Next Article in Journal
Phase Guard: A False Positive Filter for Automatic Rietveld Quantitative Phase Analysis Based on Counting Statistics in HighScore Plus
Previous Article in Journal
Study on the Drying Characteristics of Moist Fine Lignite in a Dense Gas–Solid Separation Fluidized Bed
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sedimentary Environment, Tectonic Setting, and Paleogeographic Reconstruction of the Late Jurassic Weimei Formation in Dingri, Southern Tibet

1
Research Center of Applied Geology of China Geological Survey, Chengdu 610036, China
2
Institute of Sedimentary Geology, Chengdu University of Technology, Chengdu 610059, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(10), 1040; https://doi.org/10.3390/min15101040
Submission received: 24 August 2025 / Revised: 26 September 2025 / Accepted: 29 September 2025 / Published: 30 September 2025
(This article belongs to the Section Mineral Geochemistry and Geochronology)

Abstract

The Weimei Formation, the most complete Upper Jurassic sedimentary sequence in the Tethyan Himalaya, is crucial for understanding the tectono-sedimentary evolution of the northern Indian margin. However, its depositional environment remains debated, with conflicting shallow- and deep-water interpretations. This study integrates sedimentary facies, petrography, zircon geochronology, and geochemical analyses to constrain the provenance, depositional environment, and tectonic setting of the Weimei Formation. The results reveal that the sedimentary system primarily consists of shoreface, delta, and shelf facies, with locally developed slope-incised valleys. Detrital zircon ages are concentrated at ~468 Ma and ~964 Ma, indicating a provenance mainly derived from the Indian continent. Geochemical characteristics, such as high SiO2, low Na2O–CaO–TiO2 contents, right-leaning REE patterns, and significant negative Eu anomalies, suggest the derivation of sediments from felsic upper crustal recycling within a passive continental margin. Stratigraphic comparison between southern and northern Tethyan Himalayan sub-zones reveals a paleogeographic “uplift–depression” pattern, characterized by the coexistence of shoreface–shelf deposits and slope-incised valleys. This study provides key evidence for reconstructing the Late Jurassic paleogeography of the northern Indian margin and the tectonic evolution of the Neo-Tethys Ocean.

1. Introduction

The sedimentary successions of the Tethyan Himalaya preserve important records of the northern Indian continental margin and register a series of major geological events, including the development of passive continental margin basins, the breakup of East Gondwana, the closure of the Tethys Ocean, and the India–Eurasia continental collision [1]. The Late Jurassic represents a critical period in the evolution of the Tethys Ocean, during which significant events such as the opening of the Neo-Tethys Ocean [2,3,4,5], the breakup of East Gondwana [6,7,8,9], and the activity of the Kerguelen mantle plume [10,11,12,13] occurred. Therefore, the study of Jurassic sedimentary successions is crucial for understanding the evolutionary processes of the Tethys Ocean. In southern Tibet, Jurassic marine strata are extensively exposed within the Tethyan Himalaya, among which the Upper Jurassic Weimei Formation is particularly significant, as it contains key records of the early evolution of the Neo-Tethys Ocean.
Previous studies on the Upper Jurassic Weimei Formation have explored its paleontology, sedimentology, stratigraphy, and paleomagnetism [14,15,16,17,18,19,20,21,22]. The Weimei Formation is mainly composed of thick-bedded quartz sandstone interbedded with mudstones, displaying shallow-marine depositional characteristics, and yielding abundant fossils, including planktonic ammonites, belemnites, benthic bivalves, and gastropods, accompanied by a certain number of deep-water radiolarians [23]. The quartz sandstones are mineralogically mature, with quartz contents exceeding 95%, but are poorly sorted and contain gravel-sized clasts [23]. Consequently, the depositional environment of the Weimei Formation has long been debated. Most researchers, based on faunal assemblages, sedimentary structures, and petrological features, interpret it as having formed in high-energy shoreface to shallow-marine environments [14,24,25,26]. In contrast, another viewpoint emphasizes abrupt bedding contacts, poorly developed bedding, low sandstone maturity, the presence of mud clasts, deep-water radiolarians, and planktonic cephalopods, interpreting the succession as sandy debris flow deposits in semi-deep to deep marine slope setting [23,27,28]. Additionally, some scholars, based on the presence of unconformable conglomerate horizons and turbiditic features, have proposed that the Weimei Formation developed shelf-incised valleys, forming a depositional system where shoreface–shallow-marine deposits coexisted with shelf-incised valley fills [26,29].
Given these controversies, this study focuses on the Upper Jurassic Weimei Formation in the Dingri area of the Tethyan Himalaya. Based on detailed field observations combined with petrographic, geochemical, and detrital zircon U–Pb geochronological analyses, we systematically investigate its sedimentary characteristics, tectonic setting, and provenance. On this basis, we conduct both lateral and vertical stratigraphic correlations of the Late Jurassic strata across the Tethyan Himalaya to reconstruct the sedimentary framework and paleogeography, and further discuss the tectonic evolution of the northern Indian continental margin.

2. Geological Background

The Himalayan Orogen is located at the junction between the Indian and Eurasian continents (Figure 1a) and forms a key segment of the northern Indian continental margin [30,31]. From south to north, it is subdivided into four major tectonic units: the Sub-Himalaya (SHS), Lesser Himalaya (LHS), Greater Himalaya (GHS), and Tethyan Himalaya (THS). The Tethyan Himalaya, situated between the Indus–Yarlung Tsangpo suture zone (IYSZ) and the South Tibet detachment system (STDS) [31,32], comprises Proterozoic to Eocene marine successions dominated by clastic and carbonate rocks [33,34,35]. These successions exceed 10 km in thickness, contain abundant fossil assemblages, and locally include Paleozoic and Mesozoic volcanic rocks [1,18].
The Tethyan Himalaya can be further subdivided into southern and northern subzones by the Dingri–Gamba fault (DGF) (Figure 1b). The southern subzone is dominated by low-grade metamorphosed continental margin deposits, with lithologies mainly comprising Paleozoic–Mesozoic carbonate and clastic rocks deposited in shoreface to shelf environments [34,38,39]. In contrast, the northern subzone is characterized by continental slope deposits, with lithologies primarily including sandstone, siliceous mudstone, shale, and limestone [34,38].
The Upper Jurassic Weimei Formation is exposed in the northern subzone of the Tethyan Himalaya and is widely distributed across the Saga–Dingri–Langkazi–Longzi region. First defined by Wang et al. (1980) [40], it is consists of thick-bedded sandstone interbedded with shale, with thicknesses ranging from several hundred meters to over one kilometer. In the eastern Gyangze–Longzi region, the lithology is dominated by dark-gray shale and silty shale interbedded with quartz sandstone, commonly associated with ferruginous or siliceous nodules. In contrast, in the central and western Dingri–Gyirong region, the succession includes interbeds of limestone and marlstone. The Weimei Formation yields abundant Late Jurassic ammonites, belemnites, bivalves, and gastropods, and is assigned to the Late Jurassic Tithonian Stage [20].
The Dingri area is located in the central part of the northern subzone of the Tethyan Himalaya (Figure 1c). In this region, the Upper Jurassic Weimei Formation reaches a thickness of approximately 920 m, consisting of a clastic–carbonate succession with minor tuffaceous sandstone and tuff interbeds. It is conformably overlain by the Lower Cretaceous Jiabula Formation. Based on lithological characteristics, the Weimei Formation in this area can be subdivided into three members (Figure 2a). The lower member is mainly composed of argillaceous shale and siltstone interbedded with quartz sandstone, lithic quartz sandstone, coarse-grained sandstone, and conglomerate, with quartz sandstone as the characteristic lithology (Figure 2b). The middle member is dominated by siltstone and silty shale interbedded with lithic quartz sandstone, containing abundant sandy nodules, with lithic sandstone as the characteristic lithology. (Figure 2c). The upper member comprises silty shale, calcareous siltstone, and micritic limestone, interbedded with sandy limestone and calcareous lithic sandstone, characterized by abundant limestone interbeds (Figure 2d). The Weimei Formation in this area also yields abundant Late Jurassic belemnites, ammonites, bivalves, and gastropods.

3. Analytical Methods

The sampling site is located at Yundong Village, Kema Township, Dingri County. To ensure the freshness and integrity of the samples, veins and strongly weathered outcrops were avoided during sampling. One quartz sandstone sample from the Upper Jurassic Weimei Formation (Sample ID: WMDN; coordinates: 86°41′28″ E, 28°49′10″ N) was collected for detrital zircon U–Pb dating. Additionally, five quartz sandstone samples were collected from the lower member and seven lithic sandstone samples were collected from the middle member of the Weimei Formation for whole-rock geochemical analyses.
Detrital zircon separation was conducted at the Hebei Institute of Regional Geological and Mineral Survey. After crushing to 200 mesh, zircons were extracted using flotation and electromagnetic separation methods. Under a binocular microscope, euhedral and highly transparent zircon grains were handpicked, mounted in epoxy resin, and then ground and polished. Reflected light, transmitted light, and cathodoluminescence (CL) imaging were used to select euhedral zircon grains with minimal cracks and inclusions, and with clear, uniform oscillatory zoning for analysis. Analytical work was carried out at the Laser Ablation Multi-Collector Inductively Coupled Plasma Mass Spectrometer (LA–MC–ICP–MS) Laboratory, Institute of Mineral Resources, Chinese Academy of Geological Sciences, using a laser spot size of 30 μm. To ensure analytical accuracy, two GJ-1 standards and one Plesovice zircon standard were analyzed for calibration after every five analysis points, following the experimental procedures described by Hou et al. (2007) [41]. Data pre-processing was performed using the ICP–MS DataCal v4.3 program [42], and concordia diagrams and age probability histograms were generated using Isoplot 3.0.
For geochemical analyses, twelve sandstone samples were washed, crushed, repeatedly cleaned, dried, and crushed to 200 mesh. Major element analyses were conducted at the Analytical and Testing Center of the 605 Geological Exploration Team, Sichuan Bureau of Metallurgy and Geology, using an X-ray fluorescence spectrometer (XPF–1500), with analytical precision better than 2%–3%. Trace and rare earth elements were determined after acid digestion of the samples using an inductively coupled plasma mass spectrometer (ICP–MS, Element II), with analytical errors generally within 5%.

4. Results

4.1. Zircon U–Pb Geochronology

The U–Pb dating results of detrital zircons from the Upper Jurassic quartz sandstones of the Weimei Formation are presented in Table A1. Most zircons are transparent, colorless or pale yellow, with grain sizes of ~50–200 μm and length/width ratios of 1:1–3:1. Representative zircon CL images are shown in Figure 3a. The majority of the grains display well-developed oscillatory zoning with tabular or fan-shaped structures. Their Th/U ratios are generally >0.1, indicating a magmatic genesis [43,44]. A few grains show narrow metamorphic overgrowth rims. Overall, zircons are dominantly subrounded to subangular, suggesting relatively short transport distances and a source from nearby areas.
In this study, zircons were selected based on their well-preserved crystal shape and high transparency, with a predominance of subrounded to subangular grains, indicating that the zircon particles have been less affected by metamictization [43]. For micro-domain selection, we focused on regions with clear, uniform, and clean oscillatory zoning, with most of the analytical spots located at the outermost part (Figure 3a), which typically represent younger crystallization ages [44]. This selection criterion may result in an overrepresentation of younger zircon grains, leading to an age distribution skewed towards younger ages.
A total of 106 zircon grains were analyzed, with concordance values exceeding 90%. Analytical points plot along or near the concordia line (Figure 3b). For zircons younger than 1000 Ma, 206Pb/238U ages were determined, whereas for those older than 1000 Ma, 207Pb/206Pb ages were used [45]. The zircon ages range from 306 Ma to 3435 Ma, with major concentrations between 420 and 1300 Ma. The detrital zircon age spectrum reveals two dominant age populations (Figure 3c): the first cluster falls between 460 and 519 Ma (n = 31), with a peak at ~468 Ma; the second cluster ranges from 833 to 1194 Ma (n = 45), with a peak at ~964 Ma.

4.2. Geochemistry

Whole-rock geochemical results of the Upper Jurassic sandstones from the Weimei Formation are summarized in Table 1. Quartz sandstones show high SiO2 content (95.2%–96.8%, avg. 96.2%), with relatively low contents of Al2O3 (1.12%–1.98%, avg. 1.58%), TFe2O3 (0.57%–1.60%, avg. 1.01%), MgO (0.04%–0.30%, avg. 0.12%), K2O (0.07%–0.31%, avg. 0.14%), Na2O (0.03%–0.13%, avg. 0.06%), and CaO (0.04%–0.06%, avg. 0.05%). In contrast, lithic sandstones exhibit lower SiO2 content (62.9%–70.3%, avg. 67.0%), but much higher contents of Al2O3 (4.10%–17.9%, avg. 10.1%), TFe2O3 (3.34%–7.82%, avg. 5.66%), MgO (2.50%–7.50%, avg. 4.48%), K2O (2.30%–3.31%, avg. 2.70%), Na2O (0.34%–0.73%, avg. 0.49%), and CaO (0.17%–18.4%, avg. 7.92%). The TFe2O3 + MgO (8.55%–11.0%) and K2O/Na2O (3.53–7.54) values in lithic sandstones are significantly higher than those in quartz sandstones (0.61%–1.90% and 1.63–2.86, respectively), indicating enrichment in ferromagnesian components and possibly the presence of K-feldspar or K-rich minerals.
The Al2O3/(CaO + K2O) and SiO2/Al2O3 ratios are commonly used to reflect the proportions of stable versus unstable components [46,47]. In quartz sandstones, these ratios (5.16–16.5 and 48.1–86.0) are significantly higher than in lithic sandstones (0.21–5.39 and 3.66–15.7), indicating that quartz sandstones are dominated by siliceous minerals and have higher compositional maturity, while lithic sandstones are relatively enriched in clay minerals, such as illite and kaolinite.
In the chondrite-normalized rare earth elements (REE) patterns (Figure 4a), all samples exhibit rightward-inclined trends, characterized by moderate light REE (LREE) enrichment and relatively flat heavy REE (HREE) distributions (LREE/HREE = 6.82–16.8). The samples show a pronounced negative Eu anomaly (δEu = 0.47–0.90, avg. 0.61), while Ce anomalies are negligible (δCe = 0.96–1.41, avg. 1.10). Quartz sandstones display relatively low and stable ΣREE values (53.2–58.1 ppm, avg. 56.1 ppm), whereas lithic sandstones show much larger variations (50.6–255 ppm, avg. 143 ppm). Notably, samples KMFX1, KMFX2, and KMFX3 exhibit higher ΣREE contents (209–255 ppm), likely due to the higher clay mineral contents, as clays strongly adsorb REEs [48,49]. In the Upper Continental Crust (UCC)-normalized REE patterns (Figure 4b), all samples show relatively flat distributions. However, lithic sandstones have a broader distribution range than quartz sandstones, suggesting more rapid deposition and less sedimentary recycling.
The Rb, Sr, and Ba contents of the sandstones range from 3.13 to 124 ppm, 12.7–419 ppm, and 25.9–611 ppm, respectively, all generally lower than the average UCC values [42]. The Th/U ratios range from 3.5 to 19.1 (avg. 13.2), significantly higher than the UCC (3.89) [52] and the Post-Archean Australian Shale (PASS) (4.7) [51]. Some samples (KMFX1, KMFX2, KMFX7) exhibit relatively high concentrations of Cr, Co, Ni, V, Th, and U, suggesting a possible contribution from ultramafic components in the source area. Other samples (KMFX4, KMFX5, KMFX6) show markedly lower Rb, Sr, and Ba contents, likely due to abundant carbonate components [53], consistent with their elevated CaO contents (16.3%–18.4%).

4.3. Lithofacies and Sedimentary Facies

4.3.1. Lithofacies Characteristics

Quartz sandstone is predominantly exposed in the lower member of the Weimei Formation. It is grayish white, medium- to fine-grained, and medium- to thick-bedded or massive (Figure 5a). The rock shows grain-supported textures with contact cementation. Detrital grains range from 0.25 to 0.5 mm, mostly subrounded to rounded, with relatively well sorted and high roundness, reflecting a high degree of compositional maturity and textural maturity. Detrital components are dominated by monocrystalline quartz (ca. 90%–95%), some of which exhibit secondary enlargements (Figure 5b). According to McBride, (1989) [54], teardrop-shaped secondary enlargements commonly form in shoreface or semi-exposed environments. The matrix (ca. 5%–10%) mainly consists of siliceous cement, with minor heavy minerals including tourmaline and zircon.
Lithic quartz sandstone is primarily developed in the middle member of the Weimei Formation. It occurs as medium-bedded or lenticular bodies, gray to yellowish gray, and medium- to fine-grained (Figure 5c). The rock consists of detrital grains (ca. 85%–90%) and matrix (ca. 10%–15%), showing a grain-supported texture with basal cementation. Detrital grains are subangular to subrounded with moderately sorted. The composition includes quartz, feldspar, lithic fragments, and mica (Figure 5d). Lithic fragments (ca. 20%) include volcanic rocks (e.g., andesite and mafic rocks), as well as sandstone, and mudstone. Heavy minerals such as tourmaline and zircon are occasionally present. The matrix mainly consists of siliceous–argillaceous cement.
Micritic limestone is mainly developed in the upper member of the Weimei Formation and is commonly interbedded with thin shale and mudstone. The rock is grayish brown to dark gray, 5–20 cm thick, medium- to thin-bedded, and exhibits a micritic texture (Figure 5e). Mineral composition is dominated by calcite (ca. 85%–90%), with clay (ca. 5%–10%) and minor quartz and muscovite (ca. 5%) (Figure 5f). Some samples contain up to 25%–30% detrital grains. Calcite crystals are very fine (0.01–0.03 mm), mostly irregular or compact aggregates, with occasional bands of sparry calcite and localized recrystallization forming microsparitic textures. Clay is unevenly distributed, with scattered quartz grains and platy muscovite.

4.3.2. Sedimentary Facies Analysis

Systematic fieldwork and integrated analyses of lithologic associations, sedimentary structures, and fossil assemblages reveal four depositional facies in the Weimei Formation: shoreface, braided delta, shelf, and deeply incised valley. These facies record a shallow-marine depositional system, with gravity-flow deposits locally developed in deeply incised valleys.
Shoreface facies: This facies is mainly developed in the lower member and is characterized by medium- to thick-bedded quartz conglomerates and medium- to fine-grained lithic quartz sandstone, often associated with scour surfaces. Gravels are mainly quartzose, with minor sandstone clasts, ranging from 1 to 3 cm in diameter, and are mostly subrounded. Quartz constitutes ca. 75%–95% of the detrital fraction, with a grain-supported texture indicating high compositional and textural maturity. Sedimentary structures include parallel bedding and tabular cross-bedding (Figure 6a). These features suggest a high-energy shoreface environment dominated by wave reworking, where repeated winnowing and transport generated mature quartz sand and quartzose gravels.
Braided delta facies: This facies is mainly developed in the middle member and consists of conglomerate, lithic sandstone, siltstone, and mudstone, displaying typical delta-front characteristics. Four types of depositional sequences are distinguished: (1) conglomerate–lithic sandstone–siltstone–mudstone assemblages with scour surface, large-scale trough cross-bedding (Figure 6b), and graded bedding, interpreted as subaqueous distributary channel deposits; (2) siltstone–mudstone assemblages with horizontal bedding, wavy bedding, and flaser bedding, representing subaqueous levee deposits; (3) argillaceous shale–siltstone assemblages with horizontal bedding, small-scale cross-bedding, lenticular bedding, bioclastic fragments, plant fragments (Figure 6c), and trace fossils, indicating subaqueous interdistributary bay deposits; and (4) lens-shaped, thick to massive lithic sandstones bodies with parallel bedding and wedge cross-bedding (Figure 6d), corresponding to distal sandbar deposits.
Shelf facies: This facies is mainly developed in the upper member and exhibits a clear vertical zonation. The lower part consists of rhythmically interbedded siltstone, silty shale, and micritic limestone, with horizontal bedding, wavy bedding, parallel bedding, and graded bedding, suggesting a relatively high-energy inner shelf environment. Locally interbedded calcarenite (Figure 6e) is interpreted as mixed platform deposits formed during a regressive phase. The upper part consists mainly of siltstone and shale with thin or lenticular limestone beds showing by horizontal bedding, indicative of a lower-energy outer shelf setting.
Incised valley facies: This facies is sporadically developed in the middle member, characterized by abrupt contacts and pronounced lateral erosion. Lithologic assemblages are complex, mainly including pebble-bearing sandstone, feldspathic sandstone, and greywacke, forming a depositional system of channel-lag conglomerates and channel-fill lithic quartz sandstones. Conglomerates are matrix-supported, with angular pebbles (>63% by volume), averaging 15–24 mm in size (up to 35 mm). Sandstones exhibit scour structures, mud rip-up clasts (Figure 6f), load casts (Figure 6g), and flute casts (Figure 6h). Graded bedding and flame structures are common in the pebbly sandstone (Figure 6i). These features indicate deposition by high-energy, rapid erosion–transport–accumulation processes, attributed to high-density gravity flows (e.g., high-density turbidity flows or sandy debris flows). Their spatial association with shoreface deposits suggests erosion of slope incised valleys during regressive phases, with these valleys extending landward into the shelf–shoreface environments, leading to the co-existence of deeply incised valleys and shallow-marine deposits in both space and time.

5. Discussion

5.1. Provenance Analysis

Based on detrital zircon age spectra from various tectonic units in southern Tibet (Figure 7), the Lhasa Terrane is characterized by prominent age peaks at 200–350 Ma, 450–530 Ma, and 1150–1250 Ma. These age distributions closely resemble those of detrital zircons from Western Australia [55,56,57], suggesting a tectonic affinity in terms of sedimentary provenance. In contrast, the Western Qiangtang Terrane, Zhongba Terrane, Tethyan Himalaya, and Higher Himalaya display highly consistent age peak distributions, which correlate well with those of East India, with major peaks at 450–530 Ma, 900–1050 Ma, and ~2500 Ma [35,58,59,60,61]. Together, these regions constitute the Western Qiangtang–East India–Himalaya tectonic system. The detrital zircon age peaks of the Weimei Formation sandstones in the Dingri area (~468 Ma and ~964 Ma) are consistent with this system [62,63], indicating that the sediments were mainly derived from the Indian continent.
The first peak age (~468 Ma) observed in this study is slightly younger but still close to the peak ages (500–650 Ma) recorded in the Lesser Himalaya, Higher Himalaya, Tethyan Himalaya, Lhasa Terrane, Western Australia, and the Indian continent, which are generally attributed to the Pan-African orogenic event [35,57,71]. However, as discussed in Section 4.1 (“Zircon U–Pb Geochronology”), the selection of zircons and micro-domains was biased towards younger zircon grains, resulting in an apparent 468 Ma peak, which is notably younger than the typical 500–550 Ma peak recorded in the Himalayas. Furthermore, zircon ages in the 450–500 Ma range are widely present in Mesozoic strata within the Tethyan Himalaya. For example, a significant presence of 450–500 Ma zircon ages in the Triassic Qulonggongba and Derirong Formations in the Dingri area was reported [72], while numerous 438–500 Ma zircon ages in the Lower Cretaceous strata of the Longzi area were documented [73]. Zircons within this age range are also prevalent in other Himalayan, eastern Indian, and western Australian strata. Studies indicate that the Pan-African orogeny in the East Gondwana region, specifically the Bhimphedian orogeny, occurred slightly later (460–530 Ma) compared to the East African, Kuunga–Pinjarra (Antarctica), and Ross–Delamerian orogenies (500–650 Ma) [74]. It is inferred that the ~468 Ma peak age in the Weimei Formation of the Dingri area is primarily derived from the Bhimphedian orogeny.
The second peak age (~964 Ma) is broadly distributed in the Lesser Himalaya, Higher Himalaya, Tethyan Himalaya, and Indian continent, and is considered a typical detrital zircon signature of cratonic and passive continental margin deposits [75]. This indicates that during the Late Jurassic, the Himalayan Orogen exhibited a relatively uniform provenance, primarily sourced from the Indian continent. Furthermore, the absence of the ~1170 Ma peak age in the Tethyan Himalaya, which is shared by the Lhasa Terrane and Western Australia [35,57], suggests that the Lhasa Terrane had already rifted away from the Himalayan margin during Late Jurassic.

5.2. Source Composition and Tectonic Setting

The chemical composition of sandstones is primarily controlled by the lithology of their source rocks, which record the original geochemical signatures of the sediments and provides valuable constraints on provenance characteristics. Roser and Korsch, (1988) [76] established discriminant functions F1–F2 and F3–F4 based on major elements (TiO2, Al2O3, TFe2O3, MgO, CaO, Na2O, K2O) and their ratios (TiO2/Al2O3, TFe2O3/Al2O3, MgO/Al2O3, Na2O/Al2O3, K2O/Al2O3), enabling the effective classification of four sedimentary provenance types: mafic igneous, intermediate igneous, felsic igneous, and recycled quartzose sources. The Weimei Formation sandstones in the Dingri area consistently plot within the quartzose sedimentary provenance field on both the F1–F2 (Figure 8a) and F3–F4 (Figure 8b) diagrams. Similarly, on the Co/Th–La/Sc (Figure 8c) and La/Th–Hf (Figure 8d) diagrams, the samples fall within the provenance field characterized by felsic components from passive continental margins.
The geochemical signatures of clastic sedimentary rocks not only reflect the composition of source rocks but are also closely related to tectonic settings. As a result, these signatures are widely used as discriminants for depositional environments [46,48]. In terms of major elements, sandstones deposited in different tectonic settings show significant variations such as TFe2O3 + MgO, TiO2, Al2O3/SiO2, K2O/Na2O, and Al2O3/(Na2O + CaO). For example, sandstones from active continental margins typically contain lower TFe2O3 + MgO (2.00%–5.00%) and TiO2 (0.25%–0.45%), with K2O/Na2O ratios close to 1. In contrast, sandstones deposited along a trend from oceanic island arcs to passive continental margins display decreasing TFe2O3 + MgO, TiO2, and Al2O3/SiO2 values, while K2O/Na2O and Al2O3/(Na2O + CaO) ratios progressively increase. Passive continental margin sandstones are generally enriched in SiO2 and depleted in Na2O, CaO, and TiO2 [46]. The sandstones of the Weimei Formation in the Dingri area are characterized by high SiO2 contents (avg. 79.1%) and low Na2O (avg. 0.31%), CaO (avg. 4.64%), and TiO2 (avg. 0.51%) contents, with relatively high TFe2O3 + MgO (avg. 6.39%) and K2O/Na2O ratios (avg. 4.38). These characteristics distinguish them from those in active continental margin or island arc settings but are consistent with passive continental margin deposition. This inference is further supported by the K2O/Na2O–SiO2 discrimination diagram (Figure 9a).
Compared to major elements, immobile trace elements (e.g., La, Th, Y, Zr, Ti, Co, Ni) are less affected by alteration during weathering, transport, and diagenesis, and therefore provide more reliable indicators of provenance and tectonic setting [79]. In the Ti/Zr–La/Sc diagram (Figure 9b), most Weimei Formation sandstone samples fall within the passive continental margin field, consistent with results from the La–Th–Sc, Th–Sc–Zr/10, and Th–Co–Zr/10 diagrams (Figure 9c). A few lithic sandstone samples tend toward the active continental margin field (Figure 9), which may reflect the presence of volcanic fragments (e.g., andesite and mafic rocks) (Figure 5d). Additionally, tuffaceous sandstones and tuff interlayers have been identified in the Dingri area, while basaltic interbeds are developed in the Zhongba area [74]. These volcanic fragments may be associated with contemporaneous volcanism in the Tethyan Himalaya.

5.3. Lithofacies Paleogeography

The basin infill process was influenced by sediment supply, basement subsidence, and sea-level fluctuations. Sedimentary records not only reflect provenance and depositional environments but also preserve information on tectonic evolution. Therefore, basin analysis is critical for deciphering its evolutionary history and reconstructing regional tectonic evolution [80,81].
In the Tethyan Himalaya, the Late Jurassic strata primarily consist of shoreface–shelf facies clastic rocks and carbonates, with thicknesses varying significantly from several hundred meters to over 2000 m [82]. The Tethyan Himalaya, divided by the Dingri–Gamba fault, is subdivided into southern and northern subzones, where the Menkadun and Weimei Formations were deposited during the Late Jurassic, respectively. Since the Mesozoic, the multiphase activity of the Dingri–Gamba fault has exerted significant basin-controlling effects on the sedimentary evolution of these subzones. By correlating sedimentary sequences and facies, the Late Jurassic lithofacies paleogeography can be reconstructed, further evaluating the tectonic influence of the Dingri–Gamba fault during this period.
In the northern subzone, lateral correlation of the Weimei Formation reveals minimal lithological variation, with the strata predominantly composed of clastics (sandstone, siltstone, mudstone) and carbonates (micrite, mudstone limestone), with abundant shallow-marine fossils, overall representing nearshore to shallow-marine deposits. However, the thickness of the Weimei Formation varies significantly across regions (Figure 10b): 200–300 m in Cuomei and Gyangze, 500–700 m in Saga and Gyirong [83,84], and over 1000 m in Langkazi and Dingri [83]. Laterally, stratigraphic thickness exhibits an “uplift–depression” pattern (Figure 10a), with shallow-marine assemblages dominant in the uplifted zones (Cuomei, Gyangze, Gyirong), while incised valley deposits are developed in the depressed zones (Dingri, Langkazi). This unique paleogeographic pattern reasonably explains the previous controversy regarding the depositional environment of the Weimei Formation—both shallow-marine high-energy deposits and localized deep-water gravity flow deposits can coexist within the same region.
In the southern subzone, vertical comparisons show that the Menkadun Formation and the Weimei Formation exhibit similarities in lithological assemblages, filling successions, stratigraphic thickness, and depositional environments (Figure 10c). The Menkadun Formation also comprises thick-bedded quartz sandstones and carbonates, containing a variety of shallow-marine fossils, with deposition mainly occurring in shoreface environments and locally extending into shallow-shelf settings. These comparisons indicate that the Dingri–Gamba fault did not significant control on basin configuration during the Late Jurassic.
In summary, integrated lateral and vertical comparisons reveal that the northern margin of the Indian continent was characterized by shallow-marine environments during the Late Jurassic, with several south-to-north trending incised valleys locally extending into the shallow-marine zone. As a result, a depositional mosaic developed, while the Dingri–Gamba fault did not exhibit significant basin-controlling effects.

5.4. Tectonic Evolution

The Late Jurassic was a critical stage in the tectonic evolution of the Himalayas. During this period, the Indian continent progressively rifted away from Africa, Australia, and Antarctica, forming an independent continent that drifted rapidly northward and ultimately collided with Eurasia in the Eocene, leading to the closure of the Neo-Tethys Ocean [15,85]. Paleomagnetic and detrital zircon U–Pb studies suggest that, during the Late Jurassic, the Tethyan Himalaya was situated at the northern margin of the Indian continent, with a paleolatitude of ~22.7° S [57,86,87,88,89] (Figure 11a). Against this tectonic backdrop, sedimentary systems and paleogeographic patterns in southern Tibet experienced significant changes. Stratigraphic records document the key evolutionary stages, from Pangea breakup, northward drift of India, to the expansion and subduction of the Neo-Tethys Ocean [90,91]. Therefore, the Upper Jurassic sedimentary record provides critical evidence for reconstructing both the tectonic evolution of the Neo-Tethys Ocean and the paleogeography of the northern Indian margin.
Integrated sedimentological, zircon U–Pb geochronological, and geochemical data from the Weimei Formation provide a clear record of the coupled tectonic–sedimentary processes along the northern Indian margin during the Late Jurassic. Key constraints are as follows: (1) Provenance constraints: Detrital zircon age spectra exhibit major peaks at ~468 Ma and ~964 Ma, but lack the ~1170 Ma peak. This distribution is consistent with the western Qiangtang–eastern India–Himalaya tectonic system, and contrasts with the Lhasa terrane and western Australia [35], indicating the provenance originated from the Indian craton and its passive continental margin, rather than the Lhasa terrane [22]. (2) Sedimentological constraints: The Weimei Formation is dominated by shoreface–delta–shelf facies, with locally developed deeply incised valleys [18]. The overall sedimentary pattern is characterized by the coexistence of high-energy shoreface sands and low-energy shelf mudstones, with incised valleys formed and filled by sandy gravel sediments during regression. (3) Geochemical constraints: The sandstones generally show high SiO2 and low Na2O–CaO–TiO2 contents, elevated K2O/Na2O ratios, right-leaning REE patterns with flat HREE, and significant negative Eu anomalies. High Th/U ratios further indicate felsic upper crustal and recycled sediment sources in a passive continental margin setting. Minor deviations toward “active margin/island arc” signatures reflect input of contemporaneous volcanic material.
During the Late Jurassic, sedimentation in the Tethyan Himalaya was mainly driven by regional thermal subsidence and high-frequency relative sea-level fluctuations. Along the Gyirong–Dingri–Gyangze–Cuomei transect, rhythmic alternations of shoreface–delta–shelf sandstones, shales, and carbonates developed, indicating wave-dominated shorelines under moderate to low sediment supply and accommodation space. In localized areas, such as Dingri and Langkazi, inherited basement differential subsidence, combined with relative sea-level falls, led to the formation of deeply incised valleys filled with sandy gravel deposits. Stratigraphic thicknesses exhibit an “uplift–depression” pattern: thinner, high maturity successions in Cuomei, Gyangze, and Gyirong; thicker, low maturity successions in Dingri and Langkazi, with distinct grain-size and geochemical differences. Notably, the Dingri–Gamba fault did not exert significant basin-controlling effects during this period. Overall, the Weimei Formation records the spatial coexistence of shoreface–shelf marine and incised valley–gravity-flow deposits, revealing heterogeneous sedimentation within a passive margin setting [92].
Regional evidence suggests that the breakup of East Gondwana may have initiated in the Late Jurassic. Zhu et al. (2008) [93] identified 147 Ma OIB-type gabbro–dolerite in the Cona area, associated with the activity of the Kerguelen mantle plume. Combined with sedimentological and detrital zircon evidence from the Weimei Formation, it is evident that the Tethyan Himalaya was not an isolated drifting microcontinent during the Late Jurassic [8,94], but rather an integral part of the northern Indian margin. Its tectono-sedimentary evolution was directly controlled by the northward drift of the Indian continent and the development of the Neo-Tethys Ocean. The Weimei Formation provides a long-term record of passive margin shoreface–shelf marine systems dominated by thermal subsidence, with stable provenance from the Indian craton. Local incised valley and gravity-flow deposits reflect secondary effects of sea-level falls and basement differential subsidence (Figure 11b). Thus, during the Late Jurassic, the Tethyan Himalaya retained the essential attributes of a passive continental margin, while also recording localized depositional heterogeneity, offering key evidence for understanding the evolution of the Neo-Tethys Ocean and the paleogeography of the northern Indian margin.

6. Conclusions

(1)
The sedimentary system of the Weimei Formation is dominated by shoreface–delta–shelf facies, characterized by rhythmic interbedding of sandstone, shale, and carbonate rocks. Locally developed deeply incised valleys, filled with sandy gravel deposits, indicate the coexistence of multiple depositional facies.
(2)
The detrital zircon U–Pb age spectrum exhibits major peaks at ~468 Ma and ~964 Ma, with an absence of the ~1170 Ma peak, suggesting provenance derived from the northern margin of the Indian craton rather, than the Lhasa terrane.
(3)
Geochemical characteristics of the Weimei Formation sandstones are marked by high SiO2 contents and K2O/Na2O ratios, low Na2O–CaO–TiO2 contents, right-leaning REE patterns with significant negative Eu anomalies, reflecting derivation from felsic upper crustal and recycled sediment sources within a passive continental margin.
(4)
Integrated analyses indicate that during the Late Jurassic, the Tethyan Himalaya formed part of the passive continental margin along the northern Indian continent. The sedimentary system exhibits an “uplift–depression” paleogeographic framework, with its evolutionary process closely associated with the northward drift of the Indian continent and the evolution of the Neo-Tethys Ocean.

Author Contributions

Investigation, methodology, and writing—original draft preparation, S.Y., J.W., S.C. and T.L.; project administration and data curation, S.Y. and C.X.; resources and writing—review and editing, S.Y., J.W. and H.H. funding acquisition, S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the China Geological Survey Project (No. DD20243073).

Data Availability Statement

Data are contained within the article.

Acknowledgments

We sincerely thank Xuejian Dai, Hu Li, Peng Chen, and Hongrui Dai for their assistance during the fieldwork. The authors are grateful to the anonymous reviewers for their critical and constructive reviews, which greatly improved this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. LA-ICP-MS Zircon U-Pb dating analytical data of the quartz sandstones (WMDN) of the Weimei Formation in the Dingri area.
Table A1. LA-ICP-MS Zircon U-Pb dating analytical data of the quartz sandstones (WMDN) of the Weimei Formation in the Dingri area.
PointsPbThUTh/UIsotope RatioIsotopic Age (Ma)Concordance
(ppm)207Pb/206Pb207Pb/235U206Pb/238U207Pb/206Pb207Pb/235U206Pb/238U
−11833820.400.06900.00221.71380.06090.18010.0035898661014239671994%
−2141551131.370.05620.00300.63040.03420.08120.001246111949621503798%
−31642652621.010.16000.00259.36720.15220.42410.004424572623751522792095%
−4476322.410.06430.00370.64830.03490.07400.0016754121507214601092%
−51024430.550.07690.00261.81250.06330.17100.002411176710502310181396%
−6242922041.430.05410.00160.55970.01740.07500.00103766745111466696%
−711146931.570.05800.00230.59710.02400.07490.00115288747515466697%
−826791130.700.07480.00211.81060.05580.17520.002110655710492010411199%
−91883342941.130.15700.00259.19020.22930.42380.007624242623572322783496%
−108112641.760.05130.00330.52660.03230.07530.001225714843021468791%
−11182471241.990.06240.00220.71110.02730.08260.00126877854516511793%
−121834800.430.07610.00261.88290.07390.17950.00361098671075269642098%
−13525033421.470.06010.00140.80930.02190.09760.00136065260212600899%
−1428591540.380.06990.00161.41530.03370.14710.001792648895148851098%
−15512962181.360.07260.00171.49470.04110.14940.0026100347928178981496%
−16151431381.030.05620.00270.58520.02540.07600.001246110646816472799%
−17442374230.560.05640.00140.63110.01710.08130.00124785649711504798%
−1829801200.670.07760.00181.96730.05690.18370.002511444611041910871498%
−19852760.680.06140.00270.64560.02740.07680.00146549450617477894%
−20183302460.120.23910.003318.69140.34260.56790.008931152230261828993695%
−214674910.810.12740.00266.41310.18490.36480.006720633720342520053198%
−22291911331.430.06990.00241.39770.04580.14570.002392870888198771398%
−231835800.430.07830.00251.93640.06810.17950.00261155591094249651497%
−241089821.090.05590.00330.64560.04070.08380.0018450133506255191197%
−25282272181.040.06180.00170.79060.02520.09280.00167335859214572996%
−26521242310.540.07280.00171.80010.04620.17960.002410094610461710651398%
−2759951690.560.10210.00173.82430.07620.27180.003116653115981615501696%
−2822791820.440.06440.00210.87940.03500.09940.002975469641196111795%
−298108671.600.05220.00320.55570.03690.07690.001529514144924478993%
−3040661970.340.07600.00191.72550.04670.16500.00221095441018179841296%
−3150972400.410.07430.00161.72900.05290.16860.003310504310192010041898%
−3230572600.220.06600.00150.90220.02530.09930.001780642653136101093%
−331351561521.020.23670.003218.60080.38020.57030.008330982230212029093496%
−341934850.400.07820.00231.93300.07540.17890.00391054581093269612197%
−351661352760.490.19190.002911.45750.24820.43340.007027581925612023213193%
−36252482351.050.06000.00160.61950.02020.07480.00116065749013465794%
−3760742800.260.07890.00171.92830.04860.17750.002711724310911710531596%
−38231792200.820.05590.00190.58570.02110.07610.00114507646814473699%
−39162181381.570.05790.00240.59660.02500.07490.00125288947516466798%
−402143363500.960.15840.00219.11120.18860.41750.006524392223491922492995%
−41521762570.690.07360.00181.52790.04000.15090.0021103145942169061296%
−4211131981.330.05550.00240.57400.02580.07500.00124359246117466798%
−432402864000.710.16150.00249.40410.20000.42250.005524722523782022722595%
−44140851690.500.23720.003718.81740.39770.57560.007431022630332029313096%
−451835820.430.07060.00231.74270.06240.17970.00301046671024239651796%
−466119462.600.06160.00440.62800.04320.07480.0020661156495274651293%
−472172233770.590.16130.00229.08620.15100.40870.004124692323471522091993%
−48975890.850.06080.00260.62120.02460.07490.00146329149115466994%
−491444740.590.07200.00311.45330.06940.14580.001898787911298771096%
−50362872231.280.06080.00210.89790.02960.10750.001763269651166581098%
−51182351301.810.05750.00310.65390.03340.08280.001150912351121513799%
−52191401880.750.05590.00190.57910.02350.07480.00094507646415465699%
−5335692000.350.07490.00231.45860.04790.14120.0021106557913208511292%
−54151771241.430.06380.00320.69260.03590.07900.001374410153422490891%
−55221352150.630.05880.00260.60950.02640.07540.00135619648317469896%
−561835790.440.07790.00301.92020.07130.17980.00361043751088259662097%
−57312862671.070.06010.00200.68090.02530.08200.00126093852715508796%
−584456960.580.12620.00285.81420.13650.33430.004320464019492018592195%
−59691863390.550.07480.00151.61340.03640.15630.0018106542975149361095%
−6030881650.530.06940.00191.36610.03980.14280.002192257874178611298%
−6136761380.550.08340.00212.31100.06570.20090.002612805012162011801497%
−6217103561.840.07760.00371.86420.08200.17530.003211379410682910411897%
−6323741250.590.07050.00301.37740.06250.14200.002694388879278561497%
−641172451.580.06730.00311.43610.06440.15600.003085696904279341796%
−655465900.730.16970.00339.69640.24150.41480.007125553224062322373292%
−661679531.490.07730.00331.90400.07350.18090.003611318510822610722098%
−671935850.410.07130.00221.74520.05990.17890.00421066651025229612396%
−6840631540.410.08400.00242.30610.06890.19950.003212925612142111721796%
−691672591.210.07670.00331.86260.07520.17700.002911158510682710501698%
−70291071200.890.07180.00261.68520.06520.17080.00329897210032510161798%
−71393573431.040.06020.00240.66420.02680.08010.00106138551716497695%
−7227891360.650.07430.00221.53880.04920.15060.0026105058946209041595%
−737641200.530.05930.00370.39010.02210.04860.000958913733416306691%
−7421971400.700.06910.00291.07130.04310.11310.002090286739216901193%
−751540880.450.07240.00271.37460.05380.13800.002099878878238331194%
−76413024060.740.05690.00170.59460.01880.07600.00124876747412472799%
−77311051660.630.06910.00251.34690.05170.14160.002490271866228541398%
−781834820.420.07810.00311.92220.08130.17970.00421050501089289652397%
−79722173940.550.07000.00141.36530.03360.14150.002092843874148531197%
−80172491321.890.06150.00410.67650.03790.08060.0019657143525235001195%
−811092811481.900.16800.00299.67370.19440.41770.005025392824041822502393%
−82421111710.650.08110.00302.02390.07320.18130.003012337311242510741795%
−8320121911.340.07510.00501.45590.09920.14080.00281070135912418491692%
−84571462420.600.07980.00191.95090.04730.17750.002011924810991610531195%
−8528721230.580.07910.00241.91370.06390.17550.002611745610862210421495%
−86171851491.240.06180.00380.65120.03730.07690.001573313350923478993%
−872368970.710.07450.00251.79870.06250.17530.002610576910452310411499%
−8831641330.480.07990.00252.00480.06430.18200.002811946111172210781696%
−891834800.420.07460.00241.83030.06140.17890.00401059661056229612299%
−902276800.960.08370.00312.21160.07640.19250.003112857111852411351795%
−91243191731.840.05880.00240.66380.02760.08200.001956189517175081298%
−921514650.210.06040.00280.66990.03120.08080.001761796521195011096%
−93108761850.410.17210.004310.09910.27300.42600.009625894224442522884393%
−94914584351.050.06970.00151.39500.03250.14520.002092044887148741198%
−955947630.750.29320.005024.39180.48230.60360.008434352632841930443492%
−96885661.280.06180.00620.67940.06340.08130.0021733219526385041395%
−971922173200.680.17560.002910.11410.19880.41770.005626132724451822502691%
−981126503731.740.07680.00151.93290.04870.18220.002711173910931710791598%
−992142164470.480.16010.00277.83760.19820.35530.007824572822132319603791%
−1001833820.410.07560.00321.85050.08140.17840.00371085841064299582099%
−101601863070.610.07290.00181.49270.04130.14870.0026101350927178941596%
−1021858910.640.06630.00411.38240.08570.15170.0029817130881379101696%
−103897631.530.06080.00360.63450.04010.07590.0017632129499254711094%
−10424200.200.05900.01030.65160.12560.08200.0036569389509775082199%
−105311121610.690.07130.00321.40600.06920.14300.002496591891298621496%
−1061732780.400.07780.00461.90780.12510.17970.006111431171084449653498%

References

  1. Garzanti, E. Stratigraphy and sedimentary history of the Nepal Tethys Himalaya passive margin. J. Asian Earth Sci. 1999, 17, 805–827. [Google Scholar] [CrossRef]
  2. Garzanti, E.; Hu, X. Latest Cretaceous Himalayan tectonics: Obduction, collision, or Deccan-related uplift? Gondwana Res. 2015, 28, 165–178. [Google Scholar] [CrossRef]
  3. Liu, T.; Liu, C.Z.; Wu, F.Y.; Ji, W.B.; Zhang, C.; Zhang, W.Q.; Zhang, Z.Y. Timing and mechanism of opening the Neo-Tethys Ocean: Constraints from mélanges in the Yarlung Zangbo suture zone. Sci. China Earth Sci. 2023, 66, 2807–2826. [Google Scholar] [CrossRef]
  4. Zhang, J.J.; Xie, C.M.; Duan, M.L.; Chen, H.C.; Bai, X.T.; Xu, M.W.; Shi, Z. Continental crustal growth and thickness changes caused by the Paleo- to Neo-Tethys Ocean transition: Evidence from the Early Jurassic magmatism in the Lhasa terrane. Lithos 2025, 500, 107994. [Google Scholar] [CrossRef]
  5. Zeng, X.M.; Wang, M.; Yu, Y.P.; Li, H.; Zeng, X.J.; Shen, D. Tectonic evolution of the Meso-Tethys Ocean: Insights from geochemistry and geochronology of the Jurassic ophiolitic complex in the Asa area, central Tibet. Int. Geol. Rev. 2022, 64, 351–374. [Google Scholar] [CrossRef]
  6. Liang, J.C.; Bian, W.W.; Jiao, X.W.; Peng, W.X.; Ma, J.H.; Wang, S.; Ma, Y.M.; Zhang, S.H.; Wu, H.C.; Li, H.Y.; et al. Geochronological results from the Zhela Formation volcanics of the Tethyan Himalaya and their implications for the breakup of eastern Gondwana. Sci. Rep. 2023, 13, 20035. [Google Scholar] [CrossRef]
  7. Wang, J.; Li, X.; White, L.T.; Li, Y.; Zhou, J.; Wei, S.; Wei, Z.; Zhou, M.; Fan, X. Early dextral shear motion during the breakup of East Gondwana documented in clastic dykes in the Tethys Himalaya. Tectonophysics 2024, 872, 230132. [Google Scholar] [CrossRef]
  8. Zhang, J.E.; Xiao, W.J.; Wakabayashi, J.; Windley, B.F.; Han, C.M. A fragment of Argoland from East Gondwana in the NE Himalaya. J. Geophys. Res. Solid Earth 2022, 127, e2021JB022631. [Google Scholar] [CrossRef]
  9. Roche, V.; Leroy, S.; Revillon, S.; Guillocheau, F.; Ruffet, G.; Vétel, W. Gondwana break-up controlled by tectonic inheritance and mantle plume activity: Insights from Natal rift development (South Mozambique, Africa). Tectonics 2022, 34, 290–304. [Google Scholar] [CrossRef]
  10. Shi, Y.R.; Hou, C.Y.; Anderson, J.L.; Yang, T.S.; Ma, Y.M.; Bian, W.W.; Jin, J.J. Zircon SHRIMP U-Pb age of Late Jurassic OIB-type volcanic rocks from the Tethyan Himalaya: Constraints on the initial activity time of the Kerguelen mantle plume. Acta Geochim. 2018, 37, 441–455. [Google Scholar] [CrossRef]
  11. Harry, D.L.; Tejada, M.L.G.; Lee, E.Y.; Wolfgring, E.; Wainman, C.C.; Brumsack, H.J.; Schnetger, B.; Kimura, J.I.; Riquier, L.; Borissova, I.; et al. Evolution of the Southwest Australian Rifted Continental Margin During Breakup of East Gondwana: Results From International Ocean Discovery Program Expedition 369. Geochem. Geophys. Geosyst. 2020, 21, e2020GC009144. [Google Scholar] [CrossRef]
  12. Wang, Z.; Shi, Y.R.; Yang, T.S.; Anderson, J.L.; Zhao, J.X.; Peng, W.X.; Hou, C.Y.; Sun, H.Y.; Bian, W.W.; Ma, Y.M. Early activity of the Kerguelen Mantle plume: Geochronology, geochemistry and Sr-Nd-Pb isotopes of mafic dykes and sills from the Tethyan Himalaya. Int. Geol. Rev. 2023, 65, 512–526. [Google Scholar] [CrossRef]
  13. Sushchevskaya, N.M.; Migdisova, N.A.; Antonov, A.V.; Krymsky, R.S.; Belyatsky, B.V.; Kuzmin, D.V.; Bychkova, Y.V. Geochemical features of the Quaternary lamproitic lavas of Gaussberg Volcano, East Antarctica: Result of the impact of the Kerguelen plume. Geochem. Int. 2014, 52, 1030–1048. [Google Scholar] [CrossRef]
  14. Li, J.G.; Peng, J.G.; Batten, D.J. Palynostratigraphy of a Jurassic–Cretaceous transitional succession in the Himalayan Tethys, southern Xizang (Tibet), China. Cretac. Res. 2013, 46, 123–135. [Google Scholar] [CrossRef]
  15. Krishna, J.; Pathak, D.B. Stratigraphic, biogeographic and environmental signatures in the ammonoid bearing Jurassic–Cretaceous of Himalaya on the south margin of the Tethys. Himal. Geol. 1994, 4, 151–167. [Google Scholar]
  16. Colpaert, C.P.A.M.; Li, G. Uppermost Jurassic to Lower Cretaceous benthic foraminiferal faunas of the Weimei and the Bandingsi localities of northern and southern parts of South Tibet-A preliminary analysis. Cretac. Res. 2021, 124, 104785. [Google Scholar] [CrossRef]
  17. Xia, J.; Zhong, H.M.; Tong, J. Stratigraphic framework of Jurassic period and Early Cretaceous in Luozha, South Tibet, including the function of essential sequence law of the regional geological survey. Acta Sedimentol. Sin. 2007, 25, 10–20, (In Chinese with English Abstract). [Google Scholar]
  18. Hu, X.M.; Jansa, L.; Wang, C. Upper Jurassic–Lower Cretaceous stratigraphy in southeastern Tibet: A comparison with the western Himalayas. Cretac. Res. 2008, 29, 301–315. [Google Scholar] [CrossRef]
  19. Bian, W.W.; Yang, T.S.; Ma, Y.M.; Jin, J.J.; Gao, F.; Wang, S.; Peng, W.X.; Zhang, S.H.; Wu, H.C.; Li, H.Y.; et al. Paleomagnetic and geochronological results from the Zhela and Weimei Formations lava flows of the Eastern Tethyan Himalaya: New insights into the breakup of Eastern Gondwana. J. Geophys. Res. Solid Earth 2019, 124, 44–64. [Google Scholar] [CrossRef]
  20. Xu, K.Z.; Ding, F.; Li, Q.X.; Yang, L.; Li, Y.; Dong, B.B. Petrology and geochemistry of Upper Jurassic Weimei Formation sandstones in southern Tibet: Implications for provenance and tectonic setting. Geosci. J. 2019, 23, 767–790. [Google Scholar] [CrossRef]
  21. Jiao, X.W.; Yang, T.S.; Bian, W.W.; Wang, S.; Ma, J.H.; Peng, W.X.; Zhang, S.H.; Wu, H.C.; Li, H.Y. Middle Jurassic Paleolatitude of the Tethyan Himalaya: New insights into the evolution of the Neo-Tethys Ocean. J. Geophys. Res. Solid Earth 2023, 128, e2023JB026659. [Google Scholar] [CrossRef]
  22. Jin, S.C.; Tong, Y.B.; Sun, X.X.; Zhang, Z.J.; Pei, J.L.; Hou, L.F.; Yang, Z.Y. Extensional tectonics of the Indian passive continental margin in the Middle and Late Jurassic: Constraints from detrital zircon ages in the eastern Tethyan Himalaya. J. Geodyn. 2024, 159, 102019. [Google Scholar] [CrossRef]
  23. Liu, G.; Einsele, G. Sedimentary history of the Tethyan basin in the Tibetan Himalayas. Geol. Rundsch. 1994, 83, 32–61. [Google Scholar] [CrossRef]
  24. Li, J.G.; Rao, X.; Mu, L.; Cui, X.H.; Li, X.; Luo, H.; Liu, P.X. Jurassic integrative stratigraphy, biotas, and paleogeographical evolution of the Qinghai-Tibetan Plateau and its surrounding areas. Sci. China Earth Sci. 2024, 67, 1195–1228. [Google Scholar] [CrossRef]
  25. Colpaert, C.P.A.M.; Li, G. Upper Jurassic to Lower Cretaceous benthic foraminifera from South Tibet (Tethyan Himalayas): Systematic, biostratigraphic, and palaeoecologic implications. Rev. Micropaléontologie 2023, 78, 100713. [Google Scholar] [CrossRef]
  26. Yang, L.; Ding, F.; Li, Y.; Xu, Z.K.; Xie, X.G. Sedimentary characteristics and environmental evolution of the Jurassic Weimei Formation in Yamdrok region of Southern Tibet. Xinjiang Geol. 2019, 37, 401–407, (In Chinese with English Abstract). [Google Scholar]
  27. Li, X.H.; Wang, C.S.; Hu, X.M. Verification from massive sandstone of the Upper Jurassic–Lower Cretaceous in Tibetan Tethys Himalayas. J. Mineral. Petrol. 2000, 20, 45–51, (In Chinese with English Abstract). [Google Scholar]
  28. Han, Z.; Hu, X.M.; Li, J.; Garzanti, E. Jurassic carbonate microfacies and relative sea-level changes in the Tethys Himalaya (southern Tibet). Palaeogeogr. Palaeoclimatol. Palaeoecol. 2016, 456, 1–20. [Google Scholar] [CrossRef]
  29. Xia, J.; Zhong, H.M.; Tong, J.S.; Lu, R.K. Sedimentary facies and palaeogeography of the Lhozhag region in southern Xizang during the Jurassic and Cretaceous. Sediment. Geol. Tethyan Geol. 2005, 25, 8–17, (In Chinese with English Abstract). [Google Scholar]
  30. Gansser, A. Facts and theories on the Himalayas. Eclogae Geol. Helv. 1991, 84, 33–59. [Google Scholar] [CrossRef]
  31. Yin, A.; Harrison, T.M. Geologic evolution of the Himalayan–Tibetan orogen. Annu. Rev. Earth Planet. Sci. 2000, 28, 211–280. [Google Scholar] [CrossRef]
  32. Yin, A. Cenozoic tectonic evolution of the Himalayan orogen as constrained by along-strike variation of structural geometry, exhumation history, and foreland sedimentation. Earth Sci. Front. 2006, 13, 416–515, (In Chinese with English Abstract). [Google Scholar] [CrossRef]
  33. Garzanti, E.; Frette, M.P. Stratigraphic succession of the Thakkhola region (central Nepal)—Comparison with the northwestern Tethys Himalaya. Riv. Ital. Paleontol. Stratigr. 1991, 97, 3–26. [Google Scholar]
  34. Wang, C.S.; Hu, X.M.; Sarti, M.; Scott, R.W.; Li, X.H. Upper Cretaceous oceanic red beds in southern Tibet: A major change from anoxic to oxic, deep-sea environments. Cretac. Res. 2005, 26, 21–32. [Google Scholar] [CrossRef]
  35. Gehrels, G.; Kapp, P.; Decelles, P.; Pullen, A.; Blakey, R.; Weislogel, A.; Ding, L.; Guynn, J.; Martin, A. Detrital zircon geochronology of pre-Tertiary strata in the Tibetan-Himalayan orogen. Tectonics 2011, 30, TC5016. [Google Scholar] [CrossRef]
  36. Tong, Y.B.; Pei, J.L.; Qian, T.; Sun, S.S.; Hou, L.F.; Sun, X.X.; Zhang, Z.J.; Yang, B. Three-stage India-Asia collision proposed by the thrice remagnetizations of the Tethyan Himalaya Terrane. Geophys. Res. Lett. 2024, 51, e2024GL110286. [Google Scholar] [CrossRef]
  37. Hu, X.M.; Jansa, L.; Chen, L.; Griffin, W.L.; O’Reilly, S.Y.; Wang, J.G. Provenance of Lower Cretaceous Wölong volcaniclastics in the Tibetan Tethyan Himalaya: Implications for the final breakup of eastern Gondwana. Sediment. Geol. 2010, 223, 193–205. [Google Scholar] [CrossRef]
  38. Jadoul, F.; Berra, F.; Garzanti, E. The Tethys Himalayan passive margin from Late Triassic to Early Cretaceous (South Tibet). J. Asian Earth Sci. 1998, 16, 173–194. [Google Scholar] [CrossRef]
  39. Bhargava, O.N.; Singh, B.P. Geological evolution of the Tethys Himalaya. Epis. J. Int. Geosci. 2020, 43, 404–416. [Google Scholar] [CrossRef]
  40. Wang, Y.G.; Sun, D.L.; He, G.X. New insights into stratigraphy in the Himalayan region (within China). J. Stratigr. 1980, 4, 55–59, (In Chinese with English Abstract). [Google Scholar]
  41. Hou, K.J. Laser ablation-MC-ICP-MS technique for Hf isotope microanalysis of zircon and its geological applications. Acta Petr. Sin. 2007, 23, 2595–2604. [Google Scholar]
  42. Liu, Y.S.; Gao, S.; Hu, Z.C. Continental and oceanic crust recycling induced melt-peridotite interactions in the trans-North China orogen: U-Pb dating, Hf isotopes, and trace elements in zircons from mantle xenoliths. J. Petrol. 2010, 51, 537–571. [Google Scholar] [CrossRef]
  43. Hoskin, P.W.O.; Black, L.P. Metamorphic zircon formation by solid-state recrystallization of protolith igneous zircon. J. Metamorph. Geol. 2000, 18, 423–439. [Google Scholar] [CrossRef]
  44. Hoskin, P.W.O.; Schaltegger, U. The composition of zircon and igneous and metamorphic petrogenesis. Rev. Mineral. Geochem. 2003, 53, 27–62. [Google Scholar] [CrossRef]
  45. Sircombe, K.N. Tracing provenance through the isotope ages of littoral and sedimentary detrital zircon, eastern Australia. Sediment. Geol. 1999, 124, 47–67. [Google Scholar] [CrossRef]
  46. Bhatia, M.R. Plate tectonics and geochemical composition of sandstones. J. Geol. 1983, 91, 611–627. [Google Scholar] [CrossRef]
  47. Fedo, C.M.; Nesbitt, H.W.; Young, G.M. Unraveling the effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleoweathering conditions and provenance. Geology 1995, 23, 921–924. [Google Scholar] [CrossRef]
  48. McLennan, S.M.; Taylor, S.R. Sedimentary rocks and crustal evolution: Tectonic setting and secular trends. J. Geol. 1991, 99, 1–21. [Google Scholar] [CrossRef]
  49. Condie, K.C. Another look at rare earth elements in shales. Geochim. Cosmochim. Acta 1991, 55, 2527–2531. [Google Scholar] [CrossRef]
  50. Sun, S.S.; McDonough, W.F. Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and processes. In Magmatism in the Ocean Basins; Saunders, A.D., Norry, M.J., Eds.; Geological Society: London, UK, 1989; Volume 42, pp. 313–345. [Google Scholar]
  51. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell Scientific Publications: Oxford, UK, 1985; pp. 1–312. [Google Scholar]
  52. Rudnick, R.L.; Gao, S. Composition of the continental crust. In The Crust, Treatise on Geochemistry; Rudnick, R.L., Ed.; Elsevier-Pergamon: Oxford, UK, 2003; Volume 3, pp. 1–64. [Google Scholar]
  53. Cullers, R.L. Implications of elemental concentrations for provenance, redox conditions, and metamorphic studies of shales and limestones near Pueblo, CO, USA. Chem. Geol. 2002, 191, 305–327. [Google Scholar] [CrossRef]
  54. McBride, E.F. Quartz cement in sandstones: A review. Earth-Sci. Rev. 1989, 26, 69–112. [Google Scholar] [CrossRef]
  55. Cawood, P.A.; Nemchin, A.A. Provenance record of a rift basin: U/Pb ages of detrital zircons from the Perth Basin, Western Australia. Sedimentology 2000, 134, 209–234. [Google Scholar] [CrossRef]
  56. Leier, A.L.; Kapp, P.; Gehrels, G.E.; DeCelles, P.G. Detrital zircon geochronology of Carboniferous–Cretaceous strata in the Lhasa terrane, Southern Tibet. Basin Res. 2007, 19, 361–378. [Google Scholar] [CrossRef]
  57. Zhu, D.C.; Zhao, Z.D.; Niu, Y.; Dliek, Y.; Mo, X.X. Lhasa terrane in southern Tibet came from Australia. Geology 2011, 39, 727–730. [Google Scholar] [CrossRef]
  58. DeCelles, P.G.; Gehrels, G.E.; Ouade, J.; LaReau, B.; Spurlin, M. Tectonic implications of U-Pb zircon ages of the Himalayan orogenic belt in Nepal. Science 2000, 288, 497–499. [Google Scholar] [CrossRef] [PubMed]
  59. Gehrels, G.E.; Decelles, P.G.; Ojha, T.P.; Upreti, B.N. Geologic and U-Pb geochronologic evidence for early Paleozoic tectonism in the Dadeldhura thrust sheet, far-west Nepal Himalaya. J. Asian Earth Sci. 2006, 28, 385–408. [Google Scholar] [CrossRef]
  60. Myrow, P.M.; Hughes, N.C.; Goodge, J.W.; Fanning, C.M.; Williams, I.S.; Peng, S. Extraordinary transport and mixing of sediment across Himalayan central Gondwana during the Cambrian–Ordovician. Geol. Soc. Am. Bull. 2010, 122, 1660–1670. [Google Scholar] [CrossRef]
  61. Dong, C.Y.; Li, C.; Wan, Y.S.; Wang, W.; Wu, Y.W. Detrital zircon age model of Ordovician Wenquan quartzite south of Lungmuco–Shuanghu Suture in the Oiangtang area, Tibet: Constraint on tectonic affinity and source regions. Sci. China Earth Sci. 2011, 54, 1034–1042. [Google Scholar] [CrossRef]
  62. Li, X.H.; Mattern, F.; Zhang, C.K.; Zeng, Q.; Mao, G. Multiple sources of the Upper Triassic flysch in the eastern Himalaya Orogen, Tibet, China: Implications to palaeogeography and palaeotectonic evolution. Tectonophysics 2016, 666, 12–22. [Google Scholar] [CrossRef]
  63. Wang, J.G.; Wu, F.Y.; Garzanti, E.; Hu, X.; Ji, W.Q.; Liu, Z.C.; Liu, X.C. Upper Triassic turbidites of the northern Tethyan Himalaya (Langjiexue Group): The terminal of a sediment-routing system sourced in the Gondwanide Orogen. Gondwana Res. 2016, 34, 84–98. [Google Scholar] [CrossRef]
  64. Myrow, P.M.; Hughes, N.C.; Searle, M.P.; Fanning, C.M.; Peng, S.C.; Parcha, S.K. Stratigraphic correlation of Cambrian–Ordovician deposits along the Himalaya: Implications for the age and nature of rocks in the Mount Everest region. Geol. Soc. Am. Bull. 2009, 121, 323–332. [Google Scholar] [CrossRef]
  65. Veevers, J.J.; Saeed, A.; Belousova, E.A.; Griffin, W.L. U-Pb ages and source composition by Hf-isotope and trace-element analysis of detrital zircons in Permian sandstone and modern sand from southwestern Australia and a review of the paleogeographical and denudational history of the Yilgarn Craton. Earth-Sci. Rev. 2005, 68, 245–279. [Google Scholar] [CrossRef]
  66. Ma, A.L.; Hu, X.M.; Garzanti, E.; Han, Z.; Lai, W. Sedimentary and tectonic evolution of the southern Qiangtang basin: Implications for the Lhasa–Qiangtang collision timing. J. Geophys. Res.-Solid Earth 2017, 122, 4790–4813. [Google Scholar] [CrossRef]
  67. Lai, W.; Hu, X.M.; Zhu, D.C.; An, W.; Ma, A.L. Discovery of the early Jurassic Gajia mélange in the Bangong–Nujiang suture zone: Southward subduction of the Bangong–Nujiang Ocean? Int. J. Earth Sci. 2017, 106, 1277–1288. [Google Scholar] [CrossRef]
  68. Sun, G.Y.; Hu, X.M. Tectonic affinity of Zhongba terrane: Evidences from the detrital zircon geochronology and Hf isotopes. Acta Petrol. Sin. 2012, 28, 1635–1646, (In Chinese with English Abstract). [Google Scholar]
  69. Du, X.J.; Chen, X.; Wang, C.S.; Wei, Y.S.; Li, Y.L.; Jansa, L. Geochemistry and detrital zircon U-Pb dating of Lower Cretaceous volcaniclastics in the Babazhadong section, Northern Tethyan Himalaya: Implications for the breakup of Eastern Gondwana. Cretac. Res. 2015, 52, 127–137. [Google Scholar] [CrossRef]
  70. McQuarrie, N.; Robinson, D.; Long, S.; Tobgay, T.; Grujic, D.; Gehrels, G.; Ducea, M. Preliminary stratigraphic and structural architecture of Bhutan: Implications for the along strike architecture of the Himalayan system. Earth Planet. Sci. Lett. 2008, 272, 105–117. [Google Scholar] [CrossRef]
  71. Cheng, M.; Hu, X.M.; Tang, Y.; Deng, Z.; Min, Y.Z.; Chen, S.Y.; Sun, S.J.; Zhou, H.Z. Pan-African and Early Paleozoic Orogenic Events in Southern Tibet: Evidence from Geochronology and Geochemistry of the Kangbuzhenri Gneissic Granite in the Zhegu Area. Minerals 2024, 14, 845. [Google Scholar] [CrossRef]
  72. Meng, Z.Y.; Wang, J.G.; Ji, W.Q.; Zhang, H.; Wu, F.Y.; Garzanti, E. The Langjiexue Group is an in situ sedimentary sequence rather than an exotic block: Constraints from coeval Upper Triassic strata of the Tethys Himalaya (Qulonggongba Formation). Sci. China Earth Sci. 2019, 62, 783–797, (In Chinese with English Abstract). [Google Scholar] [CrossRef]
  73. Jiao, X.; Shi, Y.R.; Yang, T.S.; Bian, W.; Wang, S.; Peng, W. U-Pb age of detrital zircons from Lower Cretaceous in eastern Tethyan Himalaya and its paleogeography. J. Earth Sci. 2021, 46, 2850–2859, (In Chinese with English Abstract). [Google Scholar]
  74. Jiao, X.W.; Yang, T.S.; Bian, W.W.; Wang, S.; Peng, W.X.; Ma, J.H.; Ding, J.K.; Liang, J.C.; Zhang, Y.B.; Li, H.Y.; et al. First paleomagnetic and geochronological results from the early Cretaceous volcanic rocks in the western Tethyan Himalaya: Contribution to the breakup of eastern Gondwana and the paleogeography of the Neo-Tethys Ocean. J. Geophys. Res. Solid Earth 2025, 130, e2024JB031095. [Google Scholar] [CrossRef]
  75. Pan, W.Y. The Sedimentary Features and Paleogeographic Implications of the Late Jurassic Quartz Sandstones in the Tethys Himalaya of Southern Tibet. Master’s Thesis, China University of Geosciences, Beijing, China, 2018. (In Chinese with English Abstract). [Google Scholar]
  76. Roser, B.P.; Korsch, R.J. Provenance signatures of sandstone-mudstone suites determined using discriminant function analysis of major-element data. Chem. Geol. 1988, 67, 119–139. [Google Scholar] [CrossRef]
  77. Gu, X.X.; Liu, J.M.; Zheng, M.H.; Tang, J.X.; Qi, L. Provenance and tectonic setting of the Proterozoic turbidites in Hunan, South China: Geochemical evidence. J. Sediment. Res. 2002, 72, 393–407. [Google Scholar] [CrossRef]
  78. Floyd, P.A.; Leveridge, B.E. Tectonic environment of the Devonian Gramscatho basin, south Cornwall: Framework mode and geochemical evidence from turbiditic sandstones. J. Geol. Soc. 1987, 144, 531–542. [Google Scholar] [CrossRef]
  79. Bhatia, M.R.; Crook, K.A.W. Trace element characteristics of graywackes and tectonic setting discrimination of sedimentary basins. Contrib. Mineral. Petrol. 1986, 92, 181–193. [Google Scholar] [CrossRef]
  80. DeCelles, P.G.; Giles, K.A. Foreland basin systems. Basin Res. 1996, 8, 105–123. [Google Scholar] [CrossRef]
  81. Sepidbar, F.; Homam, S.M.; Ghaemi, F.; Stern, R.J.; Jun, H.; Karsli, O.; Gholami, M. Cadomian tectonic evolution of Iran: Records of an unusually hot and broad extensional convergent margin on the northern margin of Gondwana. Int. Geol. Rev. 2024, 66, 1352–1372. [Google Scholar] [CrossRef]
  82. Li, S.; Li, Y.; Tan, X.; Todrani, A.; Han, Z.; Cheng, J.; Xiao, S.; Ma, X.; Li, Z.; Xu, Y.; et al. Paleomagnetic results from the Early–Middle Jurassic rocks in the Tethyan Himalayas and tectonic implications. J. Asian Earth Sci. 2024, 265, 106088. [Google Scholar] [CrossRef]
  83. Xie, L. The Background and Styles of the Mesozoic Continental Margin Rift Basin in Eastern Tethys Himalaya, South Tibet. Master’s Thesis, Chengdu University of Technology, Chengdu, China, 2017; pp. 1–76, (In Chinese with English Abstract). [Google Scholar]
  84. He, D.F. Study on Deposition Palaeogeography Feature in Late Jurassic, in Southern Tibetan Plateau. Master’s Thesis, Chengdu University of Technology, Chengdu, China, 2009; pp. 1–69, (In Chinese with English Abstract). [Google Scholar]
  85. Brookfield, M.E. The Himalayan passive margin from Precambrian to Cretaceous. Sediment. Geol. 1993, 84, 1–35. [Google Scholar] [CrossRef]
  86. Zhu, D.C.; Zhao, Z.D.; Niu, Y.; Dilek, Y.; Hou, Z.Q.; Mo, X.X. The origin and pre-Cenozoic evolution of the Tibetan Plateau. Gondwana Res. 2013, 23, 1429–1454. [Google Scholar] [CrossRef]
  87. Guo, L.; Zhang, H.F.; Harris, N.; Xu, W.C.; Pan, F.B. Detrital zircon U-Pb geochronology, trace-element and Hf-isotope geochemistry of the metasedimentary rocks in the Eastern Himalayan syntaxis: Tectonic and paleogeographic implications. Gondwana Res. 2017, 41, 207–221. [Google Scholar] [CrossRef]
  88. Zhu, Z.W.; Zhu, X.Y.; Zhang, Y.M. Palaeomagnetic Observation in Xizang and Continental Drift. Acta Geophys. Sin. 1981, 1, 40–49, (In Chinese with English Abstract). [Google Scholar]
  89. Li, J.Z.; Feng, X.T.; Zhu, T.X.; Zhuang, Z.H.; Pan, Z.X.; Zou, G.F. New Paleomagnetic Results from the Tethyan Himalaya in Southern Tibet. Prog. Nat. Sci. 2006, 16, 578–583, (In Chinese with English Abstract). [Google Scholar]
  90. Zhang, B.S.; Wei, Y.S.; Garzanti, E.; Wang, C.S.; Chen, X.; Pan, W.Y.; Liu, Q.S. Sedimentologic and stratigraphic constraints on the orientation of the Late Triassic northern Indian passive continental margin. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 533, 109234. [Google Scholar] [CrossRef]
  91. Garzanti, E.; Haas, R.; Jadoul, F. Ironstones in the Mesozoic passive margin sequence of the Tethys Himalaya (Zanskar, Northern India): Sedimentology and metamorphism. Geol. Soc. Lond. Spec. Publ. 1989, 46, 229–244. [Google Scholar] [CrossRef]
  92. Lee, E.Y.; Fathy, D.; Xiang, X.X.; Spahić, D.; Ahmed, M.S.; Fathi, E.; Sami, M. Middle Miocene syn-rift sequence on the central Gulf of Suez, Egypt: Depositional environment, diagenesis, and their roles in reservoir quality. Mar. Pet. Geol. 2025, 174, 107305. [Google Scholar] [CrossRef]
  93. Zhu, D.C.; Mo, X.X.; Pan, G.T.; Zhao, Z.D.; Dong, G.C.; Shi, Y.R.; Liao, Z.L.; Wang, L.Q.; Zhou, C.Y. Petrogenesis of the earliest Early Cretaceous mafic rocks from the Cona area of the eastern Tethyan Himalaya in south Tibet: Interaction between the incubating Kerguelen plume and the eastern Greater India lithosphere? Lithos 2008, 100, 147–173. [Google Scholar] [CrossRef]
  94. Van Hinsbergen, D.J.J.; Lippert, P.C.; Dupont-Nivet, G.; McQuarrie, N.; Doubrovine, P.V.; Spakman, W.; Torsvik, T.H. Greater India basin hypothesis and a two-stage Cenozoic collision between India and Asia. Proc. Natl. Acad. Sci. USA 2012, 109, 7659–7664. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Simplified tectonic map of the Tibetan Plateau (modified after [36]). (b) Simplified geological map of the Himalayan Orogen [37]. (c) Simplified geological map of the Dingri region. Abbreviations: AKMSZ = Ayimaqing–Kunlun–Muztagh Suture Zone; JSSZ = Jinshajiang Suture Zone; BNSZ = Bangong–Nujiang Suture Zone; YZSZ = Yarlung–Zangbo Suture Zone; STDS = South Tibet Detachment System; MCT = Main Central Thrust; MBT = Main Boundary Thrust; MFT = Main Frontal Thrust; DFG = Dingri–Gangba Fault.
Figure 1. (a) Simplified tectonic map of the Tibetan Plateau (modified after [36]). (b) Simplified geological map of the Himalayan Orogen [37]. (c) Simplified geological map of the Dingri region. Abbreviations: AKMSZ = Ayimaqing–Kunlun–Muztagh Suture Zone; JSSZ = Jinshajiang Suture Zone; BNSZ = Bangong–Nujiang Suture Zone; YZSZ = Yarlung–Zangbo Suture Zone; STDS = South Tibet Detachment System; MCT = Main Central Thrust; MBT = Main Boundary Thrust; MFT = Main Frontal Thrust; DFG = Dingri–Gangba Fault.
Minerals 15 01040 g001
Figure 2. Field photographs of the Weimei Formation in the Dingri area. (a) General view of the Weimei Formation; (b) Outcrop of the Lower Member; (c) Outcrop of the Middle Member; (d) Outcrop of the Upper Member.
Figure 2. Field photographs of the Weimei Formation in the Dingri area. (a) General view of the Weimei Formation; (b) Outcrop of the Lower Member; (c) Outcrop of the Middle Member; (d) Outcrop of the Upper Member.
Minerals 15 01040 g002
Figure 3. CL images (a), U–Pb concordia diagrams (b), and age histograms (c) of detrital zircons from the quartz sandstones of the Weimei Formation in the Dingri area.
Figure 3. CL images (a), U–Pb concordia diagrams (b), and age histograms (c) of detrital zircons from the quartz sandstones of the Weimei Formation in the Dingri area.
Minerals 15 01040 g003
Figure 4. Chondrite-normalized (a) and upper continental crust (UCC)-normalized (b) REE patterns for sandstones from the Weimei Formation in the Dingri area. Chondrite values are from Sun and McDonough, (1989) [50], and UCC values are from Taylor and McLennan (1985) [51].
Figure 4. Chondrite-normalized (a) and upper continental crust (UCC)-normalized (b) REE patterns for sandstones from the Weimei Formation in the Dingri area. Chondrite values are from Sun and McDonough, (1989) [50], and UCC values are from Taylor and McLennan (1985) [51].
Minerals 15 01040 g004
Figure 5. Field photographs and photomicrographs of the Weimei Formation in the Dingri area. (a) Outcrop of quartz sandstone; (b) Photomicrograph of quartz sandstone; (c) Outcrop of lithic quartz sandstone; (d) Photomicrograph of lithic quartz sandstone; (e) Hand specimen of micritic limestone; (f) Photomicrograph of micritic limestone. Abbreviations: Qz–Quartz; Mi–Mica; Pl–Plagioclase; Cal–Calcite; Ms–Muscovite; Lv–Volcanic debris; Ls–Sedimentary debris.
Figure 5. Field photographs and photomicrographs of the Weimei Formation in the Dingri area. (a) Outcrop of quartz sandstone; (b) Photomicrograph of quartz sandstone; (c) Outcrop of lithic quartz sandstone; (d) Photomicrograph of lithic quartz sandstone; (e) Hand specimen of micritic limestone; (f) Photomicrograph of micritic limestone. Abbreviations: Qz–Quartz; Mi–Mica; Pl–Plagioclase; Cal–Calcite; Ms–Muscovite; Lv–Volcanic debris; Ls–Sedimentary debris.
Minerals 15 01040 g005
Figure 6. Sedimentary structures of the Weimei Formation in the Dingri area. (a) Shoreface facies; (bd) Braided delta facies; (e) Shelf facies; (fi) Incised valley facies.
Figure 6. Sedimentary structures of the Weimei Formation in the Dingri area. (a) Shoreface facies; (bd) Braided delta facies; (e) Shelf facies; (fi) Incised valley facies.
Minerals 15 01040 g006
Figure 7. Comparison of detrital zircon ages from the Himalayan Orogen and adjacent regions. Data sources: Cambrian sedimentary rocks in East India [59,64]; Permian sedimentary rocks in the Perth Basin (Western Australia) from the literature [55,65]; Jurassic sedimentary rocks in the South Qiangtang [66]; Jurassic sedimentary rocks in Lhasa [35,56,67]; Middle-Lower Triassic sedimentary rocks in Zhongba Terrane [68]; Jurassic sedimentary rocks in the Tethyan Himalaya [44,69]; Cambrian–Ordovician sedimentary rocks in the Higher Himalaya [35,58]; Cambrian–Carboniferous sedimentary rocks in the Lesser Himalaya [35,70]. Abbreviations: BO = Bhimphedian Orogeny; PAO = Pan-African Orogen; EGMB, RO = Eastern Ghats Mobile Belt, Rayner Orogen; AFB = Albany Fraser Belt; ADFB = Aravalli-Dehi fold belt.
Figure 7. Comparison of detrital zircon ages from the Himalayan Orogen and adjacent regions. Data sources: Cambrian sedimentary rocks in East India [59,64]; Permian sedimentary rocks in the Perth Basin (Western Australia) from the literature [55,65]; Jurassic sedimentary rocks in the South Qiangtang [66]; Jurassic sedimentary rocks in Lhasa [35,56,67]; Middle-Lower Triassic sedimentary rocks in Zhongba Terrane [68]; Jurassic sedimentary rocks in the Tethyan Himalaya [44,69]; Cambrian–Ordovician sedimentary rocks in the Higher Himalaya [35,58]; Cambrian–Carboniferous sedimentary rocks in the Lesser Himalaya [35,70]. Abbreviations: BO = Bhimphedian Orogeny; PAO = Pan-African Orogen; EGMB, RO = Eastern Ghats Mobile Belt, Rayner Orogen; AFB = Albany Fraser Belt; ADFB = Aravalli-Dehi fold belt.
Minerals 15 01040 g007
Figure 8. Provenance discrimination diagrams for sandstones from the Weimei Formation in the Dingri area. (a) F1–F2 diagram [76]; (b) F3–F4 diagram [76]; (c) Co/Th–La/Sc diagram [77]; (d) La/Th–Hf diagram [78]. Equations for the discriminant functions: F1 = −1.773TiO2 + 0.607Al2O3 + 0.76TFe2O3 − 1.5MgO + 0.616CaO + 0.509Na2O − 1.224K2O − 9.09. F2 = 0.445TiO2 + 0.07Al2O3 − 0.25TFe2O3 − 1.142MgO + 0.438CaO + 1.475Na2O + 1.426K2O − 6.861. F3 = 30.638TiO2/Al2O3 + 12.541TFe2O3/Al2O3 + 7.329MgO/Al2O3 + 12.031Na2O/Al2O3 + 35.4K2O/Al2O3 − 6.382. F4 = 56.5TiO2/Al2O3 + 10.879TFe2O3/Al2O3 + 30.875MgO/Al2O3 + 5.404Na2O/Al2O3 + 11.112K2O/Al2O3 − 3.89.
Figure 8. Provenance discrimination diagrams for sandstones from the Weimei Formation in the Dingri area. (a) F1–F2 diagram [76]; (b) F3–F4 diagram [76]; (c) Co/Th–La/Sc diagram [77]; (d) La/Th–Hf diagram [78]. Equations for the discriminant functions: F1 = −1.773TiO2 + 0.607Al2O3 + 0.76TFe2O3 − 1.5MgO + 0.616CaO + 0.509Na2O − 1.224K2O − 9.09. F2 = 0.445TiO2 + 0.07Al2O3 − 0.25TFe2O3 − 1.142MgO + 0.438CaO + 1.475Na2O + 1.426K2O − 6.861. F3 = 30.638TiO2/Al2O3 + 12.541TFe2O3/Al2O3 + 7.329MgO/Al2O3 + 12.031Na2O/Al2O3 + 35.4K2O/Al2O3 − 6.382. F4 = 56.5TiO2/Al2O3 + 10.879TFe2O3/Al2O3 + 30.875MgO/Al2O3 + 5.404Na2O/Al2O3 + 11.112K2O/Al2O3 − 3.89.
Minerals 15 01040 g008
Figure 9. Tectonic discrimination diagrams for sandstones from the Weimei Formation in the Dingri area. (a) K2O/Na2O–SiO2 diagram [76]; (b) Ti/Zr–La/Sc diagram [79]; (c) Th–Co–Zr/10, La–Th–Sc, and Th–Sc–Zr/10 diagrams [79].
Figure 9. Tectonic discrimination diagrams for sandstones from the Weimei Formation in the Dingri area. (a) K2O/Na2O–SiO2 diagram [76]; (b) Ti/Zr–La/Sc diagram [79]; (c) Th–Co–Zr/10, La–Th–Sc, and Th–Sc–Zr/10 diagrams [79].
Minerals 15 01040 g009
Figure 10. Stratigraphic column correlation of the Late Jurassic in the Tethyan Himalaya. Sources of stratigraphic columns: (1) Cuomei area [83]; (2) Langkazi area [83]; (3) Gyangze area [84]; (4) Sakya area [84]; (5) Dingri area, Weimei Formation (this study); (6) Gyirong area [75]; (7) Dingri area, Menkadun Formation (this study). (a) Lithofacies paleogeographic map of the Gyirong–Longzi area; (b) Stratigraphic comparison of the Weimei Formation in the northern subzone; (c) Stratigraphic comparison of the Weimei Formation in the southern and northern subzone.
Figure 10. Stratigraphic column correlation of the Late Jurassic in the Tethyan Himalaya. Sources of stratigraphic columns: (1) Cuomei area [83]; (2) Langkazi area [83]; (3) Gyangze area [84]; (4) Sakya area [84]; (5) Dingri area, Weimei Formation (this study); (6) Gyirong area [75]; (7) Dingri area, Menkadun Formation (this study). (a) Lithofacies paleogeographic map of the Gyirong–Longzi area; (b) Stratigraphic comparison of the Weimei Formation in the northern subzone; (c) Stratigraphic comparison of the Weimei Formation in the southern and northern subzone.
Minerals 15 01040 g010
Figure 11. Paleogeographic reconstruction of the Tethyan Himalaya during the Late Jurassic. (a) Jurassic paleogeographic reconstruction of the Neo-Tethys Ocean [21]; (b) Facies distribution of the Tethyan Himalaya in the Late Jurassic.
Figure 11. Paleogeographic reconstruction of the Tethyan Himalaya during the Late Jurassic. (a) Jurassic paleogeographic reconstruction of the Neo-Tethys Ocean [21]; (b) Facies distribution of the Tethyan Himalaya in the Late Jurassic.
Minerals 15 01040 g011
Table 1. Major (wt%) and trace (ppm) elemental contents of sandstones from the Weimei Formation in the Dingri area.
Table 1. Major (wt%) and trace (ppm) elemental contents of sandstones from the Weimei Formation in the Dingri area.
SampleWMFX1WMFX2WMFX3WMFX4WMFX5KMFX1KMFX2KMFX3KMFX4KMFX5KMFX6KMFX7
LithologyQuartz SandstoneLithic quartz sandstone
SiO295.795.296.396.896.870.265.570.365.062.967.067.8
Na2O0.130.030.040.060.060.370.630.340.730.400.420.51
K2O0.310.080.070.140.122.793.312.492.582.352.303.08
FeO0.631.180.380.260.294.314.244.742.972.660.296.54
Fe2O30.070.291.060.350.251.982.322.080.620.460.020.55
P2O50.020.030.020.040.040.080.070.050.050.040.010.26
TiO20.160.090.150.160.161.001.130.920.260.280.181.62
MgO0.160.300.080.040.043.002.993.107.067.505.212.50
CaO0.060.040.050.060.050.170.200.1716.916.318.43.36
Al2O31.911.981.121.481.4015.017.914.34.134.104.4510.5
MnO0.000.010.020.010.010.020.020.020.050.040.020.08
LOI0.361.030.750.740.663.533.934.112.001.982.732.26
Y4.063.824.653.433.2813.613.312.39.558.727.718.4
La11.98.8510.911.911.041.251.047.212.614.010.022.3
Ce27.027.527.227.425.096.011710925.826.420.449.6
Pr2.922.502.692.742.5410.113.111.42.822.962.356.36
Nd10.69.6510.210.39.0638.548.742.21111.19.5827.4
Sm1.831.801.861.871.697.439.097.942.32.162.026.64
Eu0.290.310.260.320.311.481.681.380.460.380.321.92
Gd1.231.331.51.471.55.736.215.492.062.061.746.26
Tb0.180.210.210.200.200.710.730.660.340.320.260.90
Dy0.850.890.980.880.873.463.363.021.861.761.374.56
Ho0.170.180.190.160.160.570.610.530.380.350.310.79
Er0.460.470.500.420.381.641.721.51.071.080.922.06
Tm0.080.090.090.080.080.230.260.220.170.170.160.28
Yb0.350.420.430.310.301.51.541.390.960.950.921.62
Lu0.110.100.110.100.090.280.280.240.190.190.190.28
As0.845.1445.42.532.564.0014.710.66.247.152.122.08
Ba10225.928.458.655.053061148081.955.449.049.6
Cd0.020.040.060.030.020.050.050.030.070.030.020.11
Co1.001.146.252.943.4211.316.810.32.271.960.6824.2
Cr21.413.110.111.412.410111292.720.71810.9156
Hf3.802.424.805.034.6613.315.114.14.947.143.737.48
Mo0.410.360.370.330.340.430.570.400.220.230.200.51
Nb4.0934.224.874.0420.225.823.25.885.343.8121
Ni7.7614.312.925.328.643.648.941.67.725.574.5168.1
Pb4.484.282.34.614.1826.126.126.53.843.583.17.09
Rb13.53.763.135.484.9410412496.122.414.212.84.4
Sc1.72.191.271.121.0813.415.813.62.621.981.2610.8
Sr30.812.719.429.027.4112149100338374419144
Ta0.350.350.460.510.291.411.751.530.6440.5020.4071.4
Th4.83.514.625.355.1021.323.824.67.157.184.735.92
U0.251.010.400.350.311.931.891.460.440.520.330.77
V17.440.514.614.413.915715815228.1288.65112
Zr17910118920418043350051219221493.3254
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

Wang, J.; Yan, S.; Huang, H.; Liu, T.; Xin, C.; Chen, S. Sedimentary Environment, Tectonic Setting, and Paleogeographic Reconstruction of the Late Jurassic Weimei Formation in Dingri, Southern Tibet. Minerals 2025, 15, 1040. https://doi.org/10.3390/min15101040

AMA Style

Wang J, Yan S, Huang H, Liu T, Xin C, Chen S. Sedimentary Environment, Tectonic Setting, and Paleogeographic Reconstruction of the Late Jurassic Weimei Formation in Dingri, Southern Tibet. Minerals. 2025; 15(10):1040. https://doi.org/10.3390/min15101040

Chicago/Turabian Style

Wang, Jie, Songtao Yan, Hao Huang, Tao Liu, Chongyang Xin, and Song Chen. 2025. "Sedimentary Environment, Tectonic Setting, and Paleogeographic Reconstruction of the Late Jurassic Weimei Formation in Dingri, Southern Tibet" Minerals 15, no. 10: 1040. https://doi.org/10.3390/min15101040

APA Style

Wang, J., Yan, S., Huang, H., Liu, T., Xin, C., & Chen, S. (2025). Sedimentary Environment, Tectonic Setting, and Paleogeographic Reconstruction of the Late Jurassic Weimei Formation in Dingri, Southern Tibet. Minerals, 15(10), 1040. https://doi.org/10.3390/min15101040

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