Next Article in Journal
A Multidisciplinary Approach for the Characterization of Artificial Cavities of Historical and Cultural Interest: The Case Study of the Cloister of Sant’Agostino—Caserta, Italy
Previous Article in Journal
Impact of Land Use and Land Cover Change on Hydrological Processes in Urban Watersheds: Analysis and Forecasting for Flood Risk Management
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Provenance of Detrital Rutiles from the Triassic–Jurassic Sandstones in Franz Josef Land (Barents Sea Region, Russian High Arctic): U-Pb Ages and Trace Element Geochemistry

by
Victoria Ershova
1,2,3,*,
Andrei Prokopiev
3 and
Daniel Stockli
4
1
Institute of Earth Sciences, Saint Petersburg State University, Universitetskaya nab. 7/9, St. Petersburg 199034, Russia
2
Geological Institute of Russian Academy of Sciences, Pyzhevski Lane 7, Moscow 119017, Russia
3
Diamond and Precious Metal Geology Institute, Siberian Branch, Russian Academy of Sciences, Lenin av. 39, Yakutsk 677000, Russia
4
Department of Geological Sciences, Jackson School of Geoscience, University of Texas at Austin, Austin, TX 78712-1692, USA
*
Author to whom correspondence should be addressed.
Geosciences 2024, 14(2), 41; https://doi.org/10.3390/geosciences14020041
Submission received: 9 December 2023 / Revised: 19 January 2024 / Accepted: 23 January 2024 / Published: 3 February 2024
(This article belongs to the Section Geochemistry)

Abstract

:
Provenance study plays an important role in paleogeographic and tectonic reconstructions. Detrital zircons are commonly used to identify sediment provenance; however, a wide range of detrital zircon ages in clastic rock often represent a fingerprint of reworked older terrigenous successions rather than ages of magmatism and metamorphism in the provenance area. This study focuses on the provenance of detrital rutile grains in the Triassic–Jurassic sandstones from Franz Josef Land and shows the importance of multiproxy approaches for provenance studies. Trace element data demonstrate that most rutile grains were sourced from metapelitic rocks, with a subordinate population having a metamafic origin. The Zr-in-rutile thermometer and U-Pb geochronology suggest that detrital rutile grains were predominantly derived from rocks that underwent amphibolite facies metamorphism during the Paleozoic era, with a predominance of the Carboniferous–Permian ages. Therefore, we suggest that the provenance area for the studied sandstones on Franz Josef Land has a similar geological history to the Taimyr region and Severnaya Zemlya archipelago. We propose that this crustal domain extends across the Kara Sea and forms the basement to the north and east of FJL, representing a proximal provenance for the studied Mesozoic terrigenous rocks. This domain experienced both Middle–Late Ordovician and Carboniferous–Permian metamorphism. The comparison of U-Pb dating and the geochemistry of rutile, U-Th/He, and U-Pb dating of zircons showed that detrital rutiles are the powerful toll in provenance restoration and can give additional constrains when a provenance area locates within collisional-convergent settings.

1. Introduction

There has been a significant increase in the interest of reconstructing the geological history of sedimentary basins using isotope-geochronological, isotope-geochemical and thermochronological methods in recent years. Clastic rocks, especially sandstone, are widely distributed in various depositional environments implying their usability in diverse fundamental and applied research [1,2,3]. The U-Pb dating of detrital zircons has become a routine method for provenance studies of clastic rocks [4,5,6]. This is largely due to the high content of zircon grains in the heavy fraction of sandstones, a well-established dating technique, and the presence of numerous laboratories around the world for analytical work. At the same time, zircons also contain a number of features that complicate the reconstruction of clastic provenances, if these are the sole archives used. The main complication involves the potential for zircons to undergo repeated reworking during the sedimentary processes. Therefore, a wide range of detrital zircon ages in a clastic rock often represents a fingerprint of reworked older terrigenous successions, as opposed to the ages of magmatism and metamorphism in the provenance area [7,8,9,10,11]. Moreover, zircons are usually sourced from felsic magmatic rocks and high-grade metamorphic rocks but are rare in mafic magmatic rocks [4,5,12].
To overcome these difficulties, analytical methods have been developed for dating other minerals which are less prone to multiple phases of reworking by sedimentary processes and/or are formed in a wider range of parent rocks, such as apatites, monazite, rutiles, and titanites, among others [13,14,15]. Detrital rutiles are of particular interest, as they can not only be dated by the U-Pb (LA-ICP-MS) method, but rare earth and trace element concentrations can also be used as indicators of the composition and the metamorphic grade of the provenance area [16,17,18]. Therefore, detrital zircon and rutile minerals can yield different, but complementary insights for provenance studies. Detrital U-Pb zircon ages mainly provide evidence for magmatic, and to a lesser extent, high-grade metamorphic events, whereas rutiles commonly form in relatively high-grade metamorphic facies (e.g., [18,19]). Furthermore, detrital rutiles, in contrast to zircons, have a significantly lower closure temperature for Pb (500–650 °C) [20,21,22]. Rutiles are not only sensitive to tectonic and metamorphic processes, but they also have the additional advantage of being common in both metapelitic and metamafic rocks. The usefulness of rutiles in provenance studies has been proven by numerous researchers [18,23,24,25,26,27,28]. In these studies, U-Pb dating of rutiles has been used to derive the age of upper amphibolite to lower granulite facies metamorphic events within the provenance area. In addition, trace element compositions of detrital rutiles provide information on the provenance rock composition and temperature conditions of metamorphism.
In this study, we generated U-Pb age data and trace element composition of detrital rutiles from three Triassic–Jurassic sandstone samples from Franz Josef Land (FJL; north-eastern Barents Sea). We constrain the rock composition found in the provenance area, along with the degree and timing of metamorphism, to create a more in-depth reconstruction of the provenance of the studied sediments and their transport directions. This study yields the first detrital rutile U-Pb ages and trace element chemistry from Mesozoic sandstones of the entire Barents Sea region.
In addition, we show how the age distribution of detrital rutiles apply to the reconstruction of the tectonic settings at the time of deposition, complementing and refining interpretations based solely on detrital zircons.

2. Geological Background

FJL represents the uplifted north-eastern part of the Barents Sea shelf, representing a rare window into the Mesozoic succession of the area (Figure 1, Figure 2 and Figure 3). The Barents continental shelves supposedly comprise Neoproterozoic and Early–Middle Paleozoic basement domains [29,30,31,32]. The pre-Mesozoic stratigraphic framework of FJL is based on the stratigraphy penetrated by the deep Nagurskaya well, drilled on Alexandra Land Island in the westernmost part of the archipelago [33,34,35], along with a detailed studied of the conglomerates containing reworked clasts of older stratigraphy in the south-eastern part of FJL [36].
Precambrian rocks comprise intensely deformed quartz mica schists and phyllites, overlain above an angular unconformity by sub-horizontally bedded clastic-carbonate Carboniferous succession [34,35]. The Triassic to Cretaceous clastic succession is exposed across FJL and intruded or overlain by late Early Cretaceous mafic dykes, sills, and basalts, attributed to the High Arctic Large Igneous Province (HALIP) [37,38,39] (Figure 2 and Figure 3). The correlation of the different Mesozoic sedimentary units across FJL is debated [35,40,41]. Here, we mainly follow the stratigraphic scheme of [35,40] due to its applicability for the whole archipelago.
The Induan–Norian succession (Belozemel, Matusevich, Ermakov, Graham Bell and Heiss formations) disconformably overlies Carboniferous succession in the Nagurskaya well. The Triassic strata comprise alternating argillites, sandstones, and siltstones, with subordinate beds of clayey limestones reaching a thickness of 3500 m. The Vasiliev Formation (Rhaetian; 100–370 m in thickness) comprises polymictic sandstones and sands with beds of conglomerates and gritstones, along with rare beds of siltstones and argillites [35,40]. A markable hiatus occurred between Norian and Rhaetian deposits across FJL [42].
Figure 2. Simplified geological map of southern part of FJL with location of dated samples (simplified modified after [35,40]).
Figure 2. Simplified geological map of southern part of FJL with location of dated samples (simplified modified after [35,40]).
Geosciences 14 00041 g002
The Tegethoff Formation (Hettangian–Lower Toarcian; 60–350 m in thickness) unconformably overlies upper Triassic succession and mainly comprises coarse- to medium-grained polymictic sands and sandstones, gritstones and conglomerate beds, with occasional thin silt beds [35,40]. The Ganza Formation (Toarcian–Volgian; 40–270 m in thickness) has a patchy distribution across the archipelago due to subsequent early Cretaceous erosion. It mainly comprises alternating argillites and siltstones with occasional units of clayey limestone and rare beds of sandstone. The Lamon Formation (Oxfordian–Valanginian) comprises sandstones and sands with beds of siltstone and argillite [35,40,41].
Figure 3. Simplified stratigraphic chart of Mesozoic strata of FJL (compiled from [33,35,40,41,42]. The major unconformities are pre-Rhaetian, pre-Aalenian, pre-Aptian.
Figure 3. Simplified stratigraphic chart of Mesozoic strata of FJL (compiled from [33,35,40,41,42]. The major unconformities are pre-Rhaetian, pre-Aalenian, pre-Aptian.
Geosciences 14 00041 g003
The Armitidj Formation (Hauterivian (?)–Aptian) unconformably overlies various Triassic–Lower Cretaceous formations, comprising alternating basalts and tuffaceous sandstones. The age of formation is mainly based on the isotopic dating of mafic rocks with 40Ar-39Ar dating yielded ages ranging from 189 to 125 Ma [43,44,45,46,47], while the most reliable TIMS U-Pb zircon dates yielded a crystallization age of 122.7 Ma for a thick sill in the Severnaya Well (Graham Bell Island) [37].

3. Materials and Methods

3.1. Geological Position of Dated Samples

Sample 7-v15-9 (Wilczek Land, Ganza Cape area) was collected from outcrops located 3 km to the north-west of the Ganza Cape, where planar and cross-bedded medium- to fine-grained sandstones with rare siltstone layers (Vasiliev Formation; Rhaetian) crop out (Figure 2, Figure 3 and Figure 4 and Table 1).
Sample 11-v15-1 was collected from outcrops located 1 km to the north of the Ganza Cape, where black argillites with subordinate thin siltstone layers of the Ganza Formation crop out (Figure 3). The sample has been dated as the latest Kimmeridgian, based on ammonite findings from this locality [25] (Figure 2, Figure 3 and Figure 4 and Table 1).
Sample 15AP30 (Hall Island, Tegethoff Cape) was collected from the intercalating clays and fine-grained sandstones of the Lower Jurassic Tegethoff Formation (Figure 2, Figure 3 and Figure 4 and Table 1).

3.2. Analytical Techniques

Samples were crushed and the heavy minerals were concentrated using standard techniques. Zircon monofraction were separated following a standard scheme: grinding, sieving into size fractions, passing of the fraction of <0.25 mm through a centrifugal concentrator, electromagnetic treatment of the obtained heavy fraction, and the final treatment of the concentrate with a heavy liquid. All analyses were carried out at the UTChron geochronology facility in the Department of Geosciences at the University of Texas, Austin. The rutile grains were hand-picked from the bulk-heavy mineral separate, mounted in epoxy, and polished for LA-ICP-MS U-Pb analysis. The sample pucks were loaded into a large-volume Helex sample cell and analyzed with a magnet sector, single collector Element2 ICP-MS with a Photon Machines Analyte G.2 excimer laser (Laser source-Advanced Technology Lasers (ATL) in Wermelskirchen, Germany; the frame that the source sits in that contains all the mirrors, lenses, cameras, etc was made by Photon Machines in Bozeman, Montana, United States). Rutile U-Pb analysis was conducted using a 40-μm spot size. Rutile R10 (1090 ± 0.9 Ma; [48]) was used as a primary reference standard and rutile R19 (489.5 ± 0.9 Ma; [5]) as a secondary reference standard. A primary 238U/206Pb standard R10 rutile [48] was used to calculate ages and control for fractionation, while a secondary standard R19 rutile [5] provided an independent verification of ages. The data were reduced using the VizualAgeDRS in Iolite [49,50,51]. Pbc correction was based on the model Pb composition of [52].
A 40-μm spot size was used for trace element LA-ICP-MS determinations and concentrations were calibrated against the SRM 610 (NIST) reference glass standard. Raw trace element data were reduced using Iolite 3.7 data reduction software and Trace Element_IS [50]. The rutile U-Pb ages were corrected using a 208Pb correction [15]. Zr-in-rutile crystallization temperatures were calculated using the calibration of [53], with the Tomkins thermometer for a 10 kbar-α-quartz as a default setting, since pressure information is unavailable for detrital rutile grains [17]. Detrital rutile U-Pb analytical results are provided in Table S1.

4. Results

The most frequent rutile grains are dark yellowish to reddish brown. The crystal size is variable and, on average, ranges from approximately 70 μm to 250 μm. Most rutile grains are angular, subangular, or subrounded in shape.

4.1. Detrital Rutile Geochemistry

Trace element concentrations of 209 rutile grains from the three sandstone samples were measured and the results are listed in Table S2.

4.1.1. Sample 7-v15-9

Detrital rutile Cr and Nb concentrations exhibit large variations (Cr: 3.7–4918 ppm, Nb: 26–8320 ppm), with all of the grains plotted on the Cr-Nb diagram of [17]. Based on this diagram (Figure 4), about 80% of the detrital rutiles were sourced from metapelitic rocks and 20% from metamafic rocks. A Zr concentration of detrital rutiles ranges from 15.29 to 3718 ppm. The temperatures calculated for detrital rutiles in sample 7-v15-9 range between 464 °C and 902 °C (Figure 5).

4.1.2. Sample 15AP30

Detrital rutile Cr and Nb concentrations also exhibit large variations (Cr: 6–5582 ppm, Nb: 7–8500 ppm). Based on the Cr-Nb diagram (Figure 5), about 70% of the detrital rutiles originated from metapelitic rocks and 30% originated from metamafic rocks. The Zr concentration of detrital rutiles range from 21 to 5100 ppm. The rutile grains yielded temperatures between 481 °C and 943 °C. This temperature distribution suggests that most grains were sourced from amphibolite-facies metamorphic rocks, while only a few were derived from greenschist or granulite-facies metamorphic rocks (Figure 5).

4.1.3. Sample 11-V15-11

The detrital rutile Cr and Nb concentrations also exhibit large variations (Cr: 9–3709 ppm, Nb: 8–8390 ppm). Based on the Cr-Nb diagram (Figure 5), about 87% of the detrital rutiles were transported from metapelitic rocks and 13% from metamafic rocks. The Zr concentration of detrital rutiles range from 23 to 16,000 ppm. The calculated rutile formation temperatures for the detrital rutile grains range between 485 °C and 1122 °C. This temperature distribution suggests that most grains originated from amphibolite-facies metamorphic rocks, while only 20 grains were derived from greenschist or granulite-facies rocks (Figure 6).

4.2. U-Pb Dating of Rutile

4.2.1. Sample 7-v15-9

Most of the dated rutiles are Paleozoic in age (82%). Permian and Carboniferous rutiles are abundant, and comprise 33% of the dated population, forming multiple peaks at ca. 270 and 300–350 Ma. Devonian rutiles (28%) group at ca. 380–400 Ma. Silurian and Ordovician grains together comprise 13% of the dated population and group at ca. 440 and 490 Ma. Precambrian grains (18%) do not form any prominent peaks (Figure 7).

4.2.2. Sample 15AP30

Permian and Carboniferous rutiles comprise 27% of the dated population and form peaks at ca. 330 and 350 Ma. Devonian rutiles comprise 21% of the dated grains and form a subordinate peak at 390 Ma. Silurian, Ordovician, and Cambrian rutiles together comprise 37% of the population and form peaks at ca. 400, 430 and 459 Ma. Precambrian rutiles (10%) do not form any significant peaks (Figure 7).

4.2.3. Sample 11-V15-1

Most of the dated rutiles are Paleozoic in age (83%), with Permian rutiles (24%) forming peaks at ca. 250 and 295 Ma. The Carboniferous rutiles comprise 19% of the dated population and group at ca. 300–310 and 340–350 Ma. The Devonian–Cambrian zircons (39%) form multiple peaks at ca. 380–400 and 430–450 Ma. Precambrian zircons (16%) do not form reliable groups (Figure 7).

5. Provenance Interpretation

5.1. Previous Provenance Study

A previous provenance study of the Triassic–Cretaceous strata on FJL on the same samples was based on the U-Pb and (U-Th)/He dating of detrital zircons [36]. Detrital zircon U-Pb dating revealed that Triassic grains form a small peak at 220–230 Ma. Most of the dated zircons (54–66%) are Paleozoic in age, with a predominance of Permian and Carboniferous grains grouping at ca. 250, 295, 300–310 and 340–350 Ma. Devonian–Ordovician zircons mainly group at ca. 380–400 and 430–450 Ma. Precambrian zircons form subordinate peaks mainly between 550–650, 800–1000 and 1700–1800 Ma.
Detrital zircon (U-Th)/He (ZHe) dating previously [36] was carried out for samples 7-v15-9 and 15AP30. All obtained ZHe ages were older than the depositional ages of the studied deposits, suggesting that these rocks have not been buried deeper than 6–7 km since deposition (assuming a typical continental crust geothermal gradient of 30 °C/km). Therefore, the ZHe ages reflect the timing of exhumation and cooling in the provenance area. A total of eighteen ZHe ages showed a 180–349 Ma age range, with most of these ages grouping at approximately 225 Ma [36].

5.2. Detrital Rutile

Based on the geochemistry of detrital rutiles, most grains were derived from metapelitic rocks with a subordinate rutile population from metamafic rocks (Figure 6). The calculated rutile formation temperatures suggest that most grains were derived from amphibolite-facies metamorphic rocks, with a subordinate number from greenschist and granulite-facies metamorphic rocks. Therefore, based on our geochemical data, we suggest that the provenance area for Mesozoic clastics on FJL was predominantly composed of amphibolite-facies metamorphic rocks.
The distribution of detrital rutile ages within the dated samples is very similar, suggesting a common provenance area for the studied Triassic–Jurassic sedimentary rocks (Figure 7). The few Paleoproterozoic (ca. 1800–1600 Ma) and Mesoproterozoic (1300–1000 Ma) rutile populations can be correlated with magmatic and metamorphic rocks within the basement of the Baltica [55,56,57,58]. Neoproterozoic to Cambrian rutiles (650–520 Ma) can be correlated with the main magmatic and metamorphic event within the Timanian orogen, which formed as a result of the latest Neoproterozoic to the earliest Cambrian orogeny along the northeatsern margin of Baltica (present day coordinates) ([59,60] and refences therein). A prominent detrital rutile population comprising ages between 470 and 420 Ma could be possibly correlated with coeval magmatic and metamorphic events within the Severnaya Zemlya archipelago [61,62,63].
A Middle–Late Devonian population is prominent in sample 7-v15-9; however, a smaller population of this age can be found in the other samples. A source for the Middle–Late Devonian rutiles is unclear, since coeval metamorphic events have not been documented from onshore. However, Late Devonian-mafic magmatism and coeval rifting occurred in northern Novaya Zemlya (Barents Sea region) [64]. Furthermore, significant exhumation of the northern Kara Sea region during the Middle–Late Devonian has been reconstructed from the (U-Th)/He zircon ages [65]. Therefore, we propose that this Middle–Late Devonian tectonism was associated with some metamorphism, forming metamorphic rocks which were a source of the Middle–Late Devonian rutiles in our samples.
The most prominent detrital rutile populations indicate major metamorphic events within the provenance area occurred at ca. 350–320 Ma and again between ca. 280 and 260 Ma. These Late Paleozoic magmatic and tectonic events can be correlated to the Uralian orogeny, formed as a result of a continental collision between Kazakhstan, Siberia, and Laurussia [66,67,68]. The suture of the Uralian orogeny stretches onshore from the Aral Sea to the Polar Urals, although its northward continuation under the northern West Siberia Basin and further to the north and east are still debated [32,69,70,71,72]. However, recent studies pointed out that it continues as far as the Taimyr Peninsula (e.g., [32,71] and references therein). Furthermore, voluminous magmatism and metamorphism occurred between 315 and 288 Ma across northern Taimyr and the southern part of the Severnaya Zemlya archipelago ([71,72] and references therein).
The Upper Triassic and Lower Jurassic strata of FJL comprise a coarse-grained clastic succession with conglomerate beds and were clearly derived from a proximal provenance area. The previous study by [36] proposed that the provenance area for the studied sandstones on Franz Josef Land has a comparable geological history to the Taimyr region and Severnaya Zemlya archipelago, based on the detrital zircon U-Pb dating. They further proposed that this crustal domain extends across the Kara Sea and forms basement to the east and north of FJL, representing a proximal provenance for the studied Mesozoic clastics. Our new detrital rutile data tentatively support this model but in addition, we propose that not only magmatic but also significant metamorphic events affected the provenance area in the Middle–Late Ordovician and Carboniferous–Permian.

5.3. Comparison between U-Pb Dating of Rutiles and Zircons, (U-Th)/He Dating of Zircons and Their Application to Provenance Studies

Since both zircon and rutile are resistant to chemical and physical breakdown, they can be reworked from their host magmatic or metamorphic rocks and preserved as detrital grains within sedimentary rocks. To date, there are thousands of provenance studies based on detrital zircon U-Pb dating, mainly due to well-established analytical techniques. Some of these studies incorporated U-Pb dating of both zircons and rutiles, with even fewer accompanied by an additional geo- or thermochronometer (for example [73]). Here, we compare the data collected from detrital rutiles during this study with previous detrital zircon U-Pb and (U-Th)/He data generated from the same samples by [36], to evaluate the relative merits of the different minerals and analytical approaches for provenance studies (Figure 8).
Rutile ages are often characterized by a unimodal distribution, indicative of the youngest metamorphic event(s) within the provenance, while zircons from the same samples often yield a multimodal age distribution, suggesting a significant reworking of older grains. For example, unimodal rutile age distributions have been obtained from the Upper Cretaceous–Eocene strata of western Anatolia [74], Lower Carboniferous flysch of the Istanbul Zone [19], Late Ordovician units from Saxo-Thuringia [75], Precambrian rocks of central Australia [76], and Cambrian sandstones from Israel and Jordan [77]. Multimodal age spectra are less commonly observed in detrital rutiles, although they have been reported from modern sands of the Mississippi River [78], ice-rafted debris in the North Atlantic [79], and Neogene to recent Himalayan River deposits [27,28]. Moreover, a closer inspection of the data suggests that these non-unimodal age distributions are often better characterized as bimodal rather than multimodal, reflecting either two distinct provenance areas or complex metamorphic events within a single provenance area.
Our comparison of detrital zircon and rutile U-Pb ages suggest that rutiles yield a broadly similar age distribution to detrital zircons for the Mesozoic clastics of FJL. However, Precambrian detrital zircons are more abundant than rutiles, that is possibly explained by the ability of zircon to generally survive the rock cycle from sedimentary to high-grade metamorphic conditions, while rutile usually breakdown at the beginning of the greenschist metamorphic facies and newly forms at upper amphibolite facies conditions [80]. Furthermore, the 470–450 Ma detrital rutile population suggests a significant metamorphic event within the provenance area, yet a significant coeval detrital zircon population of this age is lacking. Therefore, the detrital rutiles clearly highlight a significant metamorphic event which would not have been interpreted from the analysis of detrital zircon U-Pb ages in isolation. The ZHe ages obtained by [36] for two samples (7-v15-9, 15AP30) indicate a significant Late Triassic exhumation event in the provenance area, along with a subordinate Late Paleozoic tectonic event. The Upper Triassic sandstones (sample 7-v15-9) contain Late Triassic ZHe and U-Pb detrital zircon ages but lack rutiles of coeval ages. The integration of all three analytical techniques can be used to suggest that the Late Triassic magmatism and exhumation, based on the detrital zircon ZHe and U-Pb ages, were not accompanied by a significant metamorphism, based on the lack of coeval rutile ages. Lower Jurassic sandstones (sample 15AP30) contain evidence for two pronounced exhumation events of latest Permian and latest Triassic age within the provenance area, but both events were not accompanied by coeval magmatism and metamorphism. The detrital rutile and zircon age spectra of the studied sandstones are both multimodal, suggestive of complex drainage systems supplying clastics from multiple provenance regions, or a single provenance region with a complex geological history involving multiple metamorphic and magmatic events [73].
A comparison of detrital zircon ZHe and U-Pb ages, along with detrital rutile U-Pb ages, suggests that the youngest significant exhumation event occurred during the latest Triassic age within the provenance area. The geodynamic cause of this Triassic/Jurassic tectonic event within the Arctic region is unclear. However, it affected an extensive region encompassing the Taimyr Severnaya Zemlya fold belt, Novaya Zemlya archipelago, and possibly extended across the central part of the Kara Sea [29,81]. Since coeval magmatism and metamorphism have not been reported across the study area, along with an absence of collisional events of this age, this pronounced tectonic event could represent an intracontinental orogen encompassing the Taimyr region and stretching across the central and possibly northern part of the Kara Sea to form the proximal provenance area for studied calstics. Such orogens are comparatively rare in the geological record and are therefore poorly understood [82]. We speculate that the onset of the cooling phase is likely related to far-field propagation and the formation of an intracontinental orogen. Therefore, reconstruction of intracontinental orogens require additional lower temperature constraints beyond the applicability of the detrital rutile and zircon U-Pb dating.
Detrital minerals represent a powerful archive for the reconstruction of the tectonic settings of sedimentary basins and providing an insight into tectonic processes affecting provenance regions. A cumulative proportion distribution against a growth-deposition ages diagram based on detrital zircon ages has been proposed by [4] and applied to detrital rutiles by [24], representing a powerful tool for deciphering tectonic settings of basins and the provenance areas. Therefore, we applied the same approach for our dataset (Figure 9) to demonstrate that both detrital rutile and zircon data plot in the convergent to collisional basin fields. The detrital rutile and zircon cumulative distributions are similar but not identical, with the large number of Precambrian detrital zircons with ages which have not been reset during the younger tectonic and metamorphic events. Consequently, detrital rutiles appear to be better candidates for reconstructing provenance areas formed in convergent to collisional settings, since younger tectonic events are better preserved compared to detrital zircons. The global database of detrital rutile data generated to date has the potential to yield important information on the reconstruction of tectonic settings within provenance areas and the tectonothermal evolution of provenance areas formed within convergent to collisional geodynamic settings.

6. Conclusions

Cr and Nb concentrations in detrital rutiles show large variations. According to the Cr-Nb diagram, a majority (70–87%) of the detrital rutiles were derived from the metapelithic with subordinate populations form the metamafic rocks (13–30%). The calculated rutile formation temperatures for detrital rutile grains showed that a majority of the detrital rutiles grains were derived from the amphibolite-facies metamorphic rocks, while only a few were derived from greenschist and granulite-facies metamorphic rocks.
Most of the dated grains are Paleozoic in age (77–85%). Permian and Carboniferous rutiles are abundant and comprise the dated population, forming multiple peaks at ca. 270 and 300–350 Ma. Devonian rutiles form a subordinate population at ca. 380–400 Ma. Silurian and Ordovician grains form groups at ca. 440 and 490 Ma. Precambrian grains do not form any significant peaks. Thus, the provenance area of the studied sandstones has the comparable geological history, structure, including multiple episodes of magmatic activity to the Taimyr region and Severnaya Zemlya archipelago, and propose that these crustal domain stretch across the Kara Sea, forming the basement to the east and north of FJL, representing the main proximal provenance for the studied Mesozoic clastics. The comparison of the results of the U-Pb dating of rutile, U-Th/He, and U-Pb dating of detrital rutiles are the powerful toll in provenance restoration and can give additional constrains when the provenance area locates within collisional-convergent settings. Moreover, low-temperature thermochronology is necessary for revealing exhumation of the provenance area in intracratonic orogen settings.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/geosciences14020041/s1, Table S1. Results of U-Pb dating of detrital rutiles. Table S2. Trace element concentrations.

Author Contributions

Conceptualization, V.E. and A.P.; methodology, V.E.; formal analysis, V.E. and D.S.; investigation, V.E., A.P. and D.S.; writing—original draft preparation V.E., A.P. and D.S. All authors have read and agreed to the published version of the manuscript.

Funding

Mesozoic stratigraphic and paleogeographic study were supported by RSF grant 21-17-00245. Interpretation of isotopic study and correlation with Taimyr-Severnaya Zemlya were supported by RSF grant 20-17-00169. Field work of AP partly supported by DPMGI SB RAS (FUFG-2024-0005).

Data Availability Statement

The original contributions presented in the study are included in the supplementary material, further inquiries can be directed to the corresponding author.

Acknowledgments

We thank two anonymous reviewers for their constructive comments, suggestions, and corrections, which very much helped to improve the manuscript. Thanks to James Barnet for editing English.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Galloway, W.E.; Hobday, D.K. Terrigenous Clastic Depositional Systems; Springer: New York, NY, USA, 1983; ISBN 978-1-4684-0172-1. [Google Scholar]
  2. Pettijohn, F.J.; Potter, P.E.; Siever, R. Sand and Sandstone; Springer: New York, NY, USA, 1987; ISBN 978-0-387-96350-. [Google Scholar]
  3. Jamil, M.; Siddiqui, N.A.; Ahmed, N.; Usman, M.; Umar, M.; Rahim, H.U.; Imran, Q.S. Facies Analysis and Sedimentary Architecture of Hybrid Event Beds in Submarine Lobes: Insights from the Crocker Fan, NW Borneo, Malaysia. JMSE 2021, 9, 1133. [Google Scholar] [CrossRef]
  4. Cawood, P.A.; Hawkesworth, C.J.; Dhuime, B. Detrital Zircon Record and Tectonic Setting. Geology 2012, 40, 875–878. [Google Scholar] [CrossRef]
  5. Fedo, C.M.; Sircombe, K.N.; Rainbird, R.H. Detrital Zircon Analysis of the Sedimentary Record. In Zircon; Hanchar, J.M., Hoskin, P.W.O., Eds.; De Gruyter: Berlin, Germany, 2003; pp. 277–304. ISBN 978-1-5015-0932-2. [Google Scholar]
  6. Gehrels, G. Detrital Zircon U-Pb Geochronology: Current Methods and New Opportunities. In Tectonics of Sedimentary Basins; Busby, C., Azor, A., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2012; pp. 45–62. [Google Scholar] [CrossRef]
  7. Hadlari, T.; Swindles, G.T.; Galloway, J.M.; Bell, K.M.; Sulphur, K.C.; Heaman, L.M.; Beranek, L.P.; Fallas, K.M. 1.8 Billion Years of Detrital Zircon Recycling Calibrates a Refractory Part of Earth’s Sedimentary Cycle. PLoS ONE 2015, 10, e0144727. [Google Scholar] [CrossRef] [PubMed]
  8. Verhaegen, J.; Von Eynatten, H.; Dunkl, I.; Weltje, G.J. Detrital Zircon Geochronology and Heavy Mineral Analysis as Complementary Provenance Tools in the Presence of Extensive Weathering, Reworking and Recycling: The Neogene of the Southern North Sea Basin. Geol. Mag. 2021, 158, 1572–1584. [Google Scholar] [CrossRef]
  9. Schwartz, T.M.; Schwartz, R.K.; Weislogel, A.L. Orogenic Recycling of Detrital Zircons Characterizes Age Distributions of North American Cordilleran Strata. Tectonics 2019, 38, 4320–4334. [Google Scholar] [CrossRef]
  10. Dickinson, W.R.; Lawton, T.F.; Gehrels, G.E. Recycling Detrital Zircons: A Case Study from the Cretaceous Bisbee Group of Southern Arizona. Geology 2009, 37, 503–506. [Google Scholar] [CrossRef]
  11. Andersen, T.; Van Niekerk, H.; Elburg, M.A. Detrital Zircon in an Active Sedimentary Recycling System: Challenging the ‘Source-to-sink’ Approach to Zircon-based Provenance Analysis. Sedimentology 2022, 69, 2436–2462. [Google Scholar] [CrossRef]
  12. Burnham, A.D. Zircon. In Reference Module in Earth Systems and Environmental Sciences; Elsevier: Amsterdam, The Netherlands, 2018; p. B978012409548910911X. ISBN 978-0-12-409548-9. [Google Scholar]
  13. Chew, D.M.; Sylvester, P.J.; Tubrett, M.N. U–Pb and Th–Pb Dating of Apatite by LA-ICPMS. Chem. Geol. 2011, 280, 200–216. [Google Scholar] [CrossRef]
  14. Gaschnig, R.M. Benefits of a Multiproxy Approach to Detrital Mineral Provenance Analysis: An Example from the Merrimack River, New England, USA. Geochem. Geophys. Geosyst. 2019, 20, 1557–1573. [Google Scholar] [CrossRef]
  15. Zack, T.; Stockli, D.F.; Luvizotto, G.L.; Barth, M.G.; Belousova, E.; Wolfe, M.R.; Hinton, R.W. In situ U–Pb rutile dating by LA-ICP-MS: 208Pb correction and prospects for geological applications. Contrib. Mineral. Petrol. 2011, 162, 515–530. [Google Scholar] [CrossRef]
  16. Zack, T.; Kronz, A.; Foley, S.F.; Rivers, T. Trace element abundances in rutiles from eclogites and associated garnet mica schists. Chem. Geol. 2002, 184, 97–122. [Google Scholar] [CrossRef]
  17. Triebold, S.; von Eynatten, H.; Zack, T. A recipe for the use of rutile in sedimentary provenance analysis. Sediment. Geol. 2012, 282, 268–275. [Google Scholar] [CrossRef]
  18. Meinhold, G. Rutile and Its Applications in Earth Sciences. Earth-Sci. Rev. 2010, 102, 1–28. [Google Scholar] [CrossRef]
  19. Okay, N.; Zack, T.; Okay, A.I.; Barth, M. Sinistral Transport along the Trans-European Suture Zone: Detrital Zircon–Rutile Geochronology and Sandstone Petrography from the Carboniferous Flysch of the Pontides. Geol. Mag. 2011, 148, 380–403. [Google Scholar] [CrossRef]
  20. Cherniak, D.J. Pb Diffusion in Rutile. Contrib. Miner. Petrol. 2000, 139, 198–207. [Google Scholar] [CrossRef]
  21. Vry, J.K.; Baker, J.A. LA-MC-ICPMS Pb–Pb Dating of Rutile from Slowly Cooled Granulites: Confirmation of the High Closure Temperature for Pb Diffusion in Rutile. Geochim. Cosmochim. Acta 2006, 70, 1807–1820. [Google Scholar] [CrossRef]
  22. Pereira, I.; Storey, C.D. Detrital rutile: Records of the deep crust, ores and fluids. Lithos 2023, 438–439, 107010. [Google Scholar] [CrossRef]
  23. Thomsen, T.B.; Knudsen, C.; Hinchey, A.M. Investigations of Detrital Zircon, Rutile and Titanite from Present-Day Labrador Drainage Basins: Fingerprinting the Grenvillean Front. GEUS Bull. 2015, 33, 77–80. [Google Scholar] [CrossRef]
  24. Pereira, I.; Storey, C.D.; Strachan, R.A.; Bento Dos Santos, T.; Darling, J.R. Detrital rutile ages can deduce the tectonic setting of sedimentary basins. Earth Planet. Sci. Lett. 2020, 537, 116193. [Google Scholar] [CrossRef]
  25. Rösel, D.; Zack, T.; Möller, A. Interpretation and Significance of Combined Trace Element and U–Pb Isotopic Data of Detrital Rutile: A Case Study from Late Ordovician Sedimentary Rocks of Saxo-Thuringia, Germany. Int. J. Earth Sci. 2019, 108, 1–25. [Google Scholar] [CrossRef]
  26. Kotowski, J.; Nejbert, K.; Olszewska-Nejbert, D. Rutile Mineral Chemistry and Zr-in-Rutile Thermometry in Provenance Study of Albian (Uppermost Lower Cretaceous) Terrigenous Quartz Sands and Sandstones in Southern Extra-Carpathian Poland. Minerals 2021, 11, 553. [Google Scholar] [CrossRef]
  27. Bracciali, L.; Najman, Y.; Parrish, R.R.; Akhter, S.H.; Millar, I. The Brahmaputra Tale of Tectonics and Erosion: Early Miocene River Capture in the Eastern Himalaya. Earth Planet. Sci. Lett. 2015, 415, 25–37. [Google Scholar] [CrossRef]
  28. Bracciali, L.; Parrish, R.R.; Najman, Y.; Smye, A.; Carter, A.; Wijbrans, J.R. Plio-Pleistocene Exhumation of the Eastern Himalayan Syntaxis and Its Domal ‘Pop-Up’. Earth-Sci. Rev. 2016, 160, 350–385. [Google Scholar] [CrossRef]
  29. Drachev, S.S. Fold Belts and Sedimentary Basins of the Eurasian Arctic. Arktos 2016, 2, 21. [Google Scholar] [CrossRef]
  30. Drachev, S.S.; Malyshev, N.A.; Nikishin, A.M. Tectonic History and Petroleum Geology of the Russian Arctic Shelves: An Overview. In Geological Society, London, Petroleum Geology Conference Series; The Geological Society of London: London, UK, 2010; Volume 7, pp. 591–619. [Google Scholar] [CrossRef]
  31. Henriksen, E.; Ryseth, A.E.; Larssen, G.B.; Heide, T.; Rønning, K.; Sollid, K.; Stoupakova, A.V. Chapter 10. Tectonostratigraphy of the Greater Barents Sea: Implications for Petroleum Systems. Geol. Soc. Lond. Mem. 2011, 35, 163–195. [Google Scholar] [CrossRef]
  32. Pease, V.; Drachev, S.; Stephenson, R.; Zhang, X. Arctic Lithosphere—A Review. Tectonophysics 2014, 628, 1–25. [Google Scholar] [CrossRef]
  33. Dibner, V.D. (Ed.) Geology of Franz Josef Land; Nor. Polarinst. Meddelelser: Tromsø, Norway, 1998; Volume 146, 199p. [Google Scholar]
  34. Gramberg, I.S.; Shkola, I.V.; Bro, E.G.; Shekhodanov, V.A.; Armishev, A.M. Parametric Wells on the Islands of the Barents and Kara Seas. Sov. Geol. 1985, 1, 95–98. (In Russian) [Google Scholar]
  35. Makariev, A.A. (Ed.) State Geological Map of the Russian Federation. Scale 1: 1,000,000 (New Series). Sheet U-37-40—Franz Josef Land (Northern Islands). Explanatory Letter; Publishing House of VSEGEI: Saint Petersburg, Russia, 2006; 272p. (In Russian) [Google Scholar]
  36. Ershova, V.; Prokopiev, A.; Stockli, D.; Kurapov, M.; Kosteva, N.; Rogov, M.; Khudoley, A.; Petrov, E.O. Provenance of the Mesozoic succession of Franz Josef Land (north-eastern Barents Sea): Paleogeographic and tectonic implications for the High Arctic. Tectonics 2022, 41, e2022TC007348. [Google Scholar] [CrossRef]
  37. Corfu, F.; Polteau, S.; Planke, S.; Faleide, J.I.; Svensen, H.; Zayoncheck, A.; Stolbov, N. U–Pb Geochronology of Cretaceous Magmatism on Svalbard and Franz Josef Land, Barents Sea Large Igneous Province. Geol. Mag. 2013, 150, 1127–1135. [Google Scholar] [CrossRef]
  38. Senger, K.; Tveranger, J.; Ogata, K.; Braathen, A.; Planke, S. Late Mesozoic Magmatism in Svalbard: A Review. Earth-Sci. Rev. 2014, 139, 123–144. [Google Scholar] [CrossRef]
  39. Levskii, L.K.; Stolbov, N.M.; Bogomolov, E.S.; Vasil’eva, I.M.; Makar’eva, E.M. Sr-Nd-Pb Isotopic Systems in Basalts of the Franz Josef Land Archipelago. Geochem. Int. 2006, 44, 327–337. [Google Scholar] [CrossRef]
  40. Makariev, A.A. (Ed.) State Geological Map of the Russian Federation. Scale 1: 1000 000 (Third Generation). Series North-Karsko-Barents Sea. Sheet U-41-44-Franz Josef Land (Eastern Islands). Explanatory Note; Publishing House of VSEGEI: Saint Petersburg, Russia, 2011; 220p. (In Russian) [Google Scholar]
  41. Kosteva, N.N. Stratigraphy of Jurassic-Cretaceous deposits of Franz Joseph Land. Arctica Antarct. 2005, 4, 16–32. [Google Scholar]
  42. Repin, Y.S.; Fedorova, A.A.; Bystrova, V.V.; Kulikova, N.K.; Polubotko, I.V. Mesozoic strata of Barents Sea sedimentary basin. In Stratigraphy and Its Role in Develop of Oil and Gas Complex of Russia; VNIGRII: Saint Petersburg, Russia, 2007; pp. 112–137. (In Russian) [Google Scholar]
  43. Abashev, V.V.; Metelkin, D.V.; Vernikovsky, V.A.; Vasyukova, E.A.; Mikhaltsov, N.E. Early Cretaceous Basalts of the Franz Josef Land Archipelago: Correspondence of New 40Ar/39Ar and Paleomagnetic Data. Dokl. Earth Sci. 2020, 493, 495–498. [Google Scholar] [CrossRef]
  44. Grachev, A.F. A new view on the origin of magmatism of the Franz Joseph Land. Izvestiya Phys. Solid Earth 2001, 9, 49–61. [Google Scholar]
  45. Karyakin, Y.V.; Sklyarov, E.V.; Travin, A.V. Plume Magmatism at Franz Josef Land. Petrology 2021, 29, 528–560. [Google Scholar] [CrossRef]
  46. Koryakin, Y.V.; Shipilov, E.V. Geochemical Specifics and 40Ar/39Ar Age of the Basaltoid Magmatism of the Alexander Land, Northbrook, Hooker, and Hayes Islands (Franz Josef Land Archipelago). Dokl. Earth Sci. 2009, 425, 260–263. [Google Scholar] [CrossRef]
  47. Shipilov, E.V.; Karyakin, Y.V. Dikes of Hayes Island (Frantz Josef Land Archipelago): Tectonic Position and Geodynamic Interpretation. Dokl. Earth Sci. 2014, 457, 814–818. [Google Scholar] [CrossRef]
  48. Luvizotto, G.L.; Zack, T.; Meyer, H.P.; Ludwig, T.; Triebold, S.; Kronz, A.; Münker, C.; Stockli, D.F.; Prowatke, S.; Klemme, S.; et al. Rutile crystals as potential trace element and isotope mineral standards for microanalysis. Chem. Geol. 2009, 261, 346–369. [Google Scholar] [CrossRef]
  49. Ludwig, K.R. Isoplot 3.00: A Geochronological Toolkit for Microsoft Excel; Berkeley Geochronology Center: Berkeley, CA, USA, 2003; Volume 70. [Google Scholar]
  50. Paton, C.; Hellstrom, J.; Paul, B.; Woodhead, J.; Hergt, J. Iolite: Freeware for the visualisation and processing of mass spectrometric data. J. Anal. At. Spectrom. 2011, 26, 2508. [Google Scholar] [CrossRef]
  51. Petrus, J.A.; Kamber, B.S. VizualAge: A Novel Approach to Laser Ablation ICP-MS U-Pb Geochronology Data Reduction. Geostand. Geoanal. Res. 2012, 36, 247–270. [Google Scholar] [CrossRef]
  52. Stacey, J.S.; Kramers, J.D. Approximation of terrestrial lead isotope evolution by a two-stage model. Earth Planet. Sci. Lett. 1975, 26, 207–221. [Google Scholar] [CrossRef]
  53. Tomkins, H.S.; Powell, R.; Ellis, D.J. The pressure dependence of the zirconium-in-rutile thermometer. J. Metamorph. Geol. 2007, 25, 703–713. [Google Scholar] [CrossRef]
  54. Bucher, K.; Grapes, R. Petrogenesis of Metamorphic Rocks; Springer: Berlin/Heidelberg, Germany, 2011; ISBN 978-3-540-74168-8. [Google Scholar]
  55. Andersson, U.B.; Sjöström, H.; Högdahl, K.H.O.; Eklund, O. The Transscandinavian Igneous Belt, evolutionary models. In The Transscandinavian Igneous Belt (TIB) in Sweden: A Review of Its Character and Evolution; Special Paper; Geological Survey of Finland: Espoo, Finland, 2004; Volume 37, pp. 104–112. [Google Scholar]
  56. Bogdanova, S.V.; Bingen, B.; Gorbatschev, R.; Kheraskova, T.N.; Kozlov, V.I.; Puchkov, V.N.; Volozh, Y.A. The East European Craton (Baltica) before and during the Assembly of Rodinia. Precambr. Res. 2008, 160, 23–45. [Google Scholar] [CrossRef]
  57. Gorbatschev, R. The Transscandinavian Igneous Belt—Introduction and background. In The Transscandinavian Igneous Belt (TIB) in Sweden: A Review of Its Character and Evolution; Special Paper; Geological Survey of Finland: Espoo, Finland, 2004; Volume 37, pp. 9–15. [Google Scholar]
  58. Korja, A.; Lahtinen, R.; Nironen, M. The Svecofennian Orogen: A Collage of Microcontinents and Island Arcs. Geol. Soc. Lond. Mem. 2006, 32, 561–578. [Google Scholar] [CrossRef]
  59. Gee, D.G.; Pease, V. The Neoproterozoic Timanide Orogen of Eastern Baltica: Introduction. Geol. Soc. Lond. Mem. 2004, 30, 1–3. [Google Scholar] [CrossRef]
  60. Kuznetsov, N.; Soboleva, A.; Udoratina, O.; Hertseva, M.; Andreichev, V. Pre-Ordovician Tectonic Evolution and Volcano–Plutonic Associations of the Timanides and Northern Pre-Uralides, Northeast Part of the East European Craton. Gondwana Res. 2007, 12, 305–323. [Google Scholar] [CrossRef]
  61. Lorenz, H.; Gee, D.G.; Whitehouse, M.J. New Geochronological Data on Palaeozoic Igneous Activity and Deformation in the Severnaya Zemlya Archipelago, Russia, and Implications for the Development of the Eurasian Arctic Margin. Geol. Mag. 2007, 144, 105–125. [Google Scholar] [CrossRef]
  62. Prokopiev, A.V.; Ershova, V.B.; Sobolev, N.N.; Korago, E.; Petrov, E.; Khudoley, A.K. New Data on Geochemistry, Age and Geodynamic Settings of Felsic and Mafic Magmatism of the Northeastern Part of October Revolution Island (Severnaya Zemlya Archipelago). AGU Chapman Conference on «Large-Scale Volcanism in the Arctic: The Role of the Mantle and Tectonics». Selfoss, Iceland. 2019. Available online: https://higherlogicdownload.s3.amazonaws.com/AGU/181026d4-2440-440a-a114-8fcd2fc3ada8/UploadedImages/Chapmans/Arctic_Volcanism/ChapmanPresentedAbstracts_ArcticVolcanism.pdf (accessed on 1 October 2023).
  63. Kurapov, M.; Ershova, V.; Khudoley, A.; Makariev, A.; Makarieva, E. The First Evidence of Late Ordovician Magmatism of the October Revolution Island (Severnaya Zemlya Archipelago, Russian High Arctic): Geochronology, Geochemistry and Geodynamic Settings. NJG 2020, 100, 1–15. [Google Scholar] [CrossRef]
  64. Pogrebitskii, Y.E. (Ed.) Novaya Zemlya and Vaygach Island. In Geological Structure and Minerageny; VNIIOkeangeologiya: St. Petersburg, Russia, 2004; 174p. [Google Scholar]
  65. Ershova, V.; Anfinson, O.; Prokopiev, A.; Khudoley, A.; Stockli, D.; Faleide, J.I.; Gaina, C.; Malyshev, N. Detrital Zircon (U-Th)/He Ages from Paleozoic Strata of the Severnaya Zemlya Archipelago: Deciphering Multiple Episodes of Paleozoic Tectonic Evolution within the Russian High Arctic. J. Geodyn. 2018, 119, 210–220. [Google Scholar] [CrossRef]
  66. Brown, D.; Spadea, P.; Puchkov, V.; Alvarez-Marron, J.; Herrington, R.; Willner, A.P.; Hetzel, R.; Gorozhanina, Y.; Juhlin, C. Arc–Continent Collision in the Southern Urals. Earth-Sci. Rev. 2006, 79, 261–287. [Google Scholar] [CrossRef]
  67. Puchkov, V.N. The Evolution of the Uralian Orogen. Geol. Soc. Lond. Spec. Publ. 2009, 327, 161–195. [Google Scholar] [CrossRef]
  68. Zonenshain, L.P.; Kuz’min, M.I.; Natapov, L.M. Plate Tectonics of the USSR Territory: A Plate Tectonic Synthesis; Geodynamics Series; American Geophysical Union: Washington, DC, USA, 1990. [Google Scholar]
  69. Scott, R.A.; Howard, J.P.; Guo, L.; Schekoldin, R.; Pease, V. Offset and Curvature of the Novaya Zemlya Fold-and-Thrust Belt, Arctic Russia. In Geological Society, London, Petroleum Geology Conference Series; The Geological Society of London: London, UK, 2010; Volume 7, pp. 645–657. [Google Scholar] [CrossRef]
  70. Şengör, A.M.C.; Natal’In, B.A.; Burtman, V.S. Evolution of the Altaid Tectonic Collage and Palaeozoic Crustal Growth in Eurasia. Nature 1993, 364, 299–307. [Google Scholar] [CrossRef]
  71. Kurapov, M.; Ershova, V.; Khudoley, A.; Luchitskaya, M.; Makariev, A.; Makarieva, E.; Vishnevskaya, I. Late Palaeozoic Magmatism of Northern Taimyr: New Insights into the Tectonic Evolution of the Russian High Arctic. Int. Geol. Rev. 2021, 63, 1990–2012. [Google Scholar] [CrossRef]
  72. Vernikovsky, V.; Vernikovskaya, A.; Proskurnin, V.; Matushkin, N.; Proskurnina, M.; Kadilnikov, P.; Larionov, A.; Travin, A. Late Paleozoic–Early Mesozoic Granite Magmatism on the Arctic Margin of the Siberian Craton during the Kara-Siberia Oblique Collision and Plume Events. Minerals 2020, 10, 571. [Google Scholar] [CrossRef]
  73. Bracciali, L. Coupled Zircon-Rutile U-Pb Chronology: LA ICP-MS Dating, Geological Significance and Applications to Sediment Provenance in the Eastern Himalayan-Indo-Burman Region. Geosciences 2019, 9, 467. [Google Scholar] [CrossRef]
  74. Mueller, M.A.; Licht, A.; Möller, A.; Condit, C.B.; Fosdick, J.C.; Ocakoğlu, F.; Campbell, C. An Expanded Workflow for Detrital Rutile Provenance Studies: An Application from the Neotethys Orogen in Anatolia; Geochronological data analysis/statistics/modelling. EGUsphere 2023. preprint. [Google Scholar] [CrossRef]
  75. Rösel, D.; Boger, S.D.; Möller, A.; Gaitzsch, B.; Barth, M.; Oalmann, J.; Zack, T. Indo-Antarctic derived detritus on the northern margin of Gondwana. Evidence for continental-scale sediment transport. Terra Nova 2014, 26, 64–71. [Google Scholar] [CrossRef]
  76. Rösel, D.; Zack, T.; Boger, S.D. LA–ICP–MS U–Pb dating of detrital rutile and zircon from the Reynolds Range. A window into the Palaeoproterozoic tectonosedimentary evolution of the North Australian Craton. Precambrian Res. 2014, 255, 381–400. [Google Scholar] [CrossRef]
  77. Avigad, D.; Morag, N.; Abbo, A.; Gerdes, A. Detrital Rutile U-Pb Perspective on the Origin of the Great Cambro-Ordovician Sandstone of North Gondwana and Its Linkage to Orogeny. Gondwana Res. 2017, 51, 17–29. [Google Scholar] [CrossRef]
  78. Allen, C.M.; Campbell, I.H. Spot dating of detrital rutile by LA–Q–ICP–MS: A powerful provenance tool. In GSA Denver Annual Meeting, Abstract Paper; The Geological Society of America (GSA): Boulder, CO, USA, 2007; pp. 196–212. [Google Scholar]
  79. Small, D.; Parrish, R.R.; Austin, W.E.N.; Cawood, P.A.; Rinterknecht, V. Provenance of North Atlantic Ice-Rafted Debris during the Last Deglaciation—A New Application of U-Pb Rutile and Zircon Geochronology. Geology 2013, 41, 155–158. [Google Scholar] [CrossRef]
  80. Zack, T.; Von Eynatten, H.; Kronz, A. Rutile Geochemistry and Its Potential Use in Quantitative Provenance Studies. Sediment. Geol. 2004, 171, 37–58. [Google Scholar] [CrossRef]
  81. Khudoley, A.K.; Verzhbitsky, V.E.; Zastrozhnov, D.A.; O’Sullivan, P.; Ershova, V.B.; Proskurnin, V.F.; Tuchkova, M.I.; Rogov, M.A.; Kyser, T.K.; Malyshev, S.V.; et al. Late Paleozoic—Mesozoic Tectonic Evolution of the Eastern Taimyr-Severnaya Zemlya Fold and Thrust Belt and Adjoining Yenisey-Khatanga Depression. J. Geodyn. 2018, 119, 221–241. [Google Scholar] [CrossRef]
  82. Raimondo, T.; Hand, M.; Collins, W.J. Compressional Intracontinental Orogens: Ancient and Modern Perspectives. Earth-Sci. Rev. 2014, 130, 128–153. [Google Scholar] [CrossRef]
Figure 1. Simplified tectonic map of study region modified after [29,30,32]. The red rectangle shows the area of Figure 2.
Figure 1. Simplified tectonic map of study region modified after [29,30,32]. The red rectangle shows the area of Figure 2.
Geosciences 14 00041 g001
Figure 4. Photo of studied sections and position of dated samples (red stars): (A) Wilczek Land, 3 km to the north-west of Ganza Cape, general view of outcrop of the Vasiliev Fm. (Rhaetian) (left) and the position of 7-v15-9 sample within cross-bedded medium- to fine-grained sandstones, (B) Wilczek Land, 1 km to the north of Ganza Cape, general view of outcrop of the Ganza Fm. (right) and the position of 11-v15-1 sample collected from the sandy siltstone bed (left); (C) Hall Island, Tegethoff Cape, general view of outcrop of the Lower Jurassic Tegethoff Formation (right) and the position of 15AP30 sample within intercalating clays and fine-grained sandstones (left).
Figure 4. Photo of studied sections and position of dated samples (red stars): (A) Wilczek Land, 3 km to the north-west of Ganza Cape, general view of outcrop of the Vasiliev Fm. (Rhaetian) (left) and the position of 7-v15-9 sample within cross-bedded medium- to fine-grained sandstones, (B) Wilczek Land, 1 km to the north of Ganza Cape, general view of outcrop of the Ganza Fm. (right) and the position of 11-v15-1 sample collected from the sandy siltstone bed (left); (C) Hall Island, Tegethoff Cape, general view of outcrop of the Lower Jurassic Tegethoff Formation (right) and the position of 15AP30 sample within intercalating clays and fine-grained sandstones (left).
Geosciences 14 00041 g004
Figure 5. Cr–Nb source rock classification of rutile. Metamafic and metapelitic rutile are discriminated following [17].
Figure 5. Cr–Nb source rock classification of rutile. Metamafic and metapelitic rutile are discriminated following [17].
Geosciences 14 00041 g005
Figure 6. Pie charts showing the percentage of different metamorphic facies for the source rocks within the provenance area of detrital rutile, according to calculated formation temperatures. Metamorphic facies boundaries have been taken from the lower pressure end of facies diagram defined by [54].
Figure 6. Pie charts showing the percentage of different metamorphic facies for the source rocks within the provenance area of detrital rutile, according to calculated formation temperatures. Metamorphic facies boundaries have been taken from the lower pressure end of facies diagram defined by [54].
Geosciences 14 00041 g006
Figure 7. Kernel Density Estimation (KDE) plots depict the U-Pb detrital rutile data from the Upper Triassic (7-v15-9), Lower Jurassic (15AP30) and Upper Jurassic (11-v15-1) samples analyzed in this paper. The age distribution since the 700 Ma was zoomed in the right part of figure. n = number of detrital rutile analyses.
Figure 7. Kernel Density Estimation (KDE) plots depict the U-Pb detrital rutile data from the Upper Triassic (7-v15-9), Lower Jurassic (15AP30) and Upper Jurassic (11-v15-1) samples analyzed in this paper. The age distribution since the 700 Ma was zoomed in the right part of figure. n = number of detrital rutile analyses.
Geosciences 14 00041 g007
Figure 8. Kernel Density Estimation (KDE) plots depicting the U-Pb detrital rutile and zircon data, as well as ZHe zircon ages from the Upper Triassic Upper Triassic (7-v15-9), Lower Jurassic (15AP30) and Upper Jurassic (11-v15-1) samples analyzed in this paper. Detrital zircon data from [36].
Figure 8. Kernel Density Estimation (KDE) plots depicting the U-Pb detrital rutile and zircon data, as well as ZHe zircon ages from the Upper Triassic Upper Triassic (7-v15-9), Lower Jurassic (15AP30) and Upper Jurassic (11-v15-1) samples analyzed in this paper. Detrital zircon data from [36].
Geosciences 14 00041 g008
Figure 9. Cumulative proportion distributions against growth-deposition ages diagram, modified from [4]. Colored fields show the different depositional tectonic settings.
Figure 9. Cumulative proportion distributions against growth-deposition ages diagram, modified from [4]. Colored fields show the different depositional tectonic settings.
Geosciences 14 00041 g009
Table 1. Summary of locations and compositions of analyzed samples.
Table 1. Summary of locations and compositions of analyzed samples.
Sample
Number
LatitudeLongitudeRock Composition
7-v15-9N 80°25.563′E 059°39.570′Medium-grained polymictic sandstone
11-v15-1N 80°24.695′E 059°42.964′Fine-grained arkosic arenite sandstone
15AP30N 80°06′13.9″E 057°53′02.1″Medium-grained polymictic sandstone
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ershova, V.; Prokopiev, A.; Stockli, D. Provenance of Detrital Rutiles from the Triassic–Jurassic Sandstones in Franz Josef Land (Barents Sea Region, Russian High Arctic): U-Pb Ages and Trace Element Geochemistry. Geosciences 2024, 14, 41. https://doi.org/10.3390/geosciences14020041

AMA Style

Ershova V, Prokopiev A, Stockli D. Provenance of Detrital Rutiles from the Triassic–Jurassic Sandstones in Franz Josef Land (Barents Sea Region, Russian High Arctic): U-Pb Ages and Trace Element Geochemistry. Geosciences. 2024; 14(2):41. https://doi.org/10.3390/geosciences14020041

Chicago/Turabian Style

Ershova, Victoria, Andrei Prokopiev, and Daniel Stockli. 2024. "Provenance of Detrital Rutiles from the Triassic–Jurassic Sandstones in Franz Josef Land (Barents Sea Region, Russian High Arctic): U-Pb Ages and Trace Element Geochemistry" Geosciences 14, no. 2: 41. https://doi.org/10.3390/geosciences14020041

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop