Next Article in Journal
Evaluation and Classification of Uranium Prospective Areas in Madagascar: A Geochemical Block-Based Approach
Previous Article in Journal
The Composition of Volatiles in Quartz and Pyrite from the Konduyak Gold Deposit (Yenisei Ridge, Russia)
Previous Article in Special Issue
Characteristics of Seismogenic Dust Particles from a Mountain and Their Significance for Paleoseismic Records in a Tufa Section: A Case Study of Jiuzhaigou, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Petrography and Geochemistry of Lower Jurassic Sandstones in the Eastern Junggar Basin: Implications for Provenance and Tectonic Setting

1
State Key Laboratory of Nuclear Resources and Environment, East China University of Technology, Nanchang 330013, China
2
Inner Mongolia Engineering Research Center of Geological Technology and Geotechnical Engineering, Inner Mongolia University of Technology, Hohhot 010051, China
3
Key Laboratory of Geological Hazards and Geotechnical Engineering Defense in Sandy and Drought Regions at Universities of Inner Mongolia Autonomous Region, Inner Mongolia University of Technology, Hohhot 010051, China
4
Langfang Integrated Natural Resources Survey Center, China Geological Survey, Langfang 065000, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(3), 279; https://doi.org/10.3390/min15030279
Submission received: 23 January 2025 / Revised: 2 March 2025 / Accepted: 6 March 2025 / Published: 9 March 2025

Abstract

:
The Junggar Basin basement comprises microcontinental blocks amalgamated through successive paleo-oceanic accretion events. Stratigraphic and provenance studies within the basin are crucial for reconstructing its evolution and understanding the closure of paleo-oceanic systems. This study presents an integrated petrographic and geochemical analysis of the Lower Jurassic Badaowan Formation sandstones in the Dongdaohaizi Depression, located in the eastern Junggar Basin. The results reveal a progressive decrease in lithic fragment content and an increase in quartz content from older to younger strata within the Badaowan Formation, indicating an increase in compositional maturity. Provenance analysis indicates that the sandstones are predominantly derived from tuffaceous rocks, granites, basalts, and minor metamorphic rocks. Heavy mineral assemblages, including zircon, chromian spinel, tourmaline, and garnet, suggest parent rocks consisting primarily of intermediate to acidic igneous rocks, mafic igneous rocks, and metamorphic rocks. Integrated petrographic and geochemical data from the surrounding areas of the Dongdaohaizi Depression confirm that the Badaowan Formation sandstones are primarily sourced from the eastern Kelameili Mountain. The continued uplift and migration of the Kelameili Mountain during the Early Jurassic played a dominant role in shaping the sedimentary provenance. LA-ICP-MS analyses reveal that the rare earth element (REE) concentrations in the Lower Jurassic sandstones are slightly lower than the average REE content of the upper continental crust. The sandstones exhibit weak differentiation between light and heavy REEs, reflecting a depositional environment characterized by anoxic reducing conditions. Geochemical results indicate a tectonic setting dominated by a passive continental margin and continental island arc in the source area. Synthesizing these findings with related studies, we propose that the Kelameili Ocean, as part of the Paleo-Asian Ocean, underwent a complex evolution involving multiple oceanic basins and microcontinental subduction–collision systems. From the Middle Ordovician to Late Silurian, the Kelameili region evolved as a passive continental margin. With the onset of subduction during the Middle Devonian to Early Carboniferous, the eastern Junggar Basin transitioned into a continental island arc system. This tectonic transition was likely driven by episodic or bidirectional subduction of the Kelameili Ocean.

1. Introduction

The Junggar Basin in northwestern China is composed of multiple microcontinental blocks amalgamated through paleo-oceanic tectonic movements. Branches of the Paleo-Asian Ocean, including the Kelameili, Armantai, and Erqis belts, began converging during the Ordovician and completely closed by the early Mesozoic [1,2,3]. The closure of these paleo-oceanic basins profoundly influenced the structural framework of the Junggar Basin and played a pivotal role in shaping the formation and evolution of Central Asia. The compressional forces resulting from oceanic closure triggered folding and faulting within certain strata, creating specific geological conditions that influenced subsequent sedimentation and hydrocarbon generation [4,5,6].
The Dongdaohaizi Depression, situated along the eastern margin of the Junggar Basin, represents a structurally complex region shaped by the tectonic evolution of the Paleo-Asian Ocean, including oceanic opening, subduction, and continental collision [7,8,9]. As a sub-unit of the Junggar Basin, its stratigraphic sequences and structural architecture reflect the dynamic interplay between regional tectonism and sedimentary processes [10,11,12]. Subduction-related volcanic activity and associated tectonic reworking have left discernible imprints in the stratigraphic record of the depression, underscoring the need for a detailed investigation of its sedimentary provenance and tectonic history [13,14,15].
The Junggar Basin is characterized by diverse sedimentary sources and depositional systems. However, studies on sediment provenance within its tectonic units often lack high-precision, quantifiable methodologies [16,17,18]. Previous research on provenance has largely focused on the basin as a whole, employing conventional methods such as gravel-to-matrix ratios, sand-to-matrix ratios, dip angle logging, and seismic data to infer sedimentary systems. These studies predominantly relied on sedimentary system maps to determine provenance, with limited application of advanced techniques [19,20,21]. In the Dongdaohaizi Depression, provenance studies have primarily been based on basin-scale depositional facies maps and analyses of orogenic belts at the basin margins. However, the limited resolution and narrow scope of these methods have led to controversies regarding the sources and extent of sediment supply within the study area [11,22,23]. This is particularly true for the Lower Jurassic, where the lithological characteristics and sediment supply mechanisms of clastic deposits remain unclear. Therefore, a quantitative analysis of sediment provenance is urgently needed to elucidate the evolution of the surrounding paleo-oceanic systems, building upon existing research.
In this study, we conducted systematic sampling and analyses, including petrographic observations, heavy mineral composition analysis, and rare earth element (REE) and trace element geochemistry. These data were integrated with other geological evidence (e.g., geophysics, isotopic chronology) to investigate the Lower Jurassic Badaowan Formation sandstones in the Dongdaohaizi Depression, located in the eastern Junggar Basin. The study provides a comprehensive petrographic and geochemical analysis of these sandstones, elucidates their provenance characteristics, and summarizes the tectonic background and evolutionary history of the provenance region. Our research not only advances the understanding of paleo-oceanic closure processes but also provides valuable insights for hydrocarbon and mineral exploration in the Dongdaohaizi Depression and the Kelameili Mountain region.

2. Geological Background

The Junggar Basin, situated in the northern part of Xinjiang, China, lies at the convergence of the Kazakhstan Plate, Siberian Plate, and Tarim Plate, serving as the eastward extension of the Kazakhstan Plate (Figure 1a,b) [24,25,26]. The basin is characterized by a triangular shape, being wider in the south and narrower in the north. Its primary tectonic units include the Ulungur Depression, Luliang Uplift, Central Depression, Eastern Uplift, Western Uplift, and Southern Margin Thrust Belt. The Dongdaohaizi Depression, a secondary tectonic unit within the Central Depression, is flanked by the Mosuowan Uplift to the west, the Baijiahai Uplift to the east, and the Fukang Depression to the south (Figure 1c).
The Dongdaohaizi Depression was formed in the early Hercynian period, characterized by extensive marine facies comprising shales and marls, as well as calc-alkaline volcanic rocks. Intense fault-volcano activity during the Early Carboniferous to Early Permian established the internal structure of the Central Depression [28]. The tectonic framework of the study area is based on a Devonian basement, with structural deformation primarily shaped by Late Hercynian and Yanshanian tectonic events [29]. Jurassic strata predominantly feature nearly northeast-trending interlayer faults, dominated by normal faults with limited horizontal extension.
The stratigraphic sequence in this depression spans from the Devonian to the Jurassic, with a significant portion of the Jurassic strata being the primary focus of this study [30]. The Devonian basement, representing the ancient continental crust, serves as the underlying foundation for the younger sedimentary deposits in the region. This basement is primarily composed of metamorphic and igneous rocks that have undergone long-term tectonic evolution. The unconformity overlying Permian and Triassic sequences form a continuous sedimentary succession, characterized by varying lithologies, including sandstones, shales, and carbonates, which reflect the tectonic and depositional changes in the basin. The Late Hercynian event, which occurred during the late Paleozoic, played a crucial role in the formation of the basement and the initial development of the basin’s structural framework [21]. This tectonic event was characterized by the closure of the Paleo-Asian Ocean, leading to the collision and amalgamation of microcontinental blocks that contributed to the formation of the basement rocks. The Yanshanian orogeny, which occurred during the Mesozoic, had a profound impact on the basin’s structural evolution [31]. It led to the reactivation of earlier faults and the formation of new tectonic structures.
The Jurassic strata in the Dongdaohaizi Depression include the Badaowan Formation, Xishanyao Formation, and Toutunhe Formation, while the Qigu Formation is absent (Figure 2a) [32]. Among these, the Badaowan Formation and Xishanyao Formation are significant coal-bearing strata [33]. The Xishanyao Formation predominantly consists of fluvial deposits, with sandstone accounting for 50% of the stratum thickness. Single sandstone layers reach a maximum thickness of 14.0 m, with most layers ranging from 2.0 m to 8.0 m, exhibiting considerable lateral thickness variation [10]. The Toutunhe Formation is dominated by continental (lake-swamp) deposits and is composed of sandstone, red and gray sandy mudstone in ascending order [34].
The Lower Jurassic Badaowan Formation (Figure 2b), representing the focus of this study, is primarily composed of fluvial and deltaic deposits. Sandstone beds account for 45% of the formation thickness, with single layers reaching up to 30.0 m in thickness, while the general range is 4.0–22.0 m [10]. The formation displays distinct cyclicity, characterized by coarsening-upward sequences. Its lithological assemblage includes conglomerate, sandstone, mudstone, and coal seams [35]. During the Early Jurassic, the Indosinian orogeny led to the formation of the sags, resulting in a regional angular unconformity between Jurassic and Triassic strata [22,36,37].
The sedimentary center of the Badaowan Formation was primarily located in the Changji Sag and Penyijingxi Sag, where its maximum thickness reaches approximately 1200 m. Lithologically, the formation can be divided into two parts: The upper section consists of interbedded mudstone and sandstone, while the middle and lower sections comprise gray-green or gray-black mudstone interbedded with conglomerate or pebbly sandstone. As the first major coal-bearing formation in the Junggar Basin, the Badaowan Formation serves as a critical hydrocarbon source rock and reservoir. Its first and third members are predominantly sandstone, whereas the second member is dominated by mudstone.
In this study, sandstone samples were systematically collected from multiple stratigraphic intervals within the Badaowan Formation across different locations in the Dongdaohaizi Depression. Specifically, samples from the third member of the Badaowan Formation were obtained from wells C1 and C3, while those from the second member were collected from wells C1 and C3. Additionally, samples from the first member were retrieved from well C3. In the western part of the Dongdaohaizi Depression, samples from the third member were collected from well Zh1 (Figure 1c and Figure 2b). The sampling depths for these wells were 4488 m, 4274 m, 4881 m, 4623 m, 4981 m, and 4658 m, respectively. The sampled intervals exhibit distinct cyclicity, characterized by a well-defined coarsening-upward sequence, and display typical fluvial facies features, including cross-bedding and scour surface.

3. Materials and Methods

3.1. Clastic Materials Statistics Method

Most sediments in clastic basins originate from source regions outside the basin. The classic sandstone QFL (quartz, feldspar, lithics) ternary diagram is among the most comprehensive and effective tools for provenance analysis [38,39,40]. Adequate sandstone samples were collected during field investigations, and suitable samples were selected for analysis [41,42,43,44,45,46]. After removing weathered surfaces, fresh rock faces were prepared for thin-section analysis. To minimize the impact of grain size variability, strict statistical methods were employed [40,47,48]. Framework grains of sandstone samples were statistically analyzed using the Gazzi-Dickinson point-counting method [49]. The maximum grid spacing was selected to ensure that 500 points were counted per thin section, with each grain counted only once. Grain sizes ranged from 0.0625 to 2 mm, while matrix and/or cement are not included in the classification systems used for analysis in this study. All particles larger than 0.0625 mm within larger lithic fragments were classified as monocrystalline grains, reducing compositional differences caused by grain size. Deep-seated igneous rock fragments were not counted as a separate category but were analyzed based on their constituent crystals, such as quartz and feldspar. The grain types considered in this study included monocrystalline quartz (Qm), polycrystalline quartz (Qp), volcanic lithics (Lv), sedimentary lithics (Ls), and metamorphic lithics (Lm).

3.2. Heavy Mineral Analysis Method

Heavy minerals, known for their high stability and resistance to abrasion, are well-preserved during transport from source areas to sedimentary basins, making them invaluable for provenance analysis [50,51,52,53,54]. Cluster analysis of heavy minerals can provide critical insights into source area characteristics [55].
Sedimentary rock samples were collected from the study area and processed through crushing and sieving to extract heavy minerals [56]. Heavy minerals were separated using heavy liquid separation and magnetic separation techniques. The heavy liquid method involved suspending samples in high-density liquids, such as bromoform (2.8 g/cm³), tetrabromoethane (3.0 g/cm³), and sodium polytungstate (3.3 g/cm³), to separate minerals based on density differences. Light minerals (e.g., quartz, feldspar) floated, while heavy minerals (e.g., garnet, zircon, ilmenite) settled at the bottom and were collected for further analysis [57,58,59,60,61]. Magnetic separation was performed using a Frantz Isodynamic Magnetic Separator, which distinguishes minerals based on their magnetic susceptibility. Separation was conducted at electromagnetic field strengths of 0.2 A, 0.6 A, and 1.2 A to progressively isolate paramagnetic and diamagnetic minerals [62,63]. The separated heavy minerals were identified using petrographic microscopy of thin sections. Diagnostic optical properties, such as color, birefringence, interference colors, and crystal habit, were used to distinguish minerals such as pyroxene, amphibole, epidote, garnet, chrome-spinel, monazite, zircon, apatite, and rutile. For example, garnet is isotropic, whereas tourmaline is birefringent.
The relative abundances of heavy minerals were statistically analyzed using the point-counting method, with at least 200 transparent mineral grains counted per sample to ensure data representativeness. Heavy mineral assemblage diagrams were constructed to characterize their composition. All analyses were conducted at Nanjing FocuMS Technology Co. Ltd., China.

3.3. Whole-Rock Geochemical Analysis

The enrichment and abundance of elements in sediments are controlled by multiple factors, including the nature of the source rocks, weathering intensity, transport distance, sorting, and diagenetic processes [38,64,65,66]. Rare earth elements (REEs) and some trace elements (e.g., Th, Sc, Cr, Co) are highly stable and resistant to dissolution during sedimentation and diagenesis. Their concentrations remain largely unchanged during transport, making them reliable indicators of source rock types [67,68,69,70,71]. In this study, whole-rock trace element and REE analyses were conducted using inductively coupled plasma mass spectrometry (ICP-MS). All analyses were performed at Nanjing FocuMS Technology Co. Ltd., China, using an Agilent 7700x ICP-MS (Agilent Technologies, Inc., Santa Clara, CA, USA). REE data were normalized to the North American Shale Composite (NASC) [72,73,74].
The analytical procedure was as follows: Sample powders were ground to 200 mesh and dried in an oven at 105 °C for 12 h. A 25 mg aliquot of the dried powder was placed in a high-pressure digestion vessel. To remove silicon, 0.5 mL of concentrated hydrofluoric acid was added, and the sample was heated at 100 °C until dry. Subsequently, 1 mL of concentrated hydrofluoric acid and 0.5 mL of concentrated nitric acid were added. The inner vessel was placed in a stainless-steel jacket, sealed, and heated in an oven at 110–120 °C for 72 h. After digestion, the vessel was opened, and the sample was dried to a moist salt state at 140 °C. To remove excess hydrofluoric acid, 1 mL of concentrated nitric acid was added and evaporated at 100 °C. Finally, 5 mL of 30% nitric acid was added, and the sample was sealed and heated at 110 °C for 4 h to ensure complete dissolution. Residues were filtered out, and the solution was diluted to 50 mL with 2% nitric acid [75,76].
To ensure accurate measurements of boron (B), which can volatilize at high temperatures, the heating temperatures used in this experiment were carefully controlled and maintained below standard conditions. International standards and blank samples were used to monitor the analytical precision and accuracy. The linearity of measured elements in standard and blank samples was satisfactory, with an analytical error of less than 2%. Repeated measurements of the same sample yielded consistent results, and the final data for each sample represent the average of three measurements [77,78].

4. Results

4.1. Clastic Composition Result

A detailed analysis and summary of thin-section data from samples of the Badaowan Formation within the Dongdaohaizi Depression and its western region were conducted (Table A1). Based on the established compositional classification of clastic rocks, the Badaowan Formation in the study area primarily consists of lithic arkose, feldspathic litharenite, and minor litharenite (Figure 3) [46,79]. Among the lithic fragment components, volcanic lithic components are predominant. Samples from sampling point C1, mainly from the third member of the Badaowan Formation, are characterized by a high proportion of volcanic lithic fragments and exhibit higher compositional maturity. Samples from C3, spanning the first and second members, show a decrease in lithic fragment content and an increase in quartz content from the first to the second members, indicating an increase in compositional maturity, although the overall compositional maturity remains relatively low. In the western region, lithic feldspathic sandstone and litharenite predominate, with minor feldspathic litharenite and lithic fragments arkose, showing significant differences from the study area. Samples from location Zh1 exhibit higher quartz content and lower lithic fragments content, indicating higher compositional maturity.
Petrographic analysis reveals that the lithic components in the study area are dominated by volcanic lithic fragments, with frequent occurrences of tuffaceous lithics. Sedimentary and metamorphic lithic fragments, such as mudstone, slate, phyllite, and quartzite fragments, are less common (Figure 4). These findings suggest that the primary source rocks in the study area are tuff, granite, basalt, and a small proportion of metamorphic rocks [80,81,82].

4.2. Heavy Mineral Analysis Result

Heavy mineral assemblages in the Badaowan Formation show distinct compositional variations between the study area and the western region (Table A2). At sampling point C1, the assemblage includes zircon, chrome-spinel, tourmaline, and garnet, suggesting source rocks dominated by intermediate to acidic igneous rocks, basic igneous rocks, and metamorphic rocks [56]. At sampling location C3, the assemblage consists of zircon, chrome-spinel, and tourmaline, reflecting source rocks primarily composed of intermediate to acidic igneous rocks and basic igneous rocks (or contact metamorphic rocks) [83]. In contrast, the assemblage at sampling location Zh2 in the western region is predominantly composed of zircon, garnet, tourmaline, and chrome-spinel, indicating source rocks mainly consisting of metamorphic rocks, intermediate to acidic igneous rocks, and basic igneous rocks (Figure 5) [84].
The heavy mineral assemblages at C1 and C3 are relatively similar, with the garnet at C1 potentially originating from deep-seated exhumation and erosion. Eastern samples exhibit greater compositional stability and higher zircon content, while samples from the western region show a higher garnet content, underscoring the differences in provenance [56,85].
Prolonged weathering can significantly alter the geochemical composition of rocks, potentially affecting the accuracy of tectonic classifications. Mobile elements, such as alkali and alkaline earth metals (e.g., Na, K, Ca), are more susceptible to leaching during weathering, whereas immobile elements (e.g., Al, Ti, Fe) tend to be retained, which can distort elemental ratios used in ternary diagrams. However, studies have shown that while weathering may cause the dissolution or redistribution of certain minerals, it generally does not alter the types of heavy minerals present.

4.3. Geochemical Analysis Result

The total REE concentrations in shales reflect the average upper crustal REE abundance, often compared against standard compositions such as NASC (North American Shale Composite) or PAAS (Post-Archean Australian Shale). In this study, REE data for the Badaowan Formation were normalized to NASC (Table A3) [72,73,74].
The total REE content in the study area and the western region reveals distinct differences. At sampling point C3, total REE ranges from 77.78 to 94.88 ppm, with an average of 89.56 ppm. At C1, total REE ranges from 148.50 to 201.97 ppm, averaging 175.43 ppm. In the western region, at location Zh1, total REE ranges from 114.17 to 128.03 ppm, with an average of 121.10 ppm (Table A3 and Table A4). Generally, REE concentrations in clastic rocks at C1 exceed the average upper crustal REE value (146.4 ppm), while those at C3 and Zh1 fall below this benchmark [86,87].
The light-to-heavy REE ratio (∑LREE/∑HREE) indicates the degree of differentiation between light and heavy REEs (Table 1) [88,89,90]. Higher ratios suggest greater differentiation and relative enrichment of light REEs. At C1, ∑LREE/∑HREE ranges from 7.58 to 8.56 (mean 8.18), while at C3, it ranges from 7.09 to 7.25 (mean 7.19). In the western region, the ratio ranges from 5.74 to 7.44 (mean 6.59). These data suggest that C1 samples exhibit greater differentiation than NASC (7.44) [72,73,74], while C3 samples show slightly lower differentiation, and the western samples display relative enrichment of heavy REEs. The REE data indicate similar provenance for C1 and C3, with notable differences compared to the western region.
The LaN/YbN ratio, reflecting the slope of the REE distribution curve, and the LaN/SmN and GdN/YbN ratios, indicating light and heavy REE fractionation, show that both the study area and the western region have relatively flat distribution curves [92,93]. This implies minimal fractionation among both light and heavy REEs (Table 1).
Eu anomalies in the Badaowan Formation samples are generally subtle, with most showing positive anomalies. The presence of plagioclase and heavy minerals in the study area likely contributes to these positive anomalies. The δCe values, ranging from 0.96 to 1.01, indicate a reducing environment during sediment deposition, consistent with Ceanom indices above −0.1 (indicating anoxic conditions).
The REE distribution patterns for C1 and C3 samples in the Badaowan Formation are similar (Figure 6), showing enrichment of light REEs and depletion of heavy REEs, with low intra-group differentiation [94,95]. Notably, samples from the second member exhibit a distinct positive Eu anomaly, likely attributable to the presence of plagioclase and heavy minerals [94,95].
The LaN/YbN ratio represents the slope of the rare earth element (REE) distribution curve normalized to the North American Shale Composite (NASC) and reflects the degree of fractionation between light and heavy REEs. Specifically, LaN/SmN and GdN/YbN indicate the fractionation of light and heavy REEs, respectively [92,93]. A higher LaN/SmN value suggests enrichment in light REEs, while a lower value indicates a greater degree of heavy REE enrichment. In our study, both the research area and the western ZH1 sampling site exhibit LaN/YbN ratios close to 1, with similar mean LaN/SmN values. Additionally, the GdN/YbN ratios for both regions are approximately 1, suggesting minimal internal fractionation of REEs (Table 2). Thus, the possibility of influence from the western provenance cannot be excluded.
Since Eu commonly exists as Eu2+ under certain geochemical conditions, it often undergoes fractionation from other trivalent REEs in geological systems [94,95]. This leads to either an enrichment (“positive Eu anomaly”) or depletion (“negative Eu anomaly”) of Eu in REE-normalized patterns [96,97,98,99]. The δEu values of the Badaowan Formation samples in our study area generally cluster around 1, though minor variations are observed among individual samples. Most samples exhibit a negative Eu anomaly, while a few display a positive Eu anomaly. The presence of hydrothermal fluids, plagioclase, and heavy minerals can all contribute to a positive Eu anomaly. Considering the regional geological background, the occurrence of plagioclase and heavy minerals is likely the primary factor responsible for the observed positive Eu anomalies.
Ce is highly sensitive to redox conditions. Under strongly oxidizing conditions, Ce fractionates from other trivalent REEs [96,97,100,101]. The Ce anomaly (δCe) is calculated as δCe = Cen/Ce*, where Ce* represents the expected Ce value derived from NASC-normalized La and Nd concentrations. In the research area, δCe values range from 0.96 to 1.01 in C1 samples (mean = 0.98), 0.99 to 1.00 in C3 samples (mean = 1.00), and 0.97 in Zh2 samples. These values suggest that both the study area and the western region predominantly reflect a reducing depositional environment.
The Ce anomaly index (Ceanom) serves as a proxy for evaluating the redox conditions of paleo-water bodies. It is calculated using the formula: Ceanom = lg[3CeN/(2LaN + NdN)], where N represents NASC-normalized values. A Ceanom value < −0.1 indicates Ce depletion, reflecting an oxidizing environment, whereas a Ceanom value > −0.1 signifies Ce enrichment, suggesting a suboxic to anoxic environment. The Ceanom values for the Badaowan Formation samples from both the study area and the western region are consistently greater than −0.1, indicating that the depositional environment was predominantly anoxic.
In contrast, the western region samples exhibit a flatter REE distribution pattern, characterized by light REE depletion and heavy REE enrichment, indicative of distinct provenance differences between the study area and the western region [98,99]. This disparity highlights variations in source material across the basin, with a general east-to-west similarity in peak values.

5. Discussion

5.1. Provenance Analysis of the Lower Jurassic in the Dongdaohaizi Depression

The Kelameili Mountains, located in the eastern Dongdaohaizi Depression, constitute a collisional orogenic belt formed following the subduction and closure of the Paleo-Kelameili Ocean [102,103,104]. This orogenic belt, which runs parallel to the Aermantai ophiolite belt, is structurally controlled by the NWW-oriented Kelameili regional deep fault. The lithological assemblage of the Kelameili orogenic belt primarily consists of intermediate–acidic to acidic eruptive rocks and volcaniclastic rocks. During the formation of the orogenic belt, extensive intermediate–acidic to felsic volcanic activity generated significant quantities of minerals such as zircon and tourmaline. The belt preserves remnants of ophiolite sequences derived from the ancient oceanic crust, collectively forming the Kelameili ophiolite belt in East Junggar. The ophiolite fragments are tectonically intercalated within Devonian and Carboniferous strata [105,106,107,108]. The ophiolite suite is composed of metamorphosed peridotites, gabbro, diabase, and mafic lavas. During the erosion of these rocks, a substantial amount of chromian spinel was produced [81]. Concurrently, with the subduction of the Paleo-Kelameili oceanic crust, minerals within the strata were segregated and transported into the basin under the multi-stage evolution. The heavy mineral assemblages derived from these processes align closely with the results of heavy mineral analysis of the Lower Jurassic Badaowan Formation sandstones in the Dongdaohaizi Depression in this study [109,110].
Among these minerals, SHRIMP U-Pb isotopic dating of zircons from the ophiolite suite yielded an age of 373 Ma [111]. The gabbro within the ophiolite has an age of 329.9 ± 1.6 Ma (LA-ICP-MS zircon U-Pb) [112], and U-Pb dating of zircons from quartz diorite yielded ages ranging from 357 to 492 Ma [113]. Similarly, U-Pb geochronological and fission-track studies from the Kelameili Orogenic Belt also provide evidence that collisional orogeny had already occurred in the Kelameili region prior to the Carboniferous [114,115,116,117]. These findings confirm that the Kelameili Mountains formed before the deposition of the Lower Jurassic Badaowan Formation, providing a source for the transportation of detrital minerals. Additionally, geochemical analyses of basalt samples from the east of Junggar Basin reveal insignificant LREE and HREE fractionation, weak Eu anomalies, and enrichment in HFSE (high-field-strength elements) [118]. The magmas are thought to have originated from a depleted mantle source metasomatized by subduction fluids, underwent mineral fractional crystallization in deep magma chambers, and experienced crustal contamination during emplacement [118]. These characteristics closely align with the geochemical results of the Badaowan Formation sandstones in the Dongdaohaizi Depression.
During the Early Jurassic, the depositional environment of the Badaowan Formation was characterized by a fluvial–lacustrine system under a humid climate, which facilitated the transport of sediments from the Kelameili region to the Dongdaohaizi Depression [119,120]. The uplift and erosion of the Kelameili Mountain provided a continuous sediment supply, while regional tectonic subsidence in the Dongdaohaizi Depression created accommodation space for deposition [1,121]. Seasonal variations in hydrodynamic conditions likely influenced sediment dispersal patterns, with higher-energy fluvial systems transporting coarser materials and lower-energy lacustrine settings favoring finer-grained deposition. This interplay between sediment supply, basin subsidence, and hydrodynamic sorting controlled the provenance signal and stratigraphic architecture of the Badaowan Formation [122].
The Ti-bearing heavy minerals in the sandstones most likely originated from mafic igneous rocks and oceanic plateau fragments [123]. Volcanic activity during the Permian Batamayineishan Formation could have contributed Ti-bearing heavy minerals to the sedimentary record. Furthermore, recent studies suggest that oceanic plateau fragments also supplied a significant amount of Ti-bearing heavy minerals, which is consistent with our interpretation that the sediments primarily originated from the Kelameili Mountains [124]. Multiple studies have demonstrated that the Kelameili region was once part of the Paleo-Kelameili Ocean, and subduction-related processes may have left behind debris from this tectonic setting. According to the law of superposition, the incorporation of detrital minerals from older formations into younger sedimentary strata requires uplift and erosion. Geological surveys have confirmed extensive exposures of Permian volcanic rocks and pre-Permian marine strata in the Kelameili region. During the evolution of the Junggar Basin, the surrounding orogenic belts underwent compressional uplift, providing a major source of detrital material to the basin. This tectonic history supports our conclusion that the provenance of the Dongdaohaizi Depression is linked to the orogenic belt along the basin’s margin.
The Hongshan and Miaogou igneous complexes are located to the west of the Darbut Fault zone (Zhayier Mountain) at the western margin of the Junggar Basin. These complexes are dominated by volcanic rocks such as boninite-like basaltic andesite and pyroxene andesite [29,125,126,127]. Geochemical studies of the western Junggar region indicate that the total rare earth element (REE) contents (ΣREE) range from 195.16 ppm to 249.18 ppm, with light rare earth elements (LREEs) ranging from 125.76 ppm to 170.99 ppm and heavy rare earth elements (HREEs) from 69.4 ppm to 80.83 ppm [128,129,130,131,132]. The (La/Yb)N ratio ranges from 3.25 to 4.69, indicating significant LREE–HREE fractionation. The (La/Sm)N values exceed 1, ranging from 2.12 to 3.28 with an average of 2.55, reflecting well-developed LREE fractionation. The (Gd/Yb)N ratio ranges from 1.00 to 1.31, suggesting no significant HREE fractionation. Trace element analysis shows weak negative Eu anomalies, with REE patterns normalized to chondrites displaying steep LREE segments and flat HREE segments, forming a right-leaning “V” shape [29,125,126,127]. These volcanic rocks are enriched in LREEs and large-ion lithophile elements (LILEs) such as Rb, K, Ba, U, and Sr, while being relatively depleted in HFSEs such as Th, Nb, Ce, and P, as well as HREEs [29,125,126,127]. These features differ from the distinct positive Eu anomalies observed in the study area.
Based on previous studies, the heavy mineral assemblages of the Tianshan (Yilinhebiergen Mountain) and Bogda regions are significantly different from those found in the Badaowan Formation sandstones, making it unlikely that sediments were derived from southern Junggar Basin (Figure 1b) [114,119,133,134,135]. Additionally, the formation of the central uplift of the Junggar Basin during the Jurassic period effectively blocked the transport of sediments from the Altai (Qinggelidi) region, northern Junggar Basin (Figure 1b) [37,136].
By comparing the petrographic and geochemical characteristics of the Badaowan Formation sandstones with those of the eastern and western igneous complexes, the sandstones of the Lower Jurassic Badaowan Formation in the Dongdaohaizi Depression exhibit greater similarity to the rocks of the Kelameili Mountains in the east. This suggests that the Kelameili region was the primary sediment source for the study area. However, given that the geochemical characteristics of the samples from sampling point Zh1 in this study exhibit some similarities with those of the western region, it is inferred that the study area shows minimal internal LREE differentiation, lacks significant HREE differentiation compared to the western region, and reflects an anoxic reducing environment. Thus, the possibility of long-distance sediment transport influence from the western provenance cannot be excluded.

5.2. Tectonic Setting of the Provenance

Previous studies have statistically analyzed the REE (rare earth element) parameters of graywackes formed under various tectonic settings [68,71,137]. A comparison between these statistics and the parameters obtained in this study is shown in Table 2. This table systematically reveals the relationship between REE distribution characteristics and the tectonic setting of sedimentary basins, as well as the provenance type [138]. The comparison reveals that, based on the abundances of La and Ce, as well as the total REE content, the samples from C1 and C3 in the study area most closely resemble those from continental island arc settings. This suggests that the tectonic setting of the provenance in the study area likely corresponds to an evolutionary stage of a continental island arc.
Trace elements in terrigenous clastic rocks, including some REEs, exhibit high stability. Elements such as La, Th, Ti, Zr, and Sc are minimally affected by geological processes during weathering, transport, and sedimentation [139,140]. Variations in their concentrations are inherently linked to tectonic settings, making them reliable indicators of the tectonic environment and evolutionary characteristics of the provenance area [98,99,141].
Based on the trace element data from the samples analyzed in this study, tectonic environment discrimination diagrams were constructed (Figure 7 and Table A4) [71,142,143,144,145]. The La-Th-Sc triangular discrimination diagram shows that most sample data points cluster in the continental island arc field. Similarly, the Th-Co-Zr/10 diagram indicates that the majority of samples also fall within the continental island arc field. The Th-Sc-Zr/10 diagram further supports this finding, with most points in the continental island arc field and a few in adjacent indeterminate zones. In contrast, the La/Y-Sc/Cr discrimination diagram reveals that the majority of the sample points plot in the passive continental margin field, with a few falling into neighboring uncertain zones.
Table 2. Comparison of REE parameters of sedimentary rocks in different tectonic settings.
Table 2. Comparison of REE parameters of sedimentary rocks in different tectonic settings.
Type/Sample and StrataLaCeΣREELa/YbLaN/YbNLREE/HREEδEu
Tectonic settingOIA819584.22.83.81.04
CIA2759146117.57.70.79
ACM377818612.58.59.10.6
PM398521015.910.88.50.56
Sample and strataC1-J1b340.582.2175.8312.770.947.581.06
C1-J1b234.069.6148.5010.630.788.412.86
C3-J1b215.731.277.7813.170.977.092.51
C3-J1b118.638.894.8810.220.757.251.23
δEu = 2 × EuN/(SmN + GdN); δCe = 2 × CeN/(LaN + PrN). ACM = active continental margin, PM = passive continental margin, CIA = continental island arc, OIA = oceanic island arc. Data of tectonic setting are from reference [138].
The results from these four diagrams show good consistency in identifying the tectonic setting of the provenance area, confirming that the Kelameili region is associated with both passive continental margin and continental island arc tectonic settings.
The principle behind this approach is that geochemical data represent a cumulative record of tectonic events over long timescales [146,147,148]. In the Lower Jurassic Badaowan Formation sandstones from the Dongdaohaizi Depression, the clastic material reflects a composite record of all tectonic events prior to the Jurassic. The proportion of lithic fragments from the Badaowan period (~10 Ma) is not necessarily greater than that from the Paleozoic to Early Jurassic (~300 Ma). Additionally, volcanic and igneous activities persisted during the Jurassic period, aligning the tectonic environment of the Junggar Basin more closely with an active continental margin [149,150]. However, our geochemical results indicate that the eastern Junggar Basin is characterized by typical passive continental margin and island arc signatures. Therefore, we suggest that these geochemical signatures reflect the tectonic setting prior to the Jurassic. While tectonic and volcanic factors during the Jurassic may have influenced the composition and origin of the sediments, our data reliably document two distinct stages of tectonic settings before the Jurassic.

5.3. Evolution and Mechanisms of the Kelameili Ocean

The tectonic evolution of the Kelameili region, located in the eastern Junggar Basin, reflects significant changes in its tectonic setting across various geological periods. This region underwent a transition from a passive continental margin to a continental island arc, demonstrating the intricate interplay between crustal evolution and plate tectonics. A comprehensive analysis of magmatic activity, sedimentary sequences, geochemical characteristics, and tectonic deformation in the Kelameili region provides a clearer timeline of its tectonic evolution and insights into its formation mechanisms [108,151].
The Kelameili region exhibited characteristics of a passive continental margin during the Middle Ordovician to Late Silurian period (~466–420 Ma) [114,152,153,154,155]. During this time, the opening and extension of a branch basin of the Paleo-Asian Ocean led to a stable tectonic regime typical of a passive continental margin. By the end of the Silurian, the breakup of the ocean basin facilitated the formation of new oceanic crust in the eastern Junggar Basin, establishing a passive continental margin depositional environment. Magmatic activity during this period was limited, with igneous rocks such as granite and basalt displaying geochemical signatures indicative of rifting and thermal melting, including low Sr/Y ratios [155,156,157]. These lithological and geochemical features suggest that the Kelameili region was situated in a relatively stable tectonic setting, distinct from island arcs or active continental margins.
With the onset of subduction of the Paleo-Asian Ocean (Kelameili Ocean), the eastern Junggar Basin transitioned from a passive continental margin to a region exhibiting island arc characteristics [154,158,159,160]. This transformation occurred during the Middle Devonian to Early Carboniferous period (~420–360 Ma), driven by the bidirectional episodic subduction of the Kelameili Ocean. Extensive island arc volcanic rocks, including basalt, andesite, and associated lithological assemblages, were formed during this period [114,161]. The progressive subduction and consumption of the Kelameili Ocean led to significant tectonic changes, marking the region’s evolution into a classic continental island arc setting.
Detrital zircon age peaks exhibit a clear concentration in the Permian–Triassic period, primarily reflecting sources from Late Permian to Early Triassic igneous or metamorphic rocks [21,23,157]. This suggests highly active tectonic activities during this period, particularly collision-related orogenesis, which led to significant magmatic activity and tectonic deformation. These findings indicate that the tectonic environment of the Junggar Basin transitioned from a passive margin to an active orogenic belt during the Permian–Triassic period.
The transformation of the Kelameili region from a passive continental margin to a continental island arc was closely linked to the subduction process of the Kelameili Ocean. Research suggests that the island arc features in the Kelameili region were primarily influenced by episodic or bidirectional subduction of the oceanic basin [154,158,159]. As the Kelameili Ocean subducted and its crust was consumed, partial melting of the mantle wedge generated large volumes of volcanic rocks enriched in aluminum, calcium, and magnesium, particularly high-alumina basalts and andesites. The formation of these magmatic rocks signaled the region’s transition into an island arc tectonic environment [162,163,164].
During this process, the tectonic setting of the Kelameili region shifted from a stable passive continental margin to an active continental margin influenced by plate subduction. Frequent and extensive magmatic activity further promoted the formation of island arc volcanism and uplift of the island arc tectonic belt. The interaction between crust and mantle not only intensified magmatic activity but also contributed to crustal thickening, shear deformation, and fold development in the region [162]. Elemental ratios highlight the magma characteristics resulting from partial melting of the mantle wedge [162]. Additionally, the high concentrations of calcium, aluminum, and magnesium in the volcanic rocks provide further evidence that the magma source was primarily derived from a subduction-affected mantle wedge [108,165].
In summary, the tectonic evolution of the Kelameili region in the eastern Junggar Basin involved a transition from a passive continental margin to a continental island arc, culminating in its collision and amalgamation with the Eurasian continent, forming an orogenic belt. This tectonic evolution not only intensified magmatic activity but also contributed to the development of uplifts, depressions, and other regional tectonic features within the Junggar Basin, providing an abundant supply of clastic sediments for basin development [109,153,154].

6. Conclusions

  • From bottom to top, the Badaowan Formation in the Dongdaohaizi Depression exhibits a decrease in lithic fragment content and an increase in quartz content, indicating an improvement in compositional maturity. The source rocks in the study area primarily include tuff, granite, basalt, and minor metamorphic rocks. The heavy mineral assemblage, dominated by zircon, chrome-spinel, tourmaline, and garnet, suggests that the source rocks are mainly intermediate to acidic magmatic rocks, basic magmatic rocks, and metamorphic rocks. The provenance of the Badaowan Formation in the study area is predominantly controlled by the Kelameili Mountain to the east, which continued to uplift and migrate during the Early Jurassic.
  • LA-ICP-MS results show that the rare earth element (REE) content of Lower Jurassic sandstones is slightly lower than the average REE content of the upper continental crust. The REE distribution exhibits weak differentiation between light and heavy rare earth elements. This geochemical signature, coupled with the overall reducing and anoxic depositional environment, reflects the passive continental margin and continental island arc tectonic settings of the source region.
  • As part of the Paleo-Asian Ocean system, the Kelameili Ocean underwent a complex evolution involving subduction and closure within a framework of multiple oceanic basins and microcontinents. During the Middle Ordovician to Late Silurian, the Kelameili region developed a passive continental margin under conditions of limited ocean basin expansion. From the Middle Devonian to Early Carboniferous, the onset of subduction of the Kelameili Ocean marked a transition in the eastern Junggar Basin from a passive continental margin to a continental island arc setting. Transition may have been caused by the episodic subduction or bidirectional subduction of the Paleo-Kelameili Ocean.

Author Contributions

Conceptualization, T.W. and F.L.; methodology, Z.Z.; software, C.Z.; validation, J.H.; investigation, J.L. and X.C.; resources, X.T. and C.S.; visualization, Y.G.; writing—original draft preparation, F.L.; writing—review and editing, T.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Qingchuang Science and Technology Plan of Shandong Higher Education Institution (2024KJG073), Funding for Basic Scientific Research Expenses for Universities Directly Under the Autonomous Region (JY20250015), and National Science and Technology Major Project (2024ZD003403).

Data Availability Statement

Research data are provided in the article and Appendix A.

Acknowledgments

The authors thank the anonymous reviewers for their constructive comments.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. Statistics of clastic materials of sandstones from the Badaowan Formation.
Table A1. Statistics of clastic materials of sandstones from the Badaowan Formation.
StrataSampleQ (%)F (%)R (%)Igneous Rocks (%)Metamorphic Rocks (%)Sedimentary Rocks (%)Maturity
J1b3C1-0156.013.830.274.019.06.01.27
C1-0256.013.830.274.020.95.11.27
C1-0356.013.830.266.617.116.31.27
C1-0456.013.830.266.617.116.31.27
C1-0561.615.223.274.019.07.01.60
C1-0650.415.234.474.017.18.91.02
C1-0750.415.234.474.020.95.11.02
C1-0861.613.824.666.620.912.51.60
C1-0961.613.824.666.619.014.41.60
C1-1050.413.835.866.620.912.51.02
J1b2C3-0145.517.736.968.014.018.00.83
C3-0250.019.430.661.212.626.21.00
C3-0340.917.741.461.214.024.80.69
C3-0440.915.943.268.012.619.40.69
C3-0540.919.439.661.214.024.80.69
C3-0650.017.732.368.012.619.41.00
C3-0750.019.430.661.214.024.81.00
C3-0840.917.741.461.215.423.40.69
C3-0940.917.741.474.812.612.60.69
C3-1045.517.736.968.014.018.00.83
J1b1C3-0146.812.740.571.019.010.00.88
C3-0251.512.735.871.020.98.11.06
C3-0346.814.039.271.020.98.10.88
C3-0442.111.546.478.117.14.80.73
C3-0542.114.043.978.120.91.00.73
C3-0651.512.735.871.017.111.91.06
C3-0742.114.043.978.117.14.80.73
C3-0851.512.735.863.917.119.01.06
C3-0946.812.740.571.017.111.90.88
C3-1046.814.039.263.920.915.20.88
Table A2. Relative content of heavy minerals of sandstones from the Badaowan Formation.
Table A2. Relative content of heavy minerals of sandstones from the Badaowan Formation.
StrataSample Heavy Mineral (%)
ZirconGarnetChrome-spinelTourmalineApatiteRutileAnataseLeucoxeneOther
J1b3C1-0137.7311.0221.6120.100.500.482.310.495.75
C1-0239.0911.7521.5219.720.470.492.340.494.13
C1-0338.7911.1921.3518.750.470.482.500.495.99
C1-0439.3411.5920.4820.680.480.462.440.454.05
C1-0540.4211.8221.7019.790.500.492.370.532.39
C1-0636.1411.7020.8719.900.500.482.450.497.49
C1-0736.9212.0822.9319.320.490.462.250.485.07
C1-0838.1412.3220.9819.000.470.482.470.495.66
C1-0940.0610.9421.2619.900.500.472.390.474.01
C1-1037.5611.9421.5819.530.490.492.340.485.58
Average38.4211.6421.4319.670.480.482.390.495.01
C3-0160.070.0427.044.820.370.873.420.353.03
C3-0257.960.0427.554.640.350.833.330.364.94
C3-0362.970.0426.385.050.350.833.270.360.75
C3-0458.380.0426.664.740.360.813.540.375.10
C3-0561.670.0428.534.710.380.873.430.372.25
C3-0661.130.0426.784.720.360.843.190.352.58
C3-0759.360.0427.394.700.380.833.250.353.69
C3-0860.200.0427.514.830.360.813.210.372.67
C3-0961.220.0424.654.730.360.803.510.354.34
C3-1063.420.0426.694.670.340.883.160.360.56
Average60.640.0426.924.760.360.843.330.362.43
Zh1-0136.8541.386.518.060.541.533.120.461.53
Zh1-0237.8040.946.387.590.551.603.300.471.37
Zh1-0336.0238.956.507.390.581.533.080.495.47
Zh1-0438.8341.126.387.630.541.673.180.480.18
Zh1-0538.3639.766.547.480.581.733.400.471.67
Zh1-0638.2442.446.456.940.551.593.070.460.26
Zh1-0736.3239.866.217.440.561.523.290.494.32
Zh1-0838.4442.746.367.420.571.593.200.480.80
Zh1-0937.5942.096.627.420.591.663.310.490.23
Zh1-1036.5442.616.397.690.541.633.400.470.72
Average37.5041.196.437.510.561.603.230.481.50
Table A3. Analysis data of rare earth elements of sandstones from the Badaowan Formation.
Table A3. Analysis data of rare earth elements of sandstones from the Badaowan Formation.
StrataSampleElement (ppm)
LaCePrNdSmEuGdTbDyHoErTmYbLu
J1b3C1-0140.4682.6810.3839.427.581.616.590.935.691.043.290.483.230.50
C1-0240.7384.769.9136.117.701.686.020.915.291.023.100.473.090.49
C1-0340.9683.1610.7137.078.091.606.310.925.381.013.250.433.420.49
C1-0439.6382.7010.6537.767.721.645.930.915.661.033.150.453.280.50
C1-0540.0274.6710.3838.218.171.656.600.885.421.073.320.463.050.50
C1-0639.1679.7010.1538.477.681.586.350.985.861.003.250.503.230.48
C1-0741.1283.449.8638.838.321.536.250.865.371.063.210.473.170.51
C1-0841.1879.0210.5537.397.671.496.770.955.571.023.150.463.330.48
C1-0939.6885.4910.3539.527.481.496.190.875.501.023.130.463.320.49
C1-1039.5184.4510.1836.837.541.506.410.905.501.003.160.473.230.46
Average40.2582.0110.3137.967.791.586.340.915.521.033.200.463.240.49
J1b2C1-1134.3668.728.4130.807.303.956.220.895.311.033.170.483.050.44
C1-1234.1566.348.3131.867.533.696.160.865.261.063.290.453.130.52
C1-1335.5269.198.3831.377.093.966.100.865.411.023.160.433.120.48
C1-1433.5173.417.7033.917.534.376.080.885.141.013.300.492.990.50
C1-1533.9572.058.3934.017.343.695.420.925.201.053.220.483.230.48
C1-1634.0169.408.3833.747.244.045.950.905.401.063.080.513.280.45
C1-1735.1573.628.2332.257.103.926.070.895.081.002.970.453.210.47
C1-1835.1768.518.4232.647.363.825.600.935.341.033.170.473.250.48
C1-1933.7267.568.4631.507.274.225.710.905.391.093.110.443.130.48
C1-2031.5470.148.0533.167.434.135.930.925.711.143.220.493.120.48
Average34.1169.908.2732.537.323.985.920.905.321.053.170.473.150.48
J1b2C3-0115.6931.323.6612.672.931.713.140.462.570.481.360.171.230.18
C3-0216.1830.273.5713.413.021.592.920.472.570.481.370.181.200.17
C3-0315.1930.903.5913.772.941.583.050.462.660.521.310.181.190.17
C3-0415.6432.143.5713.622.821.583.220.442.530.491.370.181.210.16
C3-0514.9031.773.3213.173.031.652.880.432.560.481.460.181.180.17
C3-0615.2729.823.3313.752.601.582.990.482.640.511.510.181.180.17
C3-0715.2330.833.7913.002.861.593.120.432.550.481.360.181.120.16
C3-0815.9530.643.4914.172.951.503.100.442.700.511.310.181.170.18
C3-0916.0231.253.4113.443.041.533.310.462.570.501.440.181.210.17
C3-1016.4330.603.6713.122.861.563.210.432.720.511.380.171.210.17
Average15.6530.963.5413.412.911.593.090.452.610.491.390.181.190.17
J1b1C3-1119.1839.724.6517.303.340.823.400.492.950.671.720.281.710.27
C3-1217.4439.024.5618.003.160.853.100.493.140.691.810.281.800.26
C3-1318.6538.384.6016.833.260.813.230.492.990.681.760.271.820.26
C3-1418.8437.194.7916.573.330.873.310.532.990.641.710.281.830.27
C3-1518.7737.724.7716.653.340.843.200.502.910.721.760.281.850.26
C3-1618.7137.994.7318.363.320.822.940.503.140.691.740.291.790.25
C3-1718.0839.384.3716.803.370.763.220.522.940.711.820.271.890.27
C3-1818.9138.744.7317.373.130.823.170.512.920.681.740.281.860.27
C3-1919.4538.464.5417.963.260.833.140.483.140.671.820.281.770.28
C3-2018.4639.174.7017.383.190.833.300.472.930.681.760.291.920.28
Average18.6538.584.6417.323.270.823.200.503.000.681.760.281.820.27
J1b3Zh1-0121.8245.165.7721.674.000.953.880.593.500.752.070.332.040.32
Zh1-0222.5145.535.2620.774.160.953.750.583.570.772.110.312.120.33
Zh1-0321.8945.485.5720.943.900.923.750.633.570.722.250.322.070.34
Zh1-0423.5244.795.6721.694.020.953.680.593.720.742.180.352.180.31
Zh1-0521.5346.915.4920.374.140.953.730.613.730.722.190.322.120.34
Zh1-0621.1545.075.5621.023.940.973.840.603.750.702.070.322.050.33
Zh1-0724.3145.345.6321.174.371.003.770.593.510.752.140.322.090.34
Zh1-0823.6646.945.5021.114.160.963.630.623.470.772.030.332.250.33
Zh1-0923.1447.185.7521.644.200.963.600.593.690.752.100.312.120.34
Zh1-1022.2247.195.6521.923.940.923.640.613.720.782.180.322.000.32
Average22.5745.965.5921.234.080.953.730.603.620.752.130.322.100.33
Table A4. Geochemical analysis data of trace elements of sandstones from the Badaowan Formation.
Table A4. Geochemical analysis data of trace elements of sandstones from the Badaowan Formation.
StrataSampleElement (ppm)
LaThScCoZr/10Sc/CrLa/Y
J1b3 0.520.170.270.240.520.131.23
C1-020.590.170.280.330.510.171.21
C1-030.590.170.240.220.500.171.17
C1-040.590.180.230.230.510.161.19
C1-050.600.180.280.250.510.171.23
C1-060.590.180.240.280.490.151.24
C1-070.560.180.280.310.510.201.14
C1-080.550.180.240.270.510.151.29
C1-090.620.170.300.260.520.181.12
C1-100.580.170.270.280.490.151.31
J1b2C1-110.560.190.300.270.510.141.23
C1-120.570.180.260.210.500.171.31
C1-130.550.180.270.290.470.181.28
C1-140.540.160.300.320.530.141.13
C1-150.630.160.260.280.510.201.18
C1-160.550.160.270.310.480.171.37
C1-170.560.180.230.270.510.181.14
C1-180.600.190.290.310.460.171.30
C1-190.580.180.290.320.500.141.09
C1-200.560.180.270.330.510.141.15
J1b1C3-010.590.160.270.320.500.161.18
C3-020.610.180.310.260.520.161.17
C3-030.570.170.310.260.510.171.23
C3-040.610.170.280.260.490.131.21
C3-050.590.170.300.260.500.151.12
C3-060.530.160.260.250.480.181.24
C3-070.630.190.290.280.500.151.25
C3-080.590.170.280.240.490.161.11
C3-090.540.170.270.250.500.161.20
C3-100.570.170.330.270.480.171.24

References

  1. Ashraf, A.R.; Sun, Y.; Sun, G.; Uhl, D.; Mosbrugger, V.; Li, J.; Herrmann, M. Triassic and Jurassic palaeoclimate development in the Junggar Basin, Xinjiang, Northwest China—A review and additional lithological data. Palaeobiodiversity Palaeoenvironments 2010, 90, 187–201. [Google Scholar] [CrossRef]
  2. Fu, H.; Tang, D.; Xu, H.; Xu, T.; Chen, B.; Hu, P.; Yin, Z.; Wu, P.; He, G. Geological characteristics and CBM exploration potential evaluation: A case study in the middle of the southern Junggar Basin, NW China. J. Nat. Gas Sci. Eng. 2016, 30, 557–570. [Google Scholar] [CrossRef]
  3. Xin, C.; Hua-fu, L.U.; Liang-shu, S.H.U.; Hui-min, W.; Guo-qing, Z. Study on tectonic evolution of Junggar Basin. Geol. J. China Univ. 2002, 8, 257. [Google Scholar]
  4. Lawrence, S.R. Aspects of the petroleum geology of the Junggar basin, Northwest China. Geol. Soc. Lond. Spec. Publ. 1990, 50, 545–557. [Google Scholar] [CrossRef]
  5. Wang, Y.; Jia, D.; Pan, J.; Wei, D.; Tang, Y.; Wang, G.; Wei, C.; Ma, D. Multiple-phase tectonic superposition and reworking in the Junggar Basin of northwestern China—Implications for deep-seated petroleum exploration. AAPG Bull. 2018, 102, 1489–1521. [Google Scholar] [CrossRef]
  6. Weifeng, W.; Yequan, C. Tectonic evolution and petroleum systems in the Junggar Basin. Acta Geol. Sin.-Engl. Ed. 2004, 78, 667–675. [Google Scholar] [CrossRef]
  7. Zhou, J.-B.; Wilde, S.A.; Zhao, G.-C.; Han, J. Nature and assembly of microcontinental blocks within the Paleo-Asian Ocean. Earth-Sci. Rev. 2018, 186, 76–93. [Google Scholar] [CrossRef]
  8. Wan, B.; Li, S.; Xiao, W.; Windley, B.F. Where and when did the Paleo-Asian ocean form? Precambrian Res. 2018, 317, 241–252. [Google Scholar] [CrossRef]
  9. Dobretsov, N.L.; Berzin, N.A.; Buslov, M.M. Opening and tectonic evolution of the Paleo-Asian Ocean. Int. Geol. Rev. 1995, 37, 335–360. [Google Scholar] [CrossRef]
  10. Gao, Y.; Zhang, G.; Li, S.; Guo, R.; Zeng, Z.; Cheng, S.; Xue, Z.; Li, L.; Zhou, H.; Liu, S. Provenance of the Lower Jurassic Badaowan and Sangonghe Formations in Dongdaohaizi Depression, Junggar Basin, and Its Constraint on the Karamaili Ocean. J. Mar. Sci. Eng. 2023, 11, 1375. [Google Scholar] [CrossRef]
  11. Yanping, L.I.; Hongliang, Z.O.U.; Lei, L.I.; Jiyou, F.U.; Yu, X.I.A.; Junyang, X.I.E. Petroleum exploration ideas and discoveries in upper Wuerhe Formation, Dongdaohaizi sag, Junggar Basin. Xinjiang Pet. Geol. 2022, 43, 127. [Google Scholar]
  12. Shengyu, Y.I.N.; Deyu, G.; Jing, S.U.; Zhengzhou, H.U.; Wei’an, W.U.; Ruiju, W. Genetic Types and Origins of Upper Paleozoic-reservoiered Oil in Dongdaohaizi Sag and Its Surrounding Areas in the Junggar Basin. Geol. J. China Univ. 2024, 30, 218. [Google Scholar]
  13. Lichun, K.; Yong, T.; Dewen, L.; Tao, W.; Jianhua, Q. Exploration of fan-controlled large-area lithologic oil reservoirs of Triassic Baikouquan Formation in slope zone of Mahu Depression in Junggar Basin. China Pet. Explor. 2014, 19, 14. [Google Scholar]
  14. Chen, J.; Pang, X.; Wang, X.; Wang, Y. A new method for assessing tight oil, with application to the Lucaogou Formation in the Jimusaer depression, Junggar Basin, China. AAPG Bull. 2020, 104, 1199–1229. [Google Scholar] [CrossRef]
  15. Bian, W.; Hornung, J.; Liu, Z.; Wang, P.; Hinderer, M. Sedimentary and palaeoenvironmental evolution of the Junggar Basin, Xinjiang, northwest China. Palaeobiodiversity Palaeoenvironments 2010, 90, 175–186. [Google Scholar] [CrossRef]
  16. Zhu, X.; Li, S.; Wu, D.; Zhu, S.; Dong, Y.; Zhao, D.; Wang, X.; Zhang, Q. Sedimentary characteristics of shallow-water braided delta of the Jurassic, Junggar basin, Western China. J. Pet. Sci. Eng. 2017, 149, 591–602. [Google Scholar] [CrossRef]
  17. Fang, Y.; Wu, C.; Wang, Y.; Wang, L.; Guo, Z.; Hu, H. Stratigraphic and sedimentary characteristics of the Upper Jurassic-Lower Cretaceous strata in the Junggar Basin, Central Asia: Tectonic and climate implications. J. Asian Earth Sci. 2016, 129, 294–308. [Google Scholar] [CrossRef]
  18. Feng, Y.; Jiang, S.; Wang, C. Sequence stratigraphy, sedimentary systems and petroleum plays in a low-accommodation basin: Middle to upper members of the Lower Jurassic Sangonghe Formation, Central Junggar Basin, Northwestern China. J. Asian Earth Sci. 2015, 105, 85–103. [Google Scholar] [CrossRef]
  19. Fang, Y.; Wu, C.; Guo, Z.; Hou, K.; Dong, L.; Wang, L.; Li, L. Provenance of the southern Junggar Basin in the Jurassic: Evidence from detrital zircon geochronology and depositional environments. Sediment. Geol. 2015, 315, 47–63. [Google Scholar] [CrossRef]
  20. Wang, J.; Wu, C.; Jiao, Y.; Yuan, B. Middle–Late Triassic sedimentary provenance of the southern Junggar Basin and its link with the post-orogenic tectonic evolution of Central Asia. Sci. Rep. 2021, 11, 17041. [Google Scholar] [CrossRef]
  21. Yao, Z.; Yu, H.; Yang, F.; Jianatayi, D.; Zhang, B.; Li, T.; Jia, C.; Pan, T.; Zhang, Z.; Aibibuli, N. Provenance, Depositional Environment, and Paleoclimatic Conditions of a Near-Source Fan Delta: A Case Study of the Permian Jiamuhe Formation in the Shawan Sag, Junggar Basin. Minerals 2023, 13, 1251. [Google Scholar] [CrossRef]
  22. Li, X.; Chen, G.; Wu, C.; Li, J.; Haproff, P.J.; Geng, M.; Wu, S.; Xu, S.; Li, Z.; Yang, D. Tectono-stratigraphic framework and evolution of East Junggar Basin, Central Asia. Tectonophysics 2023, 851, 229758. [Google Scholar] [CrossRef]
  23. Hu, X.; Zhang, X.; Xie, J.; Cao, H.; Zheng, X.; Zhao, Z.; Cao, J.; Pu, Q.; Li, Z.; Zhou, L. Sedimentary Characteristics and Hydrocarbon-Generation Potential of the Permian Pingdiquan Formation in Dongdaohaizi Sag, Junggar Basin, Northwest China. ACS Omega 2023, 8, 35653–35669. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, W.; Windley, B.F.; Sun, S.; Li, J.; Huang, B.; Han, C.; Yuan, C.; Sun, M.; Chen, H. A tale of amalgamation of three Permo-Triassic collage systems in Central Asia: Oroclines, sutures, and terminal accretion. Annu. Rev. Earth Planet. Sci. 2015, 43, 477–507. [Google Scholar] [CrossRef]
  25. Gao, Y.; Zhang, S.; Zhao, H.; Ren, Q.; Yang, T.; Wu, H.; Li, H. North China block underwent simultaneous true polar wander and tectonic convergence in late Jurassic: New paleomagnetic constraints. Earth Planet. Sci. Lett. 2021, 567, 117012. [Google Scholar] [CrossRef]
  26. Chew, D.M.; Spikings, R.A. Geochronology and Thermochronology Using Apatite: Time and Temperature, Lower Crust to Surface. Elements 2015, 11, 189–194. [Google Scholar] [CrossRef]
  27. Mao, J.; Pirajno, F.; Lehmann, B.; Luo, M.; Berzina, A. Distribution of porphyry deposits in the Eurasian continent and their corresponding tectonic settings. J. Asian Earth Sci. 2014, 79, 576–584. [Google Scholar] [CrossRef]
  28. Buckman, S.; Aitchison, J.C. Tectonic evolution of Palaeozoic terranes in west Junggar, Xinjiang, NW China. Geol. Soc. Lond. Spec. Publ. 2004, 226, 101–129. [Google Scholar] [CrossRef]
  29. Gao, R.; Xiao, L.; Pirajno, F.; Wang, G.-C.; He, X.-X.; Yang, G.; Yan, S.-W. Carboniferous–Permian extensive magmatism in the West Junggar, Xinjiang, northwestern China: Its geochemistry, geochronology, and petrogenesis. Lithos 2014, 204, 125–143. [Google Scholar] [CrossRef]
  30. Zhu, H.; Wicander, R.; Marshall, J.E.A. Biostratigraphic and paleogeographic significance of a palynological assemblage from the Middle Devonian Ulusubasite Formation, eastern Junggar Basin, Xinjiang, China. Rev. Palaeobot. Palynol. 2008, 152, 141–157. [Google Scholar] [CrossRef]
  31. Li, Z.; Chen, Z.; Cao, F.; Zhang, S.; Huang, N. The study of Yanshanian unconformity tectonic events in Northern Junggar Basin. IOP Conf. Ser. Earth Environ. Sci. 2019, 349, 012045. [Google Scholar] [CrossRef]
  32. Li, Y.-n.; Shao, L.; Hou, H.; Tang, Y.; Yuan, Y.; Zhang, J.; Shang, X.; Lu, J. Sequence stratigraphy, palaeogeography, and coal accumulation of the fluvio-lacustrine Middle Jurassic Xishanyao Formation in central segment of southern Junggar Basin, NW China. Int. J. Coal Geol. 2018, 192, 14–38. [Google Scholar] [CrossRef]
  33. Chen, J.; He, N.; Wang, Y.; Fu, L.; Yu, C.; Liu, Y. The sedimentary characteristics and depositional evolution in the Middle Jurassic Xishanyao Formation, Southeastern Xiayan Rise, Junggar Basin, NW China. Geol. J. 2022, 57, 2221–2234. [Google Scholar] [CrossRef]
  34. Wang, Y.-H.; Xue, C.-J.; Liu, J.-J.; Wang, J.-P.; Yang, J.-T.; Zhang, F.-F.; Zhao, Z.-N.; Zhao, Y.-J.; Liu, B. Early Carboniferous adakitic rocks in the area of the Tuwu deposit, eastern Tianshan, NW China: Slab melting and implications for porphyry copper mineralization. J. Asian Earth Sci. 2015, 103, 332–349. [Google Scholar] [CrossRef]
  35. Pan, Y.; Sha, J.; Wang, Y.; Zhang, X.; Yao, X.; Peng, B.; Rao, X. The brackish-water bivalve Waagenoperna from the Lower jurassic Badaowan Formation of the Junggar Basin and its palaeoenvironmental and palaeogeographic significance. Geosci. Front. 2013, 4, 95–103. [Google Scholar] [CrossRef]
  36. Wu, K.; Paton, D.; Zha, M. Unconformity structures controlling stratigraphic reservoirs in the north-west margin of Junggar basin, North-west China. Front. Earth Sci. 2013, 7, 55–64. [Google Scholar] [CrossRef]
  37. Yang, Y.T.; Song, C.C.; He, S. Jurassic tectonostratigraphic evolution of the Junggar basin, NW China: A record of Mesozoic intraplate deformation in Central Asia. Tectonics 2015, 34, 86–115. [Google Scholar] [CrossRef]
  38. Haughton, P.D.; Todd, S.P.; Morton, A.C. Sedimentary provenance studies. Geol. Soc. Lond. Spec. Publ. 1991, 57, 1–11. [Google Scholar] [CrossRef]
  39. Caracciolo, L. Sediment generation and sediment routing systems from a quantitative provenance analysis perspective: Review, application and future development. Earth-Sci. Rev. 2020, 209, 103226. [Google Scholar] [CrossRef]
  40. Dickinson, W.R.; Suczek, C.A. Plate tectonics and sandstone compositions. Aapg Bull. 1979, 63, 2164–2182. [Google Scholar]
  41. Jian, X.; Guan, P.; Zhang, D.W.; Zhang, W.; Feng, F.; Liu, R.J.; Lin, S.D. Provenance of Tertiary sandstone in the northern Qaidam basin, northeastern Tibetan Plateau: Integration of framework petrography, heavy mineral analysis and mineral chemistry. Sediment. Geol. 2013, 290, 109–125. [Google Scholar] [CrossRef]
  42. Yan, Z.; Wang, Z.; Wang, T.; Yan, Q.; Xiao, W.; Li, J. Provenance and Tectonic Setting of Clastic Deposits in the Devonian Xicheng Basin, Qinling Orogen, Central China. J. Sediment. Res. 2006, 76, 557–574. [Google Scholar] [CrossRef]
  43. Yan, Z.; Wang, Z.; Yan, Q.; Wang, T.; Guo, X. Geochemical Constraints On the Provenance and Depositional Setting of the Devonian Liuling Group, East Qinling Mountains, Central China: Implications for the Tectonic Evolution of the Qinling Orogenic Belt. J. Sediment. Res. 2012, 82, 9–24. [Google Scholar] [CrossRef]
  44. Le Pera, E.; Tangari, A.C.; Marinangeli, L.; Morrone, C.; Riber, L.; Andò, S. Provenance of modern sands from Baja California rivers (Mexico): Petrographic constraints from light and heavy minerals. J. Sediment. Res. 2023, 93, 617–641. [Google Scholar] [CrossRef]
  45. Pasqualone, L.; Brozzetti, F.; Mirabella, F.; Luchetti, L.; Tangari, A.C.; Barchi, M.R. Tectono-stratigraphic evolution of a deep-water foreland basin: A case study from the Marnoso-arenacea basin, central Italy. Ital. J. Geosci. 2024, 143, 105–129. [Google Scholar] [CrossRef]
  46. Garzanti, E. Petrographic classification of sand and sandstone. Earth-Sci. Rev. 2019, 192, 545–563. [Google Scholar] [CrossRef]
  47. Dickinson, W. Interpreting Provenance Relations from Detrital Modes of Sandstones. In Provenance of Arenites; Springer: Berlin/Heidelberg, Germany, 1985. [Google Scholar]
  48. Dickinson, W.R. Provenance of North American Phanerozoic sandstones in relation to tectonic setting. Geol. Soc. Am. Bull. 1983, 94, 222–235. [Google Scholar] [CrossRef]
  49. Ingersoll, R.V.; Bullard, T.F.; Ford, R.L.; Grimm, J.P.; Pickle, J.D.; Sares, S.W. The effect of grain size on detrital modes: A test of the Gazzi-Dickinson point-counting method. J. Sediment. Res. 1984, 54, 103–116. [Google Scholar]
  50. Morton, A.; Hurst, A. Correlation of sandstones using heavy minerals: An example from the Statfjord Formation of the Snorre Field, northern North Sea. Geol. Soc. Lond. Spec. Publ. 1995, 89, 3–22. [Google Scholar] [CrossRef]
  51. Hallsworth, M.C.R. Processes controlling the composition of heavy mineral assemblages in sandstones. Sediment. Geol. 1999, 124, 3–29. [Google Scholar]
  52. Hubert, J.F. A Zircon-Tourmaline-Rutile Maturity Index and the Interdependence of the Composition of Heavy Mineral Assemblages with the Gross Composition and Texture of Sandstones. J. Sediment. Res. 1962, 32, 440–450. [Google Scholar]
  53. Mange, M.A.; Wright, D.T. Heavy Minerals in Use; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  54. Tangari, A.C.; Cirillo, D.; De Luca, R.; Miriello, D.; Pugliese, E.; Le Pera, E. Heavy Minerals Distribution and Provenance in Modern Beach and Fluvial Sands of the Betic Cordillera, Southern Spain. Geosciences 2024, 14, 208. [Google Scholar] [CrossRef]
  55. Hounslow, M.W.; Morton, A.C. Evaluation of sediment provenance using magnetic mineral inclusions in clastic silicates: Comparison with heavy mineral analysis. Sediment. Geol. 2004, 171, 13–36. [Google Scholar] [CrossRef]
  56. Mange, M.A.; Morton, A.C. Geochemistry of heavy minerals. Dev. Sedimentol. 2007, 58, 345–391. [Google Scholar]
  57. Garzanti, E.; Andò, S. Plate tectonics and heavy mineral suites of modern sands. Dev. Sedimentol. 2007, 58, 741–763. [Google Scholar]
  58. Morton, A.C. Geochemical studies of detrital heavy minerals and their application to provenance research. Geol. Soc. Lond. Spec. Publ. 1991, 57, 31–45. [Google Scholar] [CrossRef]
  59. McCarty, R.; Congleton, J. Heavy liquids: Their use and methods in paleontology. In Vertebrate Paleontological Techniques: Volume 1; Cambridge University Press: Cambridge, UK, 2005; Volume 1, p. 187. [Google Scholar]
  60. Visser, M.P.E. Detection of Middle to Late Holocene Icelandic Cryptotephras in the Netherlands: Tephra Versus Biogenic Silica. Master’s Thesis, Utrecht University, Utrecht, The Netherlands, 2012. [Google Scholar]
  61. González-Yajimovich, O.; Perez-Soto, J.L.; Avila-Serrano, G.E.; Meldahl, K. Sediment transport trends in Bahí a Concepció n, Baja California Sur, Mexico, based on textural parameters and heavy mineral concentrations. Boletín De La Soc. Geológica Mex. 2010, 62, 281. [Google Scholar] [CrossRef]
  62. Afzal, Z. The Use of a Magnetic Separator (‘FRANTZ’) and Energy Dispersive Spectroscopy (EDS) for Separation of Heavy Minerals: A Case Study from South Africa and Namibia. Master’s Thesis, University of Stavanger, Stavanger, Norway, 2021. [Google Scholar]
  63. Nasir, S. Effectiveness of Frantz Machine for Separation of Heavy Minerals. Master’s Thesis, University of Stavanger, Stavanger, Norway, 2021. [Google Scholar]
  64. Gao, Y.; Li, F.; Shi, S.; Chen, Y. Determination of paleocurrent directions based on well logging technology aiming at the lower third member of the Shahejie Formation in the Chezhen Depression and its implications. Water 2021, 13, 408. [Google Scholar] [CrossRef]
  65. Banerji, U.S.; Dubey, C.P.; Goswami, V.; Joshi, K.B. Geochemical indicators in provenance estimation. In Geochemical Treasures and Petrogenetic Processes; Springer: Berlin/Heidelberg, Germany, 2022; pp. 95–121. [Google Scholar]
  66. Mondillo, N.; Chelle-Michou, C.; Putzolu, F.; Balassone, G.; Mormone, A.; Santoro, L.; Cretella, S.; Scognamiglio, G.; Tarallo, M.; Tavani, S. The mid-Cretaceous bauxites of SE France: Geochemistry, U-Pb zircon dating and their implications for the paleogeography at the junction between Alpine Tethys and Pyrenean Rift. Gondwana Res. 2025, 137, 145–170. [Google Scholar] [CrossRef]
  67. Bhatia, M.R. Plate tectonics and geochemical composition of sandstones. J. Geol. 1983, 91, 611–626. [Google Scholar] [CrossRef]
  68. Bhatia, M.R. Rare earth element geochemistry of Australian Paleozoic graywackes and mudrocks: Provenance and tectonic control. Sediment. Geol. 1985, 45, 97–113. [Google Scholar] [CrossRef]
  69. Roser, B.P.; Korsch, R.J. Determination of tectonic setting of sandstone-mudstone suites using SiO2 content and K2O/Na2O ratio. J. Geol. 1986, 94, 635–650. [Google Scholar] [CrossRef]
  70. 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]
  71. 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]
  72. Gromet, L.P.; Haskin, L.A.; Korotev, R.L.; Dymek, R.F. The “North American shale composite”: Its compilation, major and trace element characteristics. Geochim. Cosmochim. Acta 1984, 48, 2469–2482. [Google Scholar] [CrossRef]
  73. Borrego, J.; López-González, N.; Carro, B.; Lozano-Soria, O. Origin of the anomalies in light and middle REE in sediments of an estuary affected by phosphogypsum wastes (south-western Spain). Mar. Pollut. Bull. 2004, 49, 1045–1053. [Google Scholar] [CrossRef]
  74. Leybourne, M.I.; Johannesson, K.H. Rare earth elements (REE) and yttrium in stream waters, stream sediments, and Fe–Mn oxyhydroxides: Fractionation, speciation, and controls over REE+ Y patterns in the surface environment. Geochim. Cosmochim. Acta 2008, 72, 5962–5983. [Google Scholar] [CrossRef]
  75. Arslan, M.; Aslan, Z. Mineralogy, petrography and whole-rock geochemistry of the Tertiary granitic intrusions in the Eastern Pontides, Turkey. J. Asian Earth Sci. 2006, 27, 177–193. [Google Scholar] [CrossRef]
  76. Gao, P.; Zheng, Y.-F.; Zhao, Z.-F. Distinction between S-type and peraluminous I-type granites: Zircon versus whole-rock geochemistry. Lithos 2016, 258, 77–91. [Google Scholar] [CrossRef]
  77. Ratcliffe, K.T.; Morton, A.C.; Ritcey, D.H.; Evenchick, C.A. Whole-rock geochemistry and heavy mineral analysis as petroleum exploration tools in the Bowser and Sustut basins, British Columbia, Canada. Bull. Can. Pet. Geol. 2007, 55, 320–336. [Google Scholar] [CrossRef]
  78. Diwu, C.; Sun, Y.; Guo, A.; Wang, H.; Liu, X. Crustal growth in the North China Craton at ~2.5 Ga: Evidence from in situ zircon U–Pb ages, Hf isotopes and whole-rock geochemistry of the Dengfeng complex. Gondwana Res. 2011, 20, 149–170. [Google Scholar] [CrossRef]
  79. Folk, R.L. Petrology of Sedimentary Rocks; Hemphill Publishing Company: Hemphill, TX, USA, 1980. [Google Scholar]
  80. Fohey-Breting, N.K.; Barth, A.P.; Wooden, J.L.; Mazdab, F.K.; Carter, C.A.; Schermer, E.R. Relationship of voluminous ignimbrites to continental arc plutons: Petrology of Jurassic ignimbrites and contemporaneous plutons in southern California. J. Volcanol. Geotherm. Res. 2010, 189, 1–11. [Google Scholar] [CrossRef]
  81. Gualda, G.A.R.; Rivers, M. Quantitative 3D petrography using X-ray tomography: Application to Bishop Tuff pumice clasts. J. Volcanol. Geotherm. Res. 2006, 154, 48–62. [Google Scholar] [CrossRef]
  82. Kruk, N.N.; Babin, G.A.; Kruk, E.A.; Rudnev, S.N.; Kuibida, M.L. Petrology of volcanic and plutonic rocks from the Uimen-Lebed’terrain, Gorny Altai. Petrology 2008, 16, 512–530. [Google Scholar] [CrossRef]
  83. Aubrecht, R.; Mikuš, T.; Holický, I. Heavy mineral analysis of the Turonian to Maastrichtian exotics-bearing deposits in the Western Carpathians: What has changed after the Albian and Cenomanian? Geol. Carpathica 2021, 72, 505–528. [Google Scholar] [CrossRef]
  84. Faupl, P.; Pavlopoulos, A.; Klötzli, U.; Petrakakis, K. On the provenance of mid-Cretaceous turbidites of the Pindos zone (Greece): Implications from heavy mineral distribution, detrital zircon ages and chrome spinel chemistry. Geol. Mag. 2006, 143, 329–342. [Google Scholar] [CrossRef]
  85. Krippner, A.; Meinhold, G.; Morton, A.C.; Schönig, J.; von Eynatten, H. Heavy minerals and garnet geochemistry of stream sediments and bedrocks from the Almklovdalen area, Western Gneiss Region, SW Norway: Implications for provenance analysis. Sediment. Geol. 2016, 336, 96–105. [Google Scholar] [CrossRef]
  86. McLennan, S.M. Relationships between the trace element composition of sedimentary rocks and upper continental crust. Geochem. Geophys. Geosystems 2001, 2. [Google Scholar] [CrossRef]
  87. Taylor, S.R.; McLennan, S.M. The composition and evolution of the continental crust: Rare earth element evidence from sedimentary rocks. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Sci. 1981, 301, 381–399. [Google Scholar]
  88. Wright, J. Conodont apatite: Structure and geochemistry. Skelet. Biominer. Patterns Process. Evol. Trends 1990, 1, 445–459. [Google Scholar]
  89. Wright, J.; Schrader, H.; Holser, W.T. Paleoredox variations in ancient oceans recorded by rare earth elements in fossil apatite. Geochim. Cosmochim. Acta 1987, 51, 631–644. [Google Scholar] [CrossRef]
  90. Wright, J.; Seymour, R.S.; Shaw, H.F. REE and Nd Isotopes in Conodont Apatite: Variations with Geological Age and Depositional Environment; Geological Society of America: Boulder, CO, USA, 1984. [Google Scholar]
  91. Haskin, M.A.; Haskin, L.A. Rare earths in European shales: A redetermination. Science 1966, 154, 507–509. [Google Scholar] [CrossRef]
  92. Anenburg, M. Rare earth mineral diversity controlled by REE pattern shapes. Mineral. Mag. 2020, 84, 629–639. [Google Scholar] [CrossRef]
  93. Segard, M.; Bezos, A.; La, C. Petrology and Geochemistry of Crozet Hotspot Alkali Basalts: Evaluation of the Source Mineralogy; American Geophysical Union: Washington, DC, USA, 2011; p. V51E-2559. [Google Scholar]
  94. Rudnick, R.L. Restites, Eu anomalies and the lower continental crust. Geochim. Cosmochim. Acta 1992, 56, 963–970. [Google Scholar] [CrossRef]
  95. Kato, Y.; Yamaguchi, K.E.; Ohmoto, H. Rare Earth Elements in Precambrian Banded Iron Formations: Secular Changes of Ce and Eu Anomalies and Evolution of Atmospheric Oxygen; Geological Society of America: Boulder, CO, USA, 2006. [Google Scholar]
  96. Brunsmann, A.; Franz, G.; Erzinger, J. REE mobilization during small-scale high-pressure fluid–rock interaction and zoisite/fluid partitioning of La to Eu. Geochim. Cosmochim. Acta 2001, 65, 559–570. [Google Scholar] [CrossRef]
  97. Goodenough, K.M.; Schilling, J.; Jonsson, E.; Kalvig, P.; Charles, N.; Tuduri, J.; Deady, E.A.; Sadeghi, M.; Schiellerup, H.; Müller, A. Europe’s rare earth element resource potential: An overview of REE metallogenetic provinces and their geodynamic setting. Ore Geol. Rev. 2016, 72, 838–856. [Google Scholar] [CrossRef]
  98. Singh, P. Major, trace and REE geochemistry of the Ganga River sediments: Influence of provenance and sedimentary processes. Chem. Geol. 2009, 266, 242–255. [Google Scholar] [CrossRef]
  99. Hoskin, P.W.O.; Ireland, T.R. Rare earth element chemistry of zircon and its use as a provenance indicator. Geology 2000, 28, 627–630. [Google Scholar] [CrossRef]
  100. Irber, W. The lanthanide tetrad effect and its correlation with K/Rb, Eu/Eu∗, Sr/Eu, Y/Ho, and Zr/Hf of evolving peraluminous granite suites. Geochim. Cosmochim. Acta 1999, 63, 489–508. [Google Scholar] [CrossRef]
  101. Mondillo, N.; Boni, M.; Balassone, G.; Spoleto, S.; Stellato, F.; Marino, A.; Santoro, L.; Spratt, J. Rare earth elements (REE)—Minerals in the Silius fluorite vein system (Sardinia, Italy). Ore Geol. Rev. 2016, 74, 211–224. [Google Scholar] [CrossRef]
  102. Zhang, L.; Dengfa, H.E.; Zejun, Y.I.; Di, L.I. Tectonic relationship between the Kelameili range and the Dajing depression: Insights into the Carboniferous tectonic-sedimentary framework. Pet. Explor. Dev. 2020, 47, 30–45. [Google Scholar] [CrossRef]
  103. Zhao, R.; Zhang, J.; Zhou, C.; Zhang, Z.; Chen, S.; Stockli, D.F.; Olariu, C.; Steel, R.; Wang, H. Tectonic evolution of Tianshan-Bogda-Kelameili mountains, clastic wedge basin infill and chronostratigraphic divisions in the source-to-sink systems of Permian-Jurassic, southern Junggar Basin. Mar. Pet. Geol. 2020, 114, 104200. [Google Scholar] [CrossRef]
  104. Hou, Y.; Wu, C.; Chen, L.; Xie, B.; He, L.; Qin, J. New Opinion of Development Potential of Carboniferous Volcanic Gas Reservoir: Taking the Kelameili Gas Filed as an Example. Front. Energy Res. 2022, 10, 861901. [Google Scholar] [CrossRef]
  105. Yang, M.; Hongen, W.; Gaoxue, Y. Geochemical characteristics and tectonic significance of Kalamaili SSZ ophiolite from Eastern Junggar. Acta Petrol. Mineral. 2009, 28, 13. [Google Scholar]
  106. Liu, J.; Xu, X.; Wang, L.; Li, Y. Geochemistry and geochronology of the S hihuiyao ophiolite from the K elameili suture in E ast J unggar: Constraint on the tectonic evolution of the K elameili O cean. Geol. J. 2018, 53, 8–15. [Google Scholar] [CrossRef]
  107. Li, L.; Xiao, W.; Windley, B.F.; Mao, Q.; Gan, J.; Jia, X.; Yang, H.; Sang, M. Defining the Huangcaopo complex and gabbroic magmatism in the northern Harlik Mountains (NW China): Late Cambrian to latest Permian accretionary growth of the East Junggar Arc? Geol. J. 2022, 57, 1022–1045. [Google Scholar] [CrossRef]
  108. Li, L.; Xiao, W.; Zhao, G.; Yang, H.; Han, Y.; Wang, K.; Gan, J. Tearing on the southward subducting Kelameili oceanic lithosphere in the early Devonian: Evidence from the magmatism in the Harlik arc, southern Altaids. Lithos 2023, 454, 107279. [Google Scholar] [CrossRef]
  109. Tianqi, Z.; Chaodong, W.U.; Bo, Y.; Zhongkui, S.H.I.; Jialin, W.; Wen, Z.H.U.; Yanxi, Z.; Jiang, X.; Jinyong, Z.; Jun, W. New insights into multiple provenances evolution of the Jurassic from heavy minerals characteristics in southern Junggar Basin, NW China. Pet. Explor. Dev. 2019, 46, 67–81. [Google Scholar]
  110. Pirajno, F.; Seltmann, R.; Yang, Y. A review of mineral systems and associated tectonic settings of northern Xinjiang, NW China. Geosci. Front. 2011, 2, 157–185. [Google Scholar] [CrossRef]
  111. Tang, H.F.; Su, Y.P.; Liu, C.Q.; Hou, G.S.; Wang, Y.B. ircon u-pb age of the plagiogranite in kalamaili belt, northern xinjiang and its tectonic implications. Geotecton. Metallog. 2007, 31, 8. [Google Scholar]
  112. Wang, B.; Jiang, C.Y.; Li, Y.J.; Wu, H.E.; Xia, Z.D.; Lu, R.H. Geochemistry and tectonic implications of Karamaili ophiolite in East Junggar of Xinjiang. J. Mineral. Petrol. 2009, 9, 74–82. [Google Scholar]
  113. Zhao, H.; Xu, F.; Zhang, J. The kalamaili ophiolites: Age, geological characteristics and tectonic significance. Xinjiang Geol. 2012, 30, 4. [Google Scholar]
  114. Zhang, Y.; Pe-Piper, G.; Piper, D.J.W.; Guo, Z. Early Carboniferous collision of the Kalamaili orogenic belt, North Xinjiang, and its implications: Evidence from molasse deposits. Bulletin 2013, 125, 932–944. [Google Scholar] [CrossRef]
  115. Long, X.; Yuan, C.; Sun, M.; Safonova, I.; Xiao, W.; Wang, Y. Geochemistry and U–Pb detrital zircon dating of Paleozoic graywackes in East Junggar, NW China: Insights into subduction–accretion processes in the southern Central Asian Orogenic Belt. Gondwana Res. 2012, 21, 637–653. [Google Scholar] [CrossRef]
  116. Wu, Z.; Han, X.; Ji, H.; Cai, Y.; Xue, L.; Sun, S. Mesozoic-Cenozoic tectonic events of eastern Junggar Basin, NW China and their significance for uranium mineralization: Insights from seismic profiling and AFT dating analysis. Ore Geol. Rev. 2021, 139, 104488. [Google Scholar] [CrossRef]
  117. Tan, Y.; Xiao, W.; Song, D.; Mao, Q.; Gao, M.; Ao, S. Late Paleozoic intra-oceanic arc and its accretionary complex in East Junggar (NW China): Implications for multiple arc amalgamation in the southern Altaids. J. Geol. Soc. 2024, 181, jgs2023-200. [Google Scholar] [CrossRef]
  118. Li, H.; Li, Y.J.; Xu, X.Y.; Yang, G.X.; Wang, Z.P.; Xu, Q.; Wang, L.J. Petrogenesis and tectonic implications of alkali basalts in Kalamaili area, east Junggar, Xinjiang (NW China): Constraints from petrology, geochronology and geochemistry. Acta Geol. Sin. 2021, 95, 19. [Google Scholar]
  119. Ji, H.; Tao, H.; Wang, Q.; Qiu, Z.; Ma, D.; Qiu, J.; Liao, P. Early to Middle Jurassic tectonic evolution of the Bogda Mountains, Northwest China: Evidence from sedimentology and detrital zircon geochronology. J. Asian Earth Sci. 2018, 153, 57–74. [Google Scholar] [CrossRef]
  120. Zhang, P.; Lee, Y.I. Facies-related diagenesis of Jurassic sandstones, central Junggar Basin, NW China: Implications for reservoir quality evolution. Geol. Soc. Lond. Spec. Publ. 2024, 538, 261–281. [Google Scholar] [CrossRef]
  121. Sha, J.; Olsen, P.E.; Pan, Y.; Xu, D.; Wang, Y.; Zhang, X.; Yao, X.; Vajda, V. Triassic–Jurassic climate in continental high-latitude Asia was dominated by obliquity-paced variations (Junggar Basin, Ürümqi, China). Proc. Natl. Acad. Sci. USA 2015, 112, 3624–3629. [Google Scholar] [CrossRef]
  122. Ding, A.N.; Hui, R.Y.; Zhang, Z.N. Hydrocarbon potential of Jurassic source rocks in the Junggar Basin, NW China. J. Pet. Geol. 2003, 26, 307–324. [Google Scholar] [CrossRef]
  123. Zhang, K.-J.; Ji, C.; Zhou, Y.-Z.; Zhang, Y.-J. Tracing oceanic plateau relics in the basement of mainland China: A synthesis of aeromagnetic and seismic refraction data. Earth-Sci. Rev. 2024, 225, 104849. [Google Scholar] [CrossRef]
  124. Rahman, M.A.; Pownceby, M.I.; Haque, N.; Bruckard, W.J.; Zaman, M.N. Characterisation of titanium-rich heavy mineral concentrates from the Brahmaputra River basin, Bangladesh. Appl. Earth Sci. 2014, 123, 222–233. [Google Scholar] [CrossRef]
  125. Zhang, J.; Cao, J.; Hu, W.; Zhi, D.; Guo, X.; Bian, B. Insights into Carboniferous subduction-related petroleum systems in the Central Asian Orogenic Belt (CAOB) from hydrocarbons in vein calcite cements, West Junggar, northwest China. Mar. Pet. Geol. 2021, 124, 104796. [Google Scholar] [CrossRef]
  126. Li, J.; Chen, X.; Wang, Z.; Chen, W.; Li, C.; Huang, P. Late Palaeozoic mineralization and tectonic evolution of the West Junggar metallogenic belt, Central Asia: Constraints from Re–Os and 40Ar/39Ar geochronology. Int. Geol. Rev. 2017, 59, 1131–1153. [Google Scholar] [CrossRef]
  127. Wu, K.; Pei, Y.; Li, T.; Wang, X.; Liu, Y.; Liu, B.; Ma, C.; Hong, M. Structural characteristics and implication on tectonic evolution of the Daerbute strike-slip fault in West Junggar area, NW China. Front. Earth Sci. 2018, 12, 555–568. [Google Scholar] [CrossRef]
  128. Su, Y.-P.; Tang, H.-F.; Liu, C.-Q.; Hou, G.-S.; Liang, L.-L. The determination and a preliminary study of sujiquan aluminous a-type granites in east junggar, xinjiang. Acta Petrol. Mineralogica. 2006, 25, 10. [Google Scholar]
  129. Yang, H.; JuLi, W.; JianQi, W.; Meng, Y.; Pan, Y. Geochemistry and geochronology of the Miaoergou granite pluton in West Junggar, Xinjiang. Acta Petrol. Sin. 2015, 18, 505–522. [Google Scholar]
  130. Huang, P.; Chen, X.; Wang, Z.; Ye, B.; Li, X.; Yi, Y. Late Paleozoic Granitic Magmatism in West Junggar Metallogenic Belt(Xinjiang), Central Asia, and its Tectonic Implication. Geotecton. Metallog. 2016, 40, 16. [Google Scholar]
  131. Geng, H.Y.; Sun, M.; Yuan, C.; Xian, W.S.; Kong, K. Geochemical, Sr-Nd and zircon U-Pb-Hf isotopic studies of Late Carboniferous magmatism in the West Junggar, Xinjiang: Implications for ridge subduction. Geochim. Cosmochim. Acta 2009, 73, A426. [Google Scholar] [CrossRef]
  132. Chen, J.F.; Han, B.F.; Ji, J.Q. Zircon U-Pb ages and tectonic implications of Paleozoic plutons in northern West Junggar, North Xinjiang, China. Lithos Int. J. Mineral. Petrol. Geochem. 2010, 115, 137–152. [Google Scholar] [CrossRef]
  133. Zhang, Y.; Guo, Z.; Pe-Piper, G.; Piper, D.J.W. Geochemistry and petrogenesis of Early Carboniferous volcanic rocks in East Junggar, North Xinjiang: Implications for post-collisional magmatism and geodynamic process. Gondwana Res. 2015, 28, 1466–1481. [Google Scholar] [CrossRef]
  134. Wang, J.; Wu, C.; Li, Z.; Zhu, W.; Zhou, T.; Wu, J.; Wang, J. Whole-rock geochemistry and zircon Hf isotope of L ate C arboniferous–T riassic sediments in the B ogda region, NW C hina: Clues for provenance and tectonic setting. Geol. J. 2019, 54, 1853–1877. [Google Scholar] [CrossRef]
  135. Han, B.-F.; Guo, Z.-J.; Zhang, Z.-C.; Zhang, L.; Chen, J.-F.; Song, B. Age, geochemistry, and tectonic implications of a late Paleozoic stitching pluton in the North Tian Shan suture zone, western China. Bulletin 2010, 122, 627–640. [Google Scholar] [CrossRef]
  136. Li, D.; He, D.; Santosh, M.; Ma, D.; Tang, J. Tectonic framework of the northern Junggar Basin part I: The eastern Luliang Uplift and its link with the East Junggar terrane. Gondwana Res. 2015, 27, 1089–1109. [Google Scholar] [CrossRef]
  137. Bhatia, M.R.; Taylor, S.R. Trace-element geochemistry and sedimentary provinces: A study from the Tasman Geosyncline, Australia. Chem. Geol. 1981, 33, 115–125. [Google Scholar] [CrossRef]
  138. Armstrong-Altrin, J.S.; Verma, S.P. Critical evaluation of six tectonic setting discrimination diagrams using geochemical data of Neogene sediments from known tectonic settings. Sediment. Geol. 2005, 177, 115–129. [Google Scholar] [CrossRef]
  139. Wang, X.; Algeo, T.J.; Liu, W.; Xu, Z. Effects of weathering and fluvial transport on detrital trace metals. Earth-Sci. Rev. 2023, 241, 104420. [Google Scholar] [CrossRef]
  140. Kasanzu, C.; Maboko, M.A.H.; Manya, S. Geochemistry of fine-grained clastic sedimentary rocks of the Neoproterozoic Ikorongo Group, NE Tanzania: Implications for provenance and source rock weathering. Precambrian Res. 2008, 164, 201–213. [Google Scholar] [CrossRef]
  141. Roy, P.D.; Smykatz-Kloss, W. REE geochemistry of the recent playa sediments from the Thar Desert, India: An implication to playa sediment provenance. Geochemistry 2007, 67, 55–68. [Google Scholar] [CrossRef]
  142. Tanner, P.W.G.; Armstrong, H.A.; Owen, A.W. Rare earth element and La–Th–Sc analysis of cherts from the Highland Border Complex, Scotland: Geochemical determination of the sedimentary environment in greenschist facies rocks. Scott. J. Geol. 2013, 49, 15–31. [Google Scholar] [CrossRef]
  143. Kidder, D.L.; Krishnaswamy, R.; Mapes, R.H. Elemental mobility in phosphatic shales during concretion growth and implications for provenance analysis. Chem. Geol. 2003, 198, 335–353. [Google Scholar] [CrossRef]
  144. Zhang, K.-J. Secular geochemical variations of the Lower Cretaceous siliciclastic rocks from central Tibet (China) indicate a tectonic transition from continental collision to back-arc rifting. Earth Planet. Sci. Lett. 2004, 229, 73–89. [Google Scholar] [CrossRef]
  145. Zhang, K.-J.; Zhang, Y.-X.; Li, B.; Zhu, Y.-T.; Wei, R.-Z. The blueschist-bearing Qiangtang metamorphic belt (northern Tibet, China) as an in situ suture zone: Evidence from geochemical comparison with the Jinsa suture. Geology 2006, 34, 493–496. [Google Scholar] [CrossRef]
  146. Gundogdu, M.N.; Bonnot-Courtois, C.; Clauer, N. Isotopic and chemical signatures of sedimentary smectite and diagenetic clinoptilolite of a lacustrine Neogene basin near Bigadiç, western Turkey. Appl. Geochem. 1989, 4, 635–644. [Google Scholar] [CrossRef]
  147. Daly, J.S. Geochemical and Geochronological Studies in the Stora Le-Marstrand Belt of Orust, S.W. Sweden. Ph.D. Thesis, University of Keele, Keele, UK, 1978. [Google Scholar]
  148. Boger, P.D. A Study of the History of Sedimentation in the Red Sea by Means of Isotopic and Geochemical Methods. Ph.D. Thesis, The Ohio State University, Columbus, OH, USA, 1976. [Google Scholar]
  149. Simonov, V.A.; Mikolaichuk, A.V.; Safonova, I.Y.; Kotlyarov, A.V.; Kovyazin, S.V. Late Paleozoic–Cenozoic intra-plate continental basaltic magmatism of the Tienshan–Junggar region in the SW Central Asian Orogenic Belt. Gondwana Res. 2015, 27, 1646–1666. [Google Scholar] [CrossRef]
  150. Wang, J.; Su, Y.; Zheng, J.; Gao, S.; Dai, H.; Ping, X.; Xiong, Q.; Niyaz, K. Geochronology and petrogenesis of Jurassic intraplate alkali basalts in the Junggar terrane, NW China: Implication for low-volume basaltic volcanism. Lithos 2019, 324, 202–215. [Google Scholar] [CrossRef]
  151. Carroll, A.R.; Yunhai, L.; Graham, S.A.; Xuchang, X.; Hendrix, M.S.; Jinchi, C.; McKnight, C.L. Junggar basin, northwest China: Trapped Late Paleozoic ocean. Tectonophysics 1990, 181, 1–14. [Google Scholar] [CrossRef]
  152. Han, C.; Xiao, W.; Zhao, G.; Su, B.; Sakyi, P.A.; Ao, S.; Wan, B.; Zhang, J.; Zhang, Z. Late Paleozoic metallogenesis and evolution of the East Tianshan orogenic belt (NW China, Central Asia orogenic belt). Geol. Ore Depos. 2014, 56, 493–512. [Google Scholar] [CrossRef]
  153. Guocan, W.; Meng, Z.; Xionghua, Z.; Qun’an, L.; Wei, W.; Jinming, T.; Zeyou, X. Significant Paleozoic tectonic events in the northern part of the East Tianshan Mountains, Xinjiang and their implications for the evolution of CAOB: New evidence from 1: 50000 geological survey of Banfanggou and Xiaoliugou sheets. Geol. China 2019, 46, 954–976. [Google Scholar]
  154. Xiaoqi, W.; Deliang, L.; Zhensheng, L. Post-collisional volcanism of Kalamaili suture zone. Earth Sci. Front. 2009, 16, 220–230. [Google Scholar]
  155. Li, G.-Y.; Zhou, J.-B.; Wilde, S.A.; Li, L. The transition from a passive to an active continental margin in the Jiamusi Block: Constraints from Late Paleozoic sedimentary rocks. J. Geodyn. 2019, 129, 131–148. [Google Scholar] [CrossRef]
  156. Mazza, S.E.; Gazel, E.; Johnson, E.A.; Bizimis, M.; McAleer, R.; Biryol, C.B. Post-rift magmatic evolution of the eastern N orth A merican “passive-aggressive” margin. Geochem. Geophys. Geosyst. 2017, 18, 3–22. [Google Scholar] [CrossRef]
  157. Zhang, X.; Zhao, G.; Eizenhöfer, P.R.; Sun, M.; Han, Y.; Hou, W.; Liu, D.; Wang, B.; Liu, Q.; Xu, B. Varying contents of sources affect tectonic-setting discrimination of sediments: A case study from Permian sandstones in the Eastern Tianshan, northwestern China. J. Geol. 2017, 125, 299–316. [Google Scholar] [CrossRef]
  158. Song, P.; Wang, T.; Tong, Y.; Zhang, J.; Huang, H. Late Carboniferous intrusions along the Kalamaili suture zone, southwestern Central Asian Orogenic Belt (CAOB): Implications for a tectonic switch from subduction to collision. Int. Geol. Rev. 2023, 65, 1601–1621. [Google Scholar] [CrossRef]
  159. Gao, F.; Cheng, Y.; Guo, R.; Liu, X.; Li, Z.; Chen, Y.; Wang, M.; Liu, Z.; Cai, H. Slab break-off of the Kalamaili oceanic slab revealed by the latest Carboniferous mafic–ultramafic rocks in eastern North Tianshan (NW China). J. Asian Earth Sci. 2024, 273, 106274. [Google Scholar] [CrossRef]
  160. Han, Y.; Zhao, G. Final amalgamation of the Tianshan and Junggar orogenic collage in the southwestern Central Asian Orogenic Belt: Constraints on the closure of the Paleo-Asian Ocean. Earth-Sci. Rev. 2018, 186, 129–152. [Google Scholar] [CrossRef]
  161. Wanfeng, C.; Gang, G.; Xiuquan, M.; Jinrong, W.; Wanlong, H.; Binbin, Z.; Lifei, F. Geochemical Characteristics and Tectonic Significance of Early Carboniferous Volcanic Rocks of Kalamaili, Xinjiang. Adv. Earth Sci. 2016, 31, 180–191. [Google Scholar]
  162. Xu, Q.; Zhao, L.; Niu, B.; Zheng, R.; Yang, Y.; Liu, J. Early Paleozoic arc magmatism in the Kalamaili orogenic belt, Northern Xinjiang, NW China: Implications for the tectonic evolution of the East Junggar terrane. J. Asian Earth Sci. 2020, 194, 104072. [Google Scholar] [CrossRef]
  163. Luo, J.; Xiao, W.; Wakabayashi, J.; Han, C.; Zhang, J.e.; Wan, B.; Ao, S.; Zhang, Z.; Tian, Z.; Song, D. The Zhaheba ophiolite complex in Eastern Junggar (NW China): Long lived supra-subduction zone ocean crust formation and its implications for the tectonic evolution in southern Altaids. Gondwana Res. 2017, 43, 17–40. [Google Scholar] [CrossRef]
  164. Xu, X.-W.; Jiang, N.; Li, X.-H.; Qu, X.; Yang, Y.-H.; Mao, Q.; Wu, Q.; Zhang, Y.; Dong, L.-H. Tectonic evolution of the East Junggar terrane: Evidence from the Taheir tectonic window, Xinjiang, China. Gondwana Res. 2013, 24, 578–600. [Google Scholar] [CrossRef]
  165. Ye, X.-T.; Zhang, C.-L.; Zou, H.-B.; Yao, C.-Y.; Dong, Y.-G. Age and geochemistry of the Zhaheba ophiolite complex in eastern Junggar of the Central Asian Orogenic Belt (CAOB): Implications for the accretion process of the Junggar terrane. Geol. Mag. 2017, 154, 419–440. [Google Scholar] [CrossRef]
Figure 1. (a) Tectonic scheme of the Eurasian continent. (b) Tectonic map of Junggar Basin. ① and ② indicate the location of the Hongshan and Miaogou igneous complexes, respectively. (c) Tectonic map of Dongdaohaizi area. Black points indicate the sampling location in this study. Figures are revised after references [15,16,27].
Figure 1. (a) Tectonic scheme of the Eurasian continent. (b) Tectonic map of Junggar Basin. ① and ② indicate the location of the Hongshan and Miaogou igneous complexes, respectively. (c) Tectonic map of Dongdaohaizi area. Black points indicate the sampling location in this study. Figures are revised after references [15,16,27].
Minerals 15 00279 g001
Figure 2. (a) Chronostratigraphy and lithology section of Junggar Basin. (b) Chronostratigraphy and lithology section of Lower Jurassic in Dongdaohaizi Depression. Abbreviations J1b1, J1b2, and J1b3 indicate the lower, middle, and upper members of the Lower Jurassic Badaowan Formation. Figures are revised after references [15,17].
Figure 2. (a) Chronostratigraphy and lithology section of Junggar Basin. (b) Chronostratigraphy and lithology section of Lower Jurassic in Dongdaohaizi Depression. Abbreviations J1b1, J1b2, and J1b3 indicate the lower, middle, and upper members of the Lower Jurassic Badaowan Formation. Figures are revised after references [15,17].
Minerals 15 00279 g002
Figure 3. Q-F-L ternary diagram classification of sandstones from Badaowan Formation. I = Quartzarenite, II = Subarkose, III = Sublitharenite, IV = Arkose, V = Lithic arkose, VI = Litharenite, VII = Feldspathic litharenite. All data are shown in Table A1. Classification follows references [46,79].
Figure 3. Q-F-L ternary diagram classification of sandstones from Badaowan Formation. I = Quartzarenite, II = Subarkose, III = Sublitharenite, IV = Arkose, V = Lithic arkose, VI = Litharenite, VII = Feldspathic litharenite. All data are shown in Table A1. Classification follows references [46,79].
Minerals 15 00279 g003
Figure 4. (ad) Cross-polarized light photomicrograph of sandstone samples from Badaowan Formation. Lv = volcanic lithic fragments, Lm = metamorphic lithic fragments, Ls = sedimentary lithic fragments, Qm = monocrystalline quartz, Qp = polycrystalline quartz. Sampling locations and strata are shown in Figure 2.
Figure 4. (ad) Cross-polarized light photomicrograph of sandstone samples from Badaowan Formation. Lv = volcanic lithic fragments, Lm = metamorphic lithic fragments, Ls = sedimentary lithic fragments, Qm = monocrystalline quartz, Qp = polycrystalline quartz. Sampling locations and strata are shown in Figure 2.
Minerals 15 00279 g004
Figure 5. Relative contents of heavy minerals of sandstones from Badaowan Formation. Solid lines represent averaged data, and shadows represent all data. All data are shown in Table A2.
Figure 5. Relative contents of heavy minerals of sandstones from Badaowan Formation. Solid lines represent averaged data, and shadows represent all data. All data are shown in Table A2.
Minerals 15 00279 g005
Figure 6. NASC-normalized REE patterns of sandstones from Badaowan Formation. Solid icons and lines represent averaged data, and shadows represent all measured data. All data are shown in Table A3.
Figure 6. NASC-normalized REE patterns of sandstones from Badaowan Formation. Solid icons and lines represent averaged data, and shadows represent all measured data. All data are shown in Table A3.
Minerals 15 00279 g006
Figure 7. Identification diagrams of tectonic settings of sandstones from Badaowan Formation in Dongdaohaizi Depression. ACM = active continental margin, PM = passive continental margin, CIA = continental island arc, OIA = oceanic island arc. All data are shown in Table A4. Diagrams are from reference [45].
Figure 7. Identification diagrams of tectonic settings of sandstones from Badaowan Formation in Dongdaohaizi Depression. ACM = active continental margin, PM = passive continental margin, CIA = continental island arc, OIA = oceanic island arc. All data are shown in Table A4. Diagrams are from reference [45].
Minerals 15 00279 g007
Table 1. Averaged geochemical parameters of rare earth elements of sandstones from the Badaowan Formation.
Table 1. Averaged geochemical parameters of rare earth elements of sandstones from the Badaowan Formation.
Sample and StrataΣREELREEHREELREE/HREELaN/YbNδEuδCeLaN/SmNGdN/YbNCeanom
C1-J1b2175.83155.3420.487.581.002.850.980.971.15−0.03
C1-J1b3201.97180.8321.148.561.201.050.960.891.06−0.03
C3-J1b277.7868.179.617.091.242.480.991.041.00−0.02
C3-J1b194.8883.3811.507.250.961.221.001.000.99−0.02
Zh1-J1b3114.17100.6313.537.441.031.160.971.071.00−0.03
δEu = 2 × EuN/(SmN + GdN); δCe = 2 × CeN/(LaN + PrN); Ceanom = lg[3CeN/(2LaN + NdN)]. Data are normalized by NASC [91].
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

Li, F.; Zhang, Z.; Zhao, C.; Han, J.; Liu, J.; Guo, Y.; Tang, X.; Su, C.; Chang, X.; Wu, T. Petrography and Geochemistry of Lower Jurassic Sandstones in the Eastern Junggar Basin: Implications for Provenance and Tectonic Setting. Minerals 2025, 15, 279. https://doi.org/10.3390/min15030279

AMA Style

Li F, Zhang Z, Zhao C, Han J, Liu J, Guo Y, Tang X, Su C, Chang X, Wu T. Petrography and Geochemistry of Lower Jurassic Sandstones in the Eastern Junggar Basin: Implications for Provenance and Tectonic Setting. Minerals. 2025; 15(3):279. https://doi.org/10.3390/min15030279

Chicago/Turabian Style

Li, Furong, Zhi Zhang, Can Zhao, Jinqi Han, Jiaye Liu, Yaoyun Guo, Xinyu Tang, Chang Su, Xu Chang, and Tong Wu. 2025. "Petrography and Geochemistry of Lower Jurassic Sandstones in the Eastern Junggar Basin: Implications for Provenance and Tectonic Setting" Minerals 15, no. 3: 279. https://doi.org/10.3390/min15030279

APA Style

Li, F., Zhang, Z., Zhao, C., Han, J., Liu, J., Guo, Y., Tang, X., Su, C., Chang, X., & Wu, T. (2025). Petrography and Geochemistry of Lower Jurassic Sandstones in the Eastern Junggar Basin: Implications for Provenance and Tectonic Setting. Minerals, 15(3), 279. https://doi.org/10.3390/min15030279

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