO3 sample solutions and BHVO-2 standard (from USGS) was used for calibration as well as for checking the accuracy and precision of measurements. Reproducibility at 2σ level was ≤3% for REEs and ≤6% for all other trace elements (Chatterjee & Ray, Reference Chatterjee and Ray2017, Reference Chatterjee and Ray2018).

For Sr–Nd isotopic ratio analyses at the Physical Research Laboratory, Ahmedabad, India, the samples were dissolved following the standard HF-HNO3-HCl dissolution protocol for silicate rocks (e.g. Awasthi et al. Reference Awasthi, Ray, Laskar, Kumar, Sudhakar, Bhutani, Sheth and Yadava2010). Sr was separated from other REEs by cation exchange column chromatography while Nd was separated from other REEs using Ln-specific resin from Eichrom with dilute HCl (0.18 N) as elutant (Dickin, Reference Dickin2000; Awasthi et al. Reference Awasthi, Ray, Laskar, Kumar, Sudhakar, Bhutani, Sheth and Yadava2010). 87Sr/86Sr and 143Nd/144Nd were measured using a Thermo Neptune multi-collector (MC) ICP-MS (Awasthi et al. Reference Awasthi, Ray, Singh, Band and Rai2014). The measured 87Sr/86Sr and 143Nd/144Nd ratios were corrected for mass fractionation using 86Sr/88Sr of 0.1194 and 146Nd/144Nd of 0.7219. Average measured 143Nd/144Nd for an internal laboratory standard (Merck Nd) was 0.511705 ± 27 (2σ, n = 56). The average 143Nd/144Nd for BHVO-2, measured regularly during our analyses, is 0.512967 ± 0.000008 (n = 10; ±0.2 in εNd units at 2σ). Further details of analytical protocols and data presentation for comparison are provided in Chatterjee & Ray (Reference Chatterjee and Ray2017, Reference Chatterjee and Ray2018).

4. Results

4.a. Trace-element composition

Trace-element concentrations of samples from Jhumara, Jhuran and Bhuj formations are presented in Tables 15. Concentrations of most trace elements, especially that of large-ion lithophile elements (LILEs), are higher in these samples compared with the average shale composition (Wedepohl, Reference Wedepohl1971), except for Ni and Sc. The Lower Member of the Jhuran Formation possesses the highest average ∑REE of c. 317 ppm. The Upper Continental Crust (UCC) normalized trace-element concentrations in these samples are either enriched, depleted or UCC-equivalent (Fig. 2). Ba, U, Pb and Sr exhibit prominent depletion with increasing Sr depletion in younger formations. Th shows enrichment for the entire Mesozoic record. However, Nb and Ta show enrichment. The Lower Jhuran sediments show an increase in Zr and Hf, whose magnitude diminishes towards the top of the succession (Fig. 2). The Nb/Ta versus Zr/Hf of all the samples from the basin show two clusters, one group showing higher Nb/Ta and lower Zr/Hf compared with the other (Fig. 3).

Table 1. Trace-element concentrations of samples from Jhumara Formation. CN – chondrite-normalized

Table 2. Trace-element concentrations of samples from Lower Member, Jhuran Formation. CN – chondrite-normalized

Table 3. Trace-element concentrations of samples from Middle Member, Jhuran Formation. CN – chondrite-normalized

Table 4. Trace-element concentrations of samples from Upper Member, Jhuran Formation. CN – chondrite-normalized

Table 5. Trace-element concentrations of samples from Bhuj Formation. CN – chondrite-normalized

Fig. 2. Upper Continental Crust (UCC-) normalized trace-element patterns for samples from various formations, arranged in stratigraphic order, of the Mesozoic Kutch Basin. Normalizing values are from Rudnick & Gao (Reference Rudnick, Gao, Holland and Turekian2003).

Fig. 3. Nb/Ta versus Zr/Hf plot for samples from the Mesozoic Kutch Basin.

Chondrite-normalized REE patterns of Jhumara, Jhuran and Bhuj samples show broadly similar characteristics, with a prominently LREE-enriched pattern, negative Eu anomaly and positive Gd anomaly (Fig. 4). These patterns broadly overlap with those of UCC and Post-Archean Australian Shale (PAAS). Weak negative Eu anomalies, Eu/Eu* = Eu/(SmCN*NdCN)1/2 (where the subscript ‘CN’ indicates chondrite-normalized) in Jhumara, Jhuran and Bhuj samples have average values of 0.65, 0.65 and 0.61, respectively. The chondrite-normalized REE patterns for shale samples from the Jhuran Formation reveal three distinct bands for its three Lower, Middle and Upper members (Fig. 4). The samples from the Lower Member exhibit consistently high ∑REE contents, whereas those of the Upper Member show distinctly low ∑REE content. Nearly half of the samples from the Jhumara Formation exhibit REE patterns similar to those of the Upper Member of the Jhuran Formation. Samples from all three formations exhibit a low La/Th ratio (1.59–3.01 ppm); those from the Bhuj Formation exhibit the highest average (2.60). In a cross-plot of Th versus Sc, the samples from the Jhuran Formation exhibit the maximum spread of data points (Fig. 5). Most samples from Jhuran and Bhuj formations and nearly half of the samples from the Jhumara Formation bear felsic signature, while a few samples from the Jhuran (belonging to the Upper and Middle members) and Bhuj formations and most samples from the Jhumara Formation plot within the field for intermediate igneous rocks (Fig. 5) (Cullers, Reference Cullers2002).

Fig. 4. Composite log of the Jhumara, Jhuran and Bhuj formations (adapted from Biswas, Reference Biswas2005; Fursich et al. 2005; Mandal et al. Reference Mandal, Koner, Sarkar, Tawfik, Chakraborty, Bhakta and Bose2016). Chondrite-normalized REE patterns for Jhumara (lower), Jhuran (middle) and Bhuj (upper) are placed to the right of the log. Solid lines: LM – Lower Member; MM – Middle Member; UM – Upper Member. Dashed lines: MORB – mid-ocean-ridge basalt; NASC – North American Shale Composite; PAAS – Post Archean Australian Shale. Chondrite meteorite values from Boynton (1984). NASC, PAAS and MORB values from Taylor & McLennan (Reference Taylor and McLennan1985).

Fig. 5. Source-rock discrimination based on Th versus Sc plot (Cullers, Reference Cullers2002) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Average Sc/Cr ratios for samples from Jhumara, Jhuran and Bhuj formations are 0.13, 0.12 and 0.13 respectively. In a cross-plot of La/Th versus Hf, most samples plot in the field marked for passive margin sources, but a few samples occupy the field of acidic arc sources (Fig. 6). Samples from the Lower Member of Jhuran Formation and the Bhuj Formation exhibit distinctly higher concentrations of Hf compared with samples from other formations.

Fig. 6. Tectonic setting discrimination based on La/Th versus Hf (Floyd & Leveridge, Reference Floyd and Leveridge1987) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

4.b. Isotope composition

Samples from Jhumara, Jhuran and Bhuj formations bear continental crustal signatures. The value of (87Sr/86Sr)t is similar in all three formations; samples from the Bhuj Formation have the highest (87Sr/86Sr)t ranging over 0.73–0.76, followed by Jhuran Formation (0.72–0.75) and Jhumara Formation (0.72–0.73) (Fig. 7). All samples exhibit overlapping but negative εNd(0) and εNd(t) values, suggesting derivation from LREE-enriched mantle. Samples from the Jhumara Formation and Lower and Middle members of the Jhuran Formation exhibit high values of εNd(t), while those from the Upper Member of the Jhuran Formation and the Bhuj Formation demonstrate relatively low εNd(t) values (Fig. 7a). In a cross-plot of εNd(0) versus Th/Sc, samples from Jhumara, Jhuran and Bhuj formation indicate the dominance of felsic components with values similar to those of upper crust (Fig. 8). The fractionation factor f Sm/Nd values vary from −0.37 to −0.49, −0.31 to −0.50 and −0.40 to −0.46 for the Jhumara, Jhuran and Bhuj samples, respectively. Figure 9 shows temporal variations in εNd(0), depleted mantle model age (T DM) and εNd(t) in the succession. The marked decrease of εNd(0) and εNd(t) in samples from the Bhuj Formation is accompanied by an increase in T DM. The T DM age calculated from the measured data range over 1560–2170 Ma, 1430–2220 Ma and 1760–2320 Ma for the Jhumara, Jhuran and Bhuj formations, respectively (Table 6). The average T DM for the analysed samples increases through 1570, 1680 and 1800 Ma from Jhumara through Jhuran and Bhuj formations with T DM mode at 1800 Ma (Fig. 10). In a cross-plot of f Sm/Nd versus εNd(0), Jhumara, Jhuran and Bhuj samples plot within the early Proterozoic upper crustal rocks (> 1.6 Ga) and are in agreement with the values of T DM (Fig. 11).

Fig. 7. (a) εNd(t) and (b) T DM versus (87Sr/86Sr)t plot for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Fig. 8. Source-rock discrimination based on εNd versus Th/Sc plot (McLennan et al. Reference McLennan, Hemming, McDaniel, Hanson, Johnson and Basu1993) for shale samples from Jhumara, Jhuran and Bhuj Formation. Abbreviations as for Figure 4.

Fig. 9. Variations in εNd(0), T DM(Ga) and εNd(t) across Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Table 6. Sm–Nd isotope data for samples from Jhumara, Jhuran and Bhuj formations

Note: Subscripts m and t indicate the ratios measured and at the time of deposition, respectively. Initial ratios were calculated at the age of deposition, given in brackets. Sm and Nd concentrations were measured by Q-ICP-MS at Physical Research Laboratory, Ahmedabad, India. Depleted mid-ocean-ridge basalt mantle values used for T DM calculations: 143Nd/144Nd = 0.513114 and 147Sm/144Nd = 0.222 (Michard et al. Reference Michard, Gurriet, Soudant and Albarede1985).

Fig. 10. Histogram of T DM distribution for the sediments in the Mesozoic Kutch Basin.

Fig. 11. Plot of f Sm/Nd versus εNd(0) (modified after McLennan et al. Reference McLennan, Hemming, McDaniel, Hanson, Johnson and Basu1993) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

5. Discussion

Both trace-element analysis and Sr–Nd isotopic composition of sediments in the Mesozoic Kutch Basin point to their derivation primarily from upper continental sources. This is reflected in the enriched LILE and LREE, (87Sr/86Sr)t > 0.71, εNd(t) < −10 and negative f Sm/Nd (Condie, Reference Condie1993). The same is reflected in the T DM distribution of the sediments (Fig. 12). T DM mode at 1800 Ma is related to one of the major crust-forming episodes on the Earth (Condie, Reference Condie2001; Condie & Aster, Reference Condie and Aster2010). The insignificant change in εNd(t) from the Jhumara Formation to the Middle Member of the Jhuran Formation suggests a negligible shift in sediment sources. The low Th (Fig. 5), lowest (87Sr/86Sr)t and average T DM age of 1570 Ma indicate sediment inputs from a juvenile mafic source for the oldest Jhumara Formation. The basin possibly received sediments from young orogenic belt(s) during the initial phase. Importantly, the influence of such a sediment source diminishes through time, with felsic (granitic) source progressively dominating from the bottom to the top of the Mesozoic succession. The concentration of trace elements such as Th and Sc, the LREE-enriched chondrite normalized patterns and the negative Eu anomaly suggest the dominance of felsic composition at source for Jhumara, Jhuran and Bhuj formations. The high content of Th (average, 26.4 ppm) and high ∑REE in samples from the Lower Member of the Jhuran Formation are possibly related to the high content of monazite in the samples as reported by Chaudhuri et al. (Reference Chaudhuri, Banerjee and Le Pera2018) (Table 2). Low Sc/Cr ratios and the cross-plot of La/Th versus Hf indicate a passive-margin setting (Bhatia & Crook, Reference Bhatia and Crook1986; Floyd & Leveridge, Reference Floyd and Leveridge1987) (Fig. 6).

Fig. 12. Average T DM versus average εNd(t) plot for Mesozoic sediments in the Kutch Basin compared with those of potential sources in Rajasthan. Error bars for the source points are 3σ. Data sources: George & Ray (Reference George and Ray2017) and Shukla et al. (2019).

The depletion of Ba, U, Pb and Sr observed in all the formations (Fig. 2) are possibly linked to the weathering of the source region (cf. Price et al. Reference Price, Gray, Wilson, Frey and Taylor1991). The magnitude of Sr depletion increases from the bottom to the top of the succession, which could be attributed to higher weathering in the source region as a function of time (i.e. the younger formations received more sediments from highly weathered sources). The highly weathered nature of the source possibly corresponds to the progressive tectonic stability of the basin from syn-rift to post-rift. The two clusters in Nb/Ta versus Zr/Hf cross-plot suggest multiple weathering or depositional histories (Fig. 3). The high Nb/Ta (low Zr/Hf) most likely represents highly recycled sediments, that is, derivation from sedimentary source rocks; the low Nb/Ta could be attributed to derivation from felsic igneous rocks.

The Nd isotopic composition of the Bhuj Formation indicates predominant sediment derivation from crustal sources of early Proterozoic age (Figs 7, 8, 1012). T DM ages indicate the dominance of source rocks of late Palaeoproterozoic age (1700–1800 Ma), with subordinate contributions from rocks of early Mesoproterozoic (1400–1600 Ma) and early Palaeoproterozoic (2100–2400 Ma) ages (Fig. 10). The mode of T DM is c. 1800 ± 200 Ma.

Considering the overwhelming southwesterly slope of the basin (Biswas, Reference Biswas, Tandon, Pant and Casshyap1991, Reference Biswas2005; Arora et al. Reference Arora, Banerjee and Dutta2015, Reference Arora, Dutta, Gogoi and Banerjee2017; Mandal et al. Reference Mandal, Koner, Sarkar, Tawfik, Chakraborty, Bhakta and Bose2016; Desai & Biswas, Reference Desai and Biswas2018), the possible rocks which could have sourced these sediments include Nagar Parkar Igneous Suite exposed in the Nagar Parkar Ridge; Banded Gneissic Complex; Aravalli, Bhilwara, Delhi and Marwar supergroups; Erinpura Granite; and Malani volcanic rocks exposed in the Aravalli highlands (Fig. 13). The Nd isotopic composition of Mesozoic sediments in the Kutch Basin and their T DM ages overlap with those of the Proterozoic Marwar Supergroup and the Erinpura Granites (Fig. 12), suggesting these rocks were the major source of sediments to the basin. The mafic juvenile component, especially in the Jhumara Formation, possibly derives from the Nagar Parker Igneous Suite in the north, which contains mafic magmatic rocks within a largely granitic terrain (e.g. Khan et al. Reference Khan, Murata, Rehman, Zafar and Ozawa2012, Reference Khan, Murata, Jan, Rehman, Zafar, Ozawa, Qadir and Mehmood2017; Jan et al. Reference Jan, Agheem, Laghari and Anjum2017; de Wall et al. Reference de Wall, Pandit, Donhauser, Schöbel, Wang and Sharma2018; Rehman et al. Reference Rehman, Khan, Jan, Lee, Chung and Murata2018). This indicates the existence of southerly drainage during the initial phase of the basin. Subsequently, the palaeoslope became southwesterly because of the seawards tilting of the basin. Although there is substantial similarity in the source of the Middle Jurassic syn-rift sediments and the Lower Cretaceous post-rift sediments, trace-element concentrations, their ratios and Nd isotopic compositions reveal subtle variations in source characteristics of younger sediments. The increasing Hf concentration in samples from Jhuran and Bhuj formations exhibiting older sediment components suggest erosional unroofing at the source (Fig. 6). Increasing mean T DM ages exhibited by Jhumara, Jhuran and Bhuj samples further supports this interpretation.

Fig. 13. Schematic diagram of palaeodrainage pattern in the Early Cretaceous Kutch Basin and possible provenance areas.

This study indicates the dominance of felsic source rocks for the entire Mesozoic rock record. Integrating results from trace-element and Sr–Nd systematics (low Th, low (87Sr/86Sr)t and T DM < 1600 Ma) highlights additional juvenile mafic source rocks in the Jhumara Formation. The concentration of Hf indicates older source rocks in younger Jhuran and Bhuj formations. Finally, εNd(t) and T DM ages, along with the established palaeoslope, provide a more reliable correlation with the existing rocks at source area. This confirms the applicability of trace elements and Nd isotope geochemistry in tracing multiple source rocks.

6. Conclusions

Concentrations of trace elements (Th, Sc) and their ratios (Nb/Ta and Zr/Hf) indicate that multiple sources, including both igneous and sedimentary rocks, contribute sediments to Jhumara, Jhuran and Bhuj formations of the Mesozoic Kutch Basin. High (87Sr/86Sr)t, negative εNd and LREE enrichment indicate enriched crustal source of sediments. Although the felsic source rocks dominate, sediments in the Jhumara Formation exhibit inputs from a juvenile mafic source.

T DM ages indicate that rocks of late Palaeoproterozoic age (1700–1800 Ma) are the dominant sediment contributors, as well as those of early Mesoproterozoic (1400–1600 Ma) and early Palaeoproterozoic (2100–2400 Ma) ages.

Concentrations of Hf indicates the dominance of older inputs in younger sediments of Jhuran and Bhuj formations. The increasing mean T DM from bottom to top, namely 1570, 1680 and 1800 Ma for Jhumara, Jhuran and Bhuj formations, respectively, corroborates sediments from older rocks in younger formations, suggesting erosional unroofing of source rocks. Average T DM of 1800 Ma coincides with one of the major crust-forming events.

The increasing (87Sr/86Sr)t from the bottom to the top of the Mesozoic succession relates to increased weathering of the source rock, possibly linked to the increasing tectonic stability of the basin.

Considering the southwesterly palaeoslope of this basin and the mean T DM ages, rocks of the Marwar Supergroup and Erinpura Granite north and NE of the Kutch Basin constitute the main sediment source. However, minor amount of sediments are derived from the Palaeoproterozoic Banded Gneissic Complex II, Bhilwara Supergroup and Aravalli Supergroup and the Mesoproterozoic Delhi Supergroup. The mafic juvenile component in the sediments of Jhumara Formation is likely to have derived from the Nagar Parkar Igneous Suite to the north.

Acknowledgement

We thank our host institutes for infrastructure support. SB acknowledges the Science and Engineering Research Board, Government of India for financial support through grant no. SR/S4/ES-709/2014 of the Government of India. We also thank Gaurav Chauhan for assistance with sample collection and Emilia Le Pera for constructive suggestions.

References

Ahmad, AHM and Bhat, GM (2006) Petrofacies, provenance and diagenesis of the Dhosa sandstone member (Chari Formation) at Ler, Kachchh sub-basin, Western India. Journal of Asian Earth Sciences 27, 857–72, https://doi.org/10.1016/j.jseaes.2005.08.005.CrossRefGoogle Scholar
Ahmad, AHM, Noufal, KN, Masroor, AM and Khan, T (2014) Petrography and geochemistry of Jumara Dome sediments, Kachchh Basin: Implications for provenance, tectonic setting and weathering intensity. Chinese Journal of Geochemistry 33, 923, https://doi.org/10.1007/s11631-014-0656-4.CrossRefGoogle Scholar
Algeo, TJ and Maynard, JB (2004) Trace-element behavior and redox facies in core shales of Upper Pennsylvanian Kansas-type cyclothems. Chemical Geology 206, 289318, https://doi.org/10.1016/j.chemgeo.2003.12.009.CrossRefGoogle Scholar
Armstrong-Altrin, JS, Botello, AV, Villanueva, SF and Soto, LA (2019) Geochemistry of surface sediments from the northwestern Gulf of Mexico: implications for provenance and heavy metal contamination. Geological Quarterly 63, 522–38.CrossRefGoogle Scholar
Armstrong-Altrin, JS, Nagarajan, R, Madhavaraju, J, Rosalez-Hoz, L, Lee, YI, Balaram, V, Cruz-Martínez, A and Avila-Ramírez, G (2013) Geochemistry of the Jurassic and Upper Cretaceous shales from the Molango Region, Hidalgo, eastern Mexico: Implications for source-area weathering, provenance, and tectonic setting. Comptes Rendus Geoscience 345, 185202, https://doi.org/10.1016/j.crte.2013.03.004.CrossRefGoogle Scholar
Armstrong-Altrin, JS, Ramos-Vázquez, MA, Zavala-León, AC and Montiel-García, PC (2018) Provenance discrimination between Atasta and Alvarado beach sands, western Gulf of Mexico, Mexico: Constraints from detrital zircon chemistry and U–Pb geochronology. Geological Journal 53, 2824–48, https://doi.org/10.1002/gj.3122.CrossRefGoogle Scholar
Armstrong-Altrin, JS and Verma, SP (2005) Critical evaluation of six tectonic setting discrimination diagrams using geochemical data of Neogene sediments from known tectonic settings. Sedimentary Geology 177, 115–29, https://doi.org/10.1016/j.sedgeo.2005.02.004.CrossRefGoogle Scholar
Arora, A, Banerjee, S and Dutta, S (2015) Black shale in late Jurassic Jhuran Formation of Kutch: Possible indicator of oceanic anoxic event? Journal of the Geological Society of India 85, 265–78, https://doi.org/10.1007/s12594-015-0215-6.CrossRefGoogle Scholar
Arora, A, Dutta, S, Gogoi, B and Banerjee, S (2017) The effects of igneous dike intrusion on organic geochemistry of black shale and its implications: Late Jurassic Jhuran Formation, India. International Journal of Coal Geology 178, 8499, doi: 10.1016/j.coal.2017.05.002.CrossRefGoogle Scholar
Awasthi, N, Ray, JS, Laskar, AH, Kumar, A, Sudhakar, M, Bhutani, R, Sheth, HC and Yadava, MG (2010) Major ash eruptions of Barren Island volcano (Andaman Sea) during the past 72 kyr: clues from a sediment core record. Bulletin of Volcanology 72, 1131–36, https://doi.org/10.1007/s00445-010-0408-1.CrossRefGoogle Scholar
Awasthi, N, Ray, JS, Singh, AK, Band, ST and Rai, VK (2014) Provenance of the Late Quaternary sediments in the Andaman Sea: Implications for monsoon variability and ocean circulation. Geochemistry, Geophysics, Geosystems 15, 3890–906, https://doi.org/10.1002/2014GC005462.CrossRefGoogle Scholar
Banerjee, S, Chattoraj, SL, Saraswati, PK, Dasgupta, S and Sarkar, U (2012a) Mineralogy and geochemistry of lagoonal glauconites and their implications on origin and maturation: Oligocene Maniyara Fort Formation, western Kutch, India. Geological Journal 47, 357 –71.CrossRefGoogle Scholar
Banerjee, S, Chattoraj, SL, Saraswati, PK, Dasgupta, S and Sarkar, U (2012b) Substrate control on formation and maturation of glauconites in the Middle Eocene Harudi Formation, western Kutch, India. Marine and Petroleum Geology 30, 144–60, https://doi.org/10.1016/j.marpetgeo.2011.10.008.CrossRefGoogle Scholar
Bhatia, MR and Crook, KA (1986) Trace element characteristics of graywackes and tectonic setting discrimination of sedimentary basins. Contributions to Mineralogy and Petrology 92, 181–93.CrossRefGoogle Scholar
Biswas, SK (1977) Mesozoic rock stratigraphy of Kachchh. Quarterly Journal of the Geology, Mining and Metallurgy Society of India 49, 152.Google Scholar
Biswas, SK (1980) Mesozoic rock-stratigraphy of Kutch, Gujarat. Quarterly Journal of Geology, Mining and Metallurgy Society of India 49, 151.Google Scholar
Biswas, SK (1981) Basin framework, palaeo-environment and depositional history of the Mesozoic sediments of Kutch basin, western India. Quarterly Journal of Geology, Mining and Metallurgy Society of India 53, 5685.Google Scholar
Biswas, SK (1982) Rift basins in the western margin of India and their hydrocarbon prospects with special reference to Kutch basin. AAPG Bulletin 66, 1497–513.Google Scholar
Biswas, SK (1987) Regional tectonic framework, structure and evolution of the western marginal basins of India. Tectonophysics 135, 307–27.CrossRefGoogle Scholar
Biswas, SK (1991) Stratigraphy and sedimentary evolution of the Mesozoic basin of Kutch, western India. In Stratigraphy and Sedimentary Evolution of Western India (eds Tandon, SK, Pant, CC, Casshyap, SM), pp. 74103. Nainital: Gyanodaya Prakashan.Google Scholar
Biswas, SK (1993) Geology of Kutch, Vol. 1. Dehradun: KDM Institute of Petroleum Exploration, 450 p.Google Scholar
Biswas, SK (2005) A review of structure and tectonics of Kutch basin, western India, with special reference to earthquakes. Current Science 88, 1592–600.Google Scholar
Biswas, SK and Deshpande, (1968) The basement of the Mesozoic sediments of Kutch, Western India. Bulletin of Geological, Mineral and Metallurgical Society India 40, 17.Google Scholar
Boynton, WV (1984) Cosmochemistry of the rare earth elements: meteorite studies. In Rare Earth Element Geochemistry (ed. Henderson, P), pp. 63114. Amsterdam: Elsevier, https://doi.org/10.1016/B978-0-444-42148-7.50008-3.CrossRefGoogle Scholar
Chatterjee, A and Ray, JS (2017) Sources and depositional pathways of mid-Holocene sediments in the Great Rann of Kachchh, India: Implications for fluvial scenario during the Harappan Culture. Quaternary International 443, 177–87, https://doi.org/10.1016/j.quaint.2017.06.008.CrossRefGoogle Scholar
Chatterjee, A and Ray, JS (2018) Geochemistry of Harappan potteries from Kalibangan and sediments in the Ghaggar River: Clues for a dying river. Geoscience Frontiers 9, 1203–11, https://doi.org/10.1016/j.gsf.2017.07.006.CrossRefGoogle Scholar
Chaudhuri, A, Banerjee, S and Chauhan, G (2020) Compositional evolution of siliciclastic sediments recording the tectonic stability of a pericratonic rift: Mesozoic Kutch Basin, western India. Marine and Petroleum Geology 111, 476–95, https://doi.org/10.1016/j.marpetgeo.2019.08.026.CrossRefGoogle Scholar
Chaudhuri, A, Banerjee, S and Le Pera, E (2018) Petrography of Middle Jurasssic to Early Cretaceous sandstones in the Kutch Basin, western India: Implications on provenance and basin evolution. Journal of Palaeogeography 7, 214, https://doi.org/10.1186/s42501-018-0002-6.CrossRefGoogle Scholar
Collier, JS, Sansom, V, Ishizuka, O, Taylor, RN, Minshull, TA and Whitmarsh, RB (2008) Age of Seychelles–India break-up. Earth and Planetary Science Letters 272, 264–77, https://doi.org/10.1016/j.epsl.2008.04.045.CrossRefGoogle Scholar
Condie, KC (1993) Chemical composition and evolution of the upper continental crust: contrasting results from surface samples and shales. Chemical Geology 104, 137, https://doi.org/10.1016/0009-2541(93)90140-E.CrossRefGoogle Scholar
Condie, KC (2001) Continental growth during formation of Rodinia at 1.35-0.9 Ga. Gondwana Research 4, 516, https://doi.org/10.1016/S1342-937X(05)70650-X.CrossRefGoogle Scholar
Condie, KC and Aster, RC (2010) Episodic zircon age spectra of orogenic granitoids: the supercontinent connection and continental growth. Precambrian Research 180, 227–36, https://doi.org/10.1016/j.precamres.2010.03.008.CrossRefGoogle Scholar
Cullers, RL (2002) Implications of elemental concentrations for provenance, redox conditions, and metamorphic studies of shales and limestones near Pueblo, CO, USA. Chemical Geology 191, 305–27, https://doi.org/10.1016/S0009-2541(02)00133-X.CrossRefGoogle Scholar
de Wall, H, Pandit, MK, Donhauser, I, Schöbel, S, Wang, W and Sharma, KK (2018) Evolution and tectonic setting of the Malani–Nagarparkar igneous suite: a Neoproterozoic silicic-dominated large igneous province in NW India–SE Pakistan. Journal of Asian Earth Sciences 160, 136–58, https://doi.org/10.1016/j.jseaes.2018.04.016.CrossRefGoogle Scholar
Desai, BG and Biswas, SK (2018) Postrift deltaic sedimentation in western Kachchh Basin: insights from ichnology and sedimentology. Palaeogeography, Palaeoclimatology, Palaeoecology 504, 104–24, https://doi.org/10.1016/j.palaeo.2018.05.013.CrossRefGoogle Scholar
Dickin, AP (2000) Crustal formation in the Grenville Province: Nd-isotope evidence. Canadian Journal of Earth Sciences 37, 165–81, https://doi.org/10.1139/e99-039.CrossRefGoogle Scholar
Dickinson, WR (1988) Provenance and sediment dispersal in relation to paleotectonics and paleogeography of sedimentary basins. In New Perspectives in Basin Analysis (eds Kleinspehn, KL and Paola, C), pp. 325. New York, NY: Springer, https://doi.org/10.1007/978-1-4612-3788-4_1.CrossRefGoogle Scholar
Dickinson, WR and Suczek, CA (1979) Plate tectonics and sandstone compositions. AAPG Bulletin 63, 2164–82.Google Scholar
Floyd, PA and Leveridge, BE (1987) Tectonic environment of the Devonian Gramscatho basin, south Cornwall: framework mode and geochemical evidence from turbiditic sandstones. Journal of the Geological Society 144, 531–42, https://doi.org/10.1144/gsjgs.144.4.0531.CrossRefGoogle Scholar
Fürsich, FT, Singh, IB, Joachimski, M, Krumm, S, Schlirf, M and Schlirf, S (2005) Palaeoclimate reconstructions of the Middle Jurassic of Kachchh (western India): an integrated approach based on palaeoecological, oxygen isotopic, and clay mineralogical data. Palaeogeography, Palaeoclimatology, Palaeoecology 217, 289309, doi: 10.1016/j.palaeo.2004.11.026.CrossRefGoogle Scholar
George, BG and Ray, JS (2017) Provenance of sediments in the Marwar Supergroup, Rajasthan, India: Implications for basin evolution and Neoproterozoic global events. Journal of Asian Earth Sciences 147, 254–70, https://doi.org/10.1016/j.jseaes.2017.07.027.CrossRefGoogle Scholar
Hofer, G, Wagreich, M and Neuhuber, S (2013) Geochemistry of fine-grained sediments of the upper Cretaceous to Paleogene Gosau Group (Austria, Slovakia): Implications for paleoenvironmental and provenance studies. Geoscience Frontiers 4, 449–68, https://doi.org/10.1016/j.gsf.2012.11.009.CrossRefGoogle Scholar
Jan, MQ, Agheem, MH, Laghari, A and Anjum, S (2017) Geology and petrography of the Nagar Parkar igneous complex, southeastern Sindh, Pakistan: the Kharsar body. Journal of the Geological Society of India 89, 91–8, https://doi.org/10.1007/s12594-017-0564-4.CrossRefGoogle Scholar
Khan, T, Murata, M, Jan, MQ, Rehman, HU, Zafar, M, Ozawa, H, Qadir, A and Mehmood, S (2017) Felsic dykes in the Neoproterozoic Nagar Parkar Igneous Complex, SE Sindh, Pakistan: geochemistry and tectonic settings. Arabian Journal of Geosciences 10, 308, https://doi.org/10.1007/s12517-017-3077-y.CrossRefGoogle Scholar
Khan, T, Murata, M, Rehman, HU, Zafar, M and Ozawa, H (2012) Nagarparker granites showing Rodinia remnants in the southeastern part of Pakistan. Journal of Asian Earth Sciences 59, 3951, https://doi.org/10.1016/j.jseaes.2012.05.028.CrossRefGoogle Scholar
Mandal, A, Koner, A, Sarkar, S, Tawfik, HA, Chakraborty, N, Bhakta, S and Bose, PK (2016) Physico-chemical tuning of palaeogeographic shifts: Bhuj formation, Kutch, India. Marine and Petroleum Geology 78, 474–92, https://doi.org/10.1016/j.marpetgeo.2016.10.003.CrossRefGoogle Scholar
McLennan, SM, Hemming, S, McDaniel, DK and Hanson, GN (1993) Geochemical approaches to sedimentation, provenance, and tectonics. In Processes Controlling the Composition of Clastic Sediments (eds Johnson, MJ and Basu, A), pp. 2140. Geological Society of America, Special Paper no. 285.CrossRefGoogle Scholar
McLennan, SM, Taylor, SR and Eriksson, KA (1983) Geochemistry of Archean shales from the Pilbara Supergroup, western Australia. Geochimica et Cosmochimica Acta 47, 1211–22, https://doi.org/10.1016/0016-7037(83)90063-7.CrossRefGoogle Scholar
Michard, A, Gurriet, P, Soudant, M and Albarede, F (1985) Nd isotopes in French Phanerozoic shales: external vs. internal aspects of crustal evolution. Geochimica et Cosmochimica Acta 49, 601–10, doi: 10.1016/0016-7037(85)90051-1.CrossRefGoogle Scholar
Nesbitt, W II and Young, GM (1982) Early Proterozoic climates and plate. Nature 299, 715–17, https://doi.org/10.1038/299715a0.CrossRefGoogle Scholar
Price, RC, Gray, CM, Wilson, RE, Frey, FA and Taylor, SR (1991) The effects of weathering on rare-earth element, Y and Ba abundances in Tertiary basalts from southeastern Australia. Chemical Geology 93, 245–65, https://doi.org/10.1016/0009-2541(91)90117-A.CrossRefGoogle Scholar
Ramakrishnan, M and Vaidyanadhan, R (2008) Geology of India , pp. 261–333. Bangalore: Geological Society of India.Google Scholar
Ramos-Vázquez, MA and Armstrong-Altrin, JS (2019) Sediment chemistry and detrital zircon record in the Bosque and Paseo del Mar coastal areas from the southwestern Gulf of Mexico. Marine and Petroleum Geology 110, 650–75.CrossRefGoogle Scholar
Rehman, HU, Khan, T, Jan, MQ, Lee, HY, Chung, SL and Murata, M (2018) Timing and span of the continental crustal growth in SE Pakistan: evidence from LA-ICP-MS U–Pb zircon ages from granites of the Nagar Parkar Igneous Complex. Gondwana Research 61, 172–86, https://doi.org/10.1016/j.gr.2018.04.018.CrossRefGoogle Scholar
Rudnick, RL and Gao, S (2003) Composition of the continental crust. In Treatise of Geochemistry (eds Holland, HD and Turekian, KK), pp. 164. Amsterdam: Elsevier. https://doi.org/10.1016/B0-08-043751-6/03016-4.Google Scholar
Saraswati, PK, Khanolkar, S and Banerjee, S (2018) Paleogene stratigraphy of Kutch, India: an update about progress in foraminiferal biostratigraphy. Geodinamica Acta 30, 100–18, https://doi.org/10.1080/09853111.2017.1408263.CrossRefGoogle Scholar
Shukla, AD, George, BG and Ray, JS (2020) Evolution of the Proterozoic Vindhyan Basin, Rajasthan, India: insights from geochemical provenance of siliciclastic sediments. International Geology Review 62(2), 153–67, https://doi.org/10.1080/00206814.2019.1594412. CrossRefGoogle Scholar
Storey, M, Mahoney, JJ, Saunders, AD, Duncan, RA, Kelley, SP and Coffin, MF (1995) Timing of hot spot-related volcanism and the breakup of Madagascar and India. Science 267, 852–55, https://doi.org/10.1126/science.267.5199.852.CrossRefGoogle Scholar
Taylor, SR and McLennan, SM (1985) The Continental Crust: Its Composition and Evolution. Oxford: Blackwell Scientific Publications.Google Scholar
Torsvik, TH, Tucker, RD, Ashwald, LD, Eidea, EA, Rakotosolofod, NA and De Wite, MJ (1998) Late Cretaceous magmatism in Madagascar: palaeomagnetic evidence for a stationary Marion hotspot. Earth Planetary Science Letters 164, 221–32, https://doi.org/10.1016/S0012-821X(98)00206-4.CrossRefGoogle Scholar
Tribovillard, N, Algeo, TJ, Lyons, T and Riboulleau, A (2006) Trace metals as paleoredox and paleoproductivity proxies: an update. Chemical Geology 232, 1232, https://doi.org/10.1016/j.chemgeo.2006.02.012.CrossRefGoogle Scholar
Valdiya, KS (2015) The Making of India: Geodynamic Evolution. Heidelberg: Springer.Google Scholar
Verma, SP and Armstrong-Altrin, JS (2016) Geochemical discrimination of siliciclastic sediments from active and passive margin settings. Sedimentary Geology 332, 112, https://doi.org/10.1016/j.sedgeo.2015.11.011.CrossRefGoogle Scholar
Wedepohl, KH (1971) Environmental influences on the chemical composition of shales and clays. Physics and Chemistry of the Earth 8, 305–33, https://doi.org/10.1016/0079-1946(71)90020-6.CrossRefGoogle Scholar
Figure 0

Fig. 1. Geological map of Kutch mainland showing extents of Cenozoic and Mesozoic outcrops (after Biswas, 1977, 1981). Mesozoic succession of the Kutch Basin, western India (after Biswas, 1977, 1981).

Figure 1

Table 1. Trace-element concentrations of samples from Jhumara Formation. CN – chondrite-normalized

Figure 2

Table 2. Trace-element concentrations of samples from Lower Member, Jhuran Formation. CN – chondrite-normalized

Figure 3

Table 3. Trace-element concentrations of samples from Middle Member, Jhuran Formation. CN – chondrite-normalized

Figure 4

Table 4. Trace-element concentrations of samples from Upper Member, Jhuran Formation. CN – chondrite-normalized

Figure 5

Table 5. Trace-element concentrations of samples from Bhuj Formation. CN – chondrite-normalized

Figure 6

Fig. 2. Upper Continental Crust (UCC-) normalized trace-element patterns for samples from various formations, arranged in stratigraphic order, of the Mesozoic Kutch Basin. Normalizing values are from Rudnick & Gao (2003).

Figure 7

Fig. 3. Nb/Ta versus Zr/Hf plot for samples from the Mesozoic Kutch Basin.

Figure 8

Fig. 4. Composite log of the Jhumara, Jhuran and Bhuj formations (adapted from Biswas, 2005; Fursich et al. 2005; Mandal et al.2016). Chondrite-normalized REE patterns for Jhumara (lower), Jhuran (middle) and Bhuj (upper) are placed to the right of the log. Solid lines: LM – Lower Member; MM – Middle Member; UM – Upper Member. Dashed lines: MORB – mid-ocean-ridge basalt; NASC – North American Shale Composite; PAAS – Post Archean Australian Shale. Chondrite meteorite values from Boynton (1984). NASC, PAAS and MORB values from Taylor & McLennan (1985).

Figure 9

Fig. 5. Source-rock discrimination based on Th versus Sc plot (Cullers, 2002) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Figure 10

Fig. 6. Tectonic setting discrimination based on La/Th versus Hf (Floyd & Leveridge, 1987) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Figure 11

Fig. 7. (a) εNd(t) and (b) TDM versus (87Sr/86Sr)t plot for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Figure 12

Fig. 8. Source-rock discrimination based on εNd versus Th/Sc plot (McLennan et al.1993) for shale samples from Jhumara, Jhuran and Bhuj Formation. Abbreviations as for Figure 4.

Figure 13

Fig. 9. Variations in εNd(0), TDM(Ga) and εNd(t) across Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Figure 14

Table 6. Sm–Nd isotope data for samples from Jhumara, Jhuran and Bhuj formations

Figure 15

Fig. 10. Histogram of TDM distribution for the sediments in the Mesozoic Kutch Basin.

Figure 16

Fig. 11. Plot of fSm/Nd versus εNd(0) (modified after McLennan et al.1993) for shale samples from Jhumara, Jhuran and Bhuj formations. Abbreviations as for Figure 4.

Figure 17

Fig. 12. Average TDM versus average εNd(t) plot for Mesozoic sediments in the Kutch Basin compared with those of potential sources in Rajasthan. Error bars for the source points are 3σ. Data sources: George & Ray (2017) and Shukla et al. (2019).

Figure 18

Fig. 13. Schematic diagram of palaeodrainage pattern in the Early Cretaceous Kutch Basin and possible provenance areas.