Next Article in Journal
Distribution, Sources, and Ecological Risk Assessment of Potentially Toxic Elements in Surface Sediments of Dongzhai Harbor, Hainan Island, China
Previous Article in Journal
Mechanical, Chloride Resistance, and Microstructural Properties of Basalt Fiber-Reinforced Fly Ash–Silica Fume Composite Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genesis of the Upper Jurassic Continental Red Sandstones in the Yongjin Area of the Central Junggar Basin: Evidence from Petrology and Geochemistry

1
Key Laboratory of Deep Petroleum Intelligent Exploration and Development, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing 100029, China
2
University of Chinese Academy of Sciences, Beijing 101408, China
3
SINOPEC Exploration & Production Research Institute, Beijing 102206, China
4
National Center for Geological Experiment and Testing, China Geological Survey, Beijing 100037, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(4), 347; https://doi.org/10.3390/min15040347
Submission received: 11 February 2025 / Revised: 21 March 2025 / Accepted: 25 March 2025 / Published: 27 March 2025
(This article belongs to the Topic Recent Advances in Diagenesis and Reservoir 3D Modeling)

Abstract

:
The sandstone sections in the Upper Jurassic red beds of the Yongjin area in the central Junggar Basin are important oil and gas reservoirs. The debate over whether red beds are of primary depositional or secondary diagenetic origin persists, leading to uncertainties in the interpretation of reservoir sedimentary facies. This study uses core samples and employs thin section microscope observations, scanning electron microscopy, X-ray diffraction, and major and trace element analyses to investigate the formation period and paleoclimate conditions of red beds and explore the origin of red sandstone. The Upper Jurassic red beds are mainly deposited in arid delta plain environments. The framework grains of the red sandstone are composed of quartz (averaging 22.6%), feldspar (averaging 16.3%), and rock fragments (averaging 36.7%). The rock fragments in the sandstone are mainly composed of intermediate basic volcanic rocks and cryptocrystalline acid volcanic rocks, which are rich in mafic silicate minerals such as olivine, pyroxene, ilmenite, and magnetite. In situ hematitization of ilmenite is observed in the rock fragments, suggesting that the in situ alteration of mafic silicate minerals in the parent rock is the main source of iron ions for hematite. Tiny hematite crystals (2.1 μm) are observed in clay mineral micropores via SEM. Abundant mixed-layer illite/smectite clay indicates early smectite transformation, providing a minor source of iron ions for hematite. Hematite in the red sandstone occurs as a grain-coating type, predating quartz overgrowth, feldspar overgrowth, and (ferroan) calcite and (ferroan) dolomite precipitation. Residual hematite coatings between detrital grain point contacts indicate that hematite is a product of syn-sedimentary or very early diagenetic precipitation, ruling out the possibility that red sandstone formation was caused by later atmospheric water leaching during the fold and thrust belt stage. The average chemical index of alteration (CIA) for the red sandstone is 52.2, whereas the CIA for the red mudstone averages 59.5, and the chemical index of weathering (CIW) reached a maximum of 69. These values indicate that the rocks have undergone mild chemical weathering in arid climates. Additionally, the ratios of trace elements indicate that the water bodies were in an oxidizing state during the sedimentary period. The arid climate and oxidative water conditions were ideal for hematite preservation, thus facilitating red bed formation. The red bed sediments in the study area represent a direct response to the Late Jurassic aridification event and can be compared to global climate change. The results have important implications for stratigraphic correlation and interpretation of reservoir sedimentary facies in the study area while also providing a valuable case study for global research on red beds.

1. Introduction

Red beds have been reported in numerous regions worldwide. Notable examples include the Jurassic Navajo red sandstone of the Colorado Plateau in the United States [1], the Cretaceous red sandstones of the Neuquén Basin in Argentina [2], the Triassic Buntsandstein red sandstones of Central Europe and Germany [3], the early Cambrian red sandstones in the Aydıncık region of southern Turkey [4], and the Permian red sandstones in the Ordos Basin of China [5].
The lithology of the red beds varies, potentially comprising conglomerates, sandstones, siltstones, and mudstones. Essentially, the red beds exhibit distinctive coloration due to the presence of iron oxides—primarily hematite crystals—coating the clastic particles [4,6,7,8]. Consequently, red beds are commonly considered indicators of a strongly oxidizing environment [7]. They primarily develop within fluvial, lacustrine, and aeolian deposits in continental basins [9,10,11,12,13,14].
In addition to serving as indicators of oxidizing environments, red sandstones are important hosts for oil, gas, and metal resources globally [11]. They play a crucial role in the formation of metallic and hydrocarbon deposits within sedimentary basins [11,15,16,17,18]. Consequently, extensive research has been conducted on the genesis of red beds [9,19,20,21,22,23]. As early as the early twentieth century, Tomlinson [19] investigated the genesis of red beds and discovered that the coloring agents within these deposits had already formed during sedimentation and were transported and deposited as mechanical sediments. Krynine [24], Folk [25], and Besly [26] further examined the conditions for the formation of red beds, suggesting that these deposits developed in hot/arid or hot/humid sedimentary environments, closely related to paleoclimate conditions. However, Walker [21] and Eren [8] reported that the infiltration of early oxidizing fluids post-sedimentation can lead to the alteration of iron-rich silicate minerals, resulting in the formation of hematite, which can also give rise to red beds. They argued that the presence of red beds more accurately reflects the environmental conditions after deposition rather than a specific climatic type, as these formations can develop under varying climatic conditions. Miot et al. [27] and Eren et al. [4] found that microorganisms can induce hematite precipitation during early diagenesis. Jiang et al. [28] discovered that during the burial of diagenesis, goethite can be dehydrated to form hematite as the temperature rises. Zhang et al. [29] discovered that hematite can be formed during tectonic uplift. These studies indicate that hematite can form at different stages of burial diagenesis. Recently, Shahrokhi et al. [30] and Liu et al. [31] used advanced inductively coupled plasma mass spectrometry to perform in situ elemental content analysis and quantitative diffuse reflectance spectroscopy to distinguish different types of ferrooxides and demonstrated that hematite is jointly controlled by terrigenous input, early diagenetic oxidation, and clay mineral modification. They proposed that hematite has multiple sources and refuted previous views about sediments turning red during late diagenesis, suggesting diagenetic hematite forms only in the early diagenetic stage. Overall, researchers attribute the formation of continental red beds mainly to hot/arid or hot/humid environments, inherited colors from source rock weathering products, and burial diagenesis [32]. Notably, most research has focused on outcrops at the surface, with few investigations into the genesis of red beds in deeper subsurface environments.
Globally, oil and gas exploration is developing to deep and ultra-deep depths [33,34,35]. Deep and ultra-deep oil and gas reservoirs are diverse and include carbonate, clastic, and volcanic reservoirs. Deep marine carbonates and continental deltaic clastic reservoirs are key exploration objects [33,35]. In deep oil and gas reservoir research, the development mechanisms of deep carbonate reservoirs, such as deep burial leaching and early diagenetic permeability reflux, have been extensively studied [34,36,37]. In contrast, the development mechanisms of deep clastic reservoirs have received less attention. Red sandstone, as a key clastic reservoir, is widely developed in the deep zones of oil-bearing basins, for instance, the deep red sandstone reservoirs in the Permian Rotligende formation in the North German Basin of Germany and the Shahejie formation in the Bohai Bay Basin of China, both exceeding 4000 m in burial depth. Previous studies have mostly focused on outcropping red sandstone, with less attention on the deep red sandstone. Therefore, studying the genesis of deep red sandstone reservoirs can help us understand the development mechanisms of deep clastic reservoirs.
The Yongjin area in the hinterland of the Junggar Basin is rich in deep and ultra-deep Jurassic oil and gas resources and has great exploration and development potential. The Upper Jurassic strata are characterized by a sequence of red beds, with red sandstone serving as the primary reservoir for oil and gas. The origin of these red beds has long been a subject of debate. One perspective posits that red beds are formed through oxidation during the simultaneous transport and deposition of sediments under arid climatic conditions [38,39]. Another viewpoint suggests that red beds developed post-deposition, resulting from the uplift of the basin during the Late Jurassic, which brought the strata close to the surface, where they were subjected to weathering and leaching by atmospheric water [40]. Recent research [41] posits that the lower portion of the red beds at the top of the Jurassic is formed by oxidation during deposition, whereas the upper section resulted from superimposed groundwater oxidation on the preexisting localized red beds developed during sedimentation. The controversy surrounding the genesis of Upper Jurassic red beds has resulted in uncertainties regarding the depositional environment of the reservoir in the study area, thereby limiting the characterization and prediction of hydrocarbon reservoirs.
On the basis of detailed petrological and geochemical studies of Upper Jurassic red sandstone, this paper discusses the formation period and coloring mechanism of the red sandstone, investigates the geochemical characteristics of the red beds to infer the paleoclimate and paleoenvironmental conditions during their formation, and further discusses the genesis of the red beds at the top of the Jurassic in the Yongjin area. This study is not only important for deep reservoir prediction and exploration in the Yongjin area but also serves as a valuable case study for global research on red beds.

2. Geological Setting

2.1. Tectonic Background

The Junggar Basin is the second largest inland basin in China; it is located in the northern part of the Xinjiang Uygur Autonomous Region and is a regional negative tectonic unit surrounded by folded mountain ranges (Figure 1a). The basin is bordered on the north by the Hala Alate Mountains and the Qinggelidi Mountains, west by the Zhayier Mountains and the Yilinheibiergen Mountains, and east by the Kelameili Mountains and the Bogeda Mountains. It is characterized by a marginal basement that is higher than the interior, with the northern margin elevated relative to the southern margin. The northern section experiences uplift, whereas the southern section subsides, constrained by concealed faults, resulting in a configuration that resembles an isosceles triangle [42,43,44,45,46]. The basin primarily comprises six first-order structural belts: three uplift belts, namely, the Western Uplift Belt, the Luliang Uplift Belt, and the Eastern Uplift Belt; two depression belts, the Central Depression and the Wulungu Depression; and the Southern Foreland Thrust Belt of the Northern Tianshan Mountains [47]. The Central Depression can be further subdivided into four depressions and six uplifts, as illustrated in Figure 1a.
The Yongjin area is located in the southeastern part of the Central Depression of the Shawan Sag, within the basin’s interior, at the southern flank of the Chemo Uplift and the western segment of the Fukang Sag (Figure 1a) [48]. The Chemo Uplift is an important structure that formed during the Yanshanian period and spans the Dzungarian Basin. Its evolutionary process can be divided into four main stages: the initial formation stage from the late early Permian to the Early Jurassic; the intense uplift and erosion stage during the Middle to Late Jurassic; the concealed burial stage in the Early Cretaceous; and the folding and decline stage from the Late Cretaceous to the Neogene [49,50,51].

2.2. Jurassic Stratigraphy

Drilling data indicate that the Jurassic thickness in the Yongjin area ranges from approximately 430 to 630 m. From top to bottom, it includes the Kalaza Formation (J3k), the Qigu Formation (J3q), the Toutunhe Formation (J2t), the Xishanyao Formation (J2x), the Sangonghe Formation (J1s), and the Badawan Formation (J1b) (Figure 1c). The Lower Jurassic comprises the Badawan Formation and the Sangonghe Formation. The Badawan Formation is primarily composed of gray-black mudstone, sandstone, and siltstone, with thicknesses ranging from approximately 150 to 200 m. The Sangonghe Formation is primarily composed of black coal seams, mudstone, and gray conglomerate, with thicknesses ranging from approximately 100 to 220 m (Figure 1c). During the Middle to Late Jurassic, the formation of the Chemo Uplift resulted in considerable uplift and erosion in the Yongjin area of the basin. This process led to the absence of the Upper Jurassic Kalaza Formation and the Qigu Formation, whereas the Middle Jurassic Xishanyao Formation is preserved only in the central-eastern part of the study area. This formation is primarily composed of gray sandstone and mudstone interbedded with black coal seams and has a thickness of approximately 100–110 m.
The red beds overlie the Xishanyao Formation. No evidence of an obvious unconformable contact between the two units was found in the drill core. The red beds are mainly composed of brownish-red sandstone and mudstone, with intercalations of oil-bearing gray sandstone. The thickness of the residual red beds decreases from south to north, gradually pinching out toward the north, with a range of 62–102 m (Figure 2) [48,52,53,54]. The red beds are in unconformable contact with the overlying gray formations of the chalk group (Figure 1c). Currently, the chronostratigraphic affiliation of this set of red beds remains a matter of debate [38]. Xie et al. [55] conducted a comprehensive analysis using characteristics such as paleobiological fossil assemblages, clay minerals, and logging curves. They concluded that the red beds belong to the Xishanyao Formation and formed as a weathering crust through weathering and leaching processes after sedimentation. Yang Zhi et al. [56] argued that the sedimentation period of the Toutunhe Formation was characterized by arid and hot conditions, which could have led to the formation of such red beds, whereas the sedimentation period of the Xishanyao Formation was characterized by a warm and humid climate, making the deposition of red beds unlikely. Therefore, this paper provisionally refers to this set of strata as the “Upper Jurassic”. This naming aims to reflect the uncertainty in the current academic discussion and provide a foundation for further research.

2.3. Sedimentology

Currently, the upper red beds of the Jurassic system represent the primary stratigraphic target for oil and gas exploration and development in the Yongjin region. Previous studies have extensively examined the sedimentology of red bed sections via drill cores, logging, and seismic data [41,48,53,57,58].
Observations of drill cores indicate that the lithofacies types and vertical distributions of red bed sections are generally consistent. The upper part is characterized by thick layers of brownish-red mudstone, primarily exhibiting massive bedding, with local occurrences of horizontal bedding. Bioturbation structures are also observed within the mudstone, primarily comprising a fossil assemblage of Palaeophycus, Planolites, and Sellaulichnus. These bioturbation structures are formed by worms during habitation, foraging, and crawling on sedimentary substrates [59]. The central region is predominantly composed of brownish-red mudstone interbedded with reddish-brown and light reddish-brown fine sandstone, very fine sandstone, and siltstone. The primary sedimentary structures observed include parallel bedding and ripple bedding. In the lower section, dark gray and gray medium sandstone, along with very fine sandstone, are interbedded with brownish-red mudstone. The gray sandstone features parallel bedding and cross-bedding. Notably, the gray sandstone is invariably saturated with oil. Furthermore, no plant debris has been observed in red mudstone, red sandstone, or oil-bearing gray sandstone (Figure 3).
The interpretation of these sedimentary environments remains highly controversial. Some scholars suggest that red beds developed primarily in littoral and shallow lacustrine sedimentary environments under humid to semi-arid conditions [60,61]. The alternating appearance of reddish-brown and gray colors may reflect fluctuations in the paleoclimate between humid and arid conditions. However, paleontological evidence from the upper red bed section suggests that, in contrast to the Xishanyao Formation, this section is nearly devoid of spore-pollen fossils and exhibits scarce paleontological traces. This idea indicates that the paleoclimate during the deposition of the red beds was likely characterized by persistent aridity [58]. The widespread development of reddish-brown mudstone further indicates the sustained development of arid environments. Therefore, other scholars believe that the upper red bed section primarily formed under arid conditions as shallow braided river delta deposits, gradually transitioning upward into shallow meandering river delta deposits [52,53,58].

3. Materials and Methods

To analyze the reasons for the formation of the upper red beds in the Deep Jurassic, we made a detailed observation of the cores from nine drilled wells in the study area. It was found that the cores from these nine coring wells showed consistent developmental characteristics from top to bottom. The upper section of the core consists predominantly of thick-bedded red mudstone with evident blocky bedding. The central region is predominantly composed of grey medium sandstone, fine sandstone interbedded with reddish-brown fine sandstone, very fine sandstone, and siltstone. Among these, the gray sandstones primarily exhibit parallel and cross-bedding, while the red sandstones mainly display parallel and ripple bedding. The lower section is characterized by red mudstone interbedded with red muddy siltstone. Overall, the cores exhibit multiple fining-upward sequences. Based on the above observations, we collected samples from nine wells (Y2, Y6, YJ1-1, YJ1-3, YJ12, YJ14, YJ3-X17, YJ15, and YJ303). This collection included 7 samples of red mudstone and 23 samples of red sandstone. The sandstone samples cover fine sand, very fine sand, and siltstone with calcareous and muddy cementation. These samples were used for thin section petrology, SEM, XRD, and major–trace element characterization analyses.

3.1. Petrography

To obtain the petrographic characteristics of the red beds in the study area, standard thin sections were prepared from 22 core plugs and rock samples collected from eight wells, including Y2, YJ1-1, and YJ12, in the Yongjin area. The thin sections were stained with a mixture of alizarin red and potassium ferrocyanide to differentiate between different types of authigenic carbonate minerals. Under an Axio Scope A1 ZEISS petrographic microscope (Carl Zeiss Microscopy Deutschland GmbH, Oberkochen, Germany), the content of clastic particle components was quantified via the point counting method. At the same time, the petrographic structural characteristics of the red sandstone were observed, including particle sorting, rounding, contact relationships, types of cement, and their relative spatial coexistence relationships.

3.2. SEM

Eight representative sandstone samples were selected. Portions of these samples were prepared as probe slices, while the remainder were cut into units approximately 1 cm in diameter and polished. After the samples were covered with a conductive coating (carbon coating), they were adhered to the specimen stage of the electron microscope via a conductive adhesive. Under an acceleration voltage of 15 kV and a spot size of 5.5, the working distance was adjusted to approximately 7 mm. The surface mineral morphology of the particles was observed via a Nova NanoSEM 450 scanning electron microscope (Thermo Fisher Scientific, Hillsborough, NJ, USA), resulting in backscattered and morphological images of the samples. The energy spectrum elemental maps of the sample minerals were obtained using an X-ray energy spectrometer equipped with an X-MAXN80 instrument (Oxford Instruments, Oxford, UK). Additionally, two unpolished fragments of red sandstone were selected. After the surfaces were coated with a conductive film (approximately 17 nm thick), they were adhered to the sample stage of a scanning electron microscope using a conductive adhesive. The mineral morphological characteristics of the particle surfaces were observed via a ZEISS scanning electron microscope (Carl Zeiss Microscopy Deutschland GmbH, Oberkochen, Germany) at an accelerating voltage of 15 kV.

3.3. XRD Diffraction

To identify the types of clay minerals in the samples, ten representative samples of red sandstone and mudstone were selected for XRD diffraction analysis of the clay minerals. The selected samples were ground to 200 mesh via a disk vibration mill. For each sample, 10 g were weighed and mixed with 200 mL of deionized water. After thorough mixing, 10% H2O2 was added to remove organic matter, followed by the addition of 2% hydrochloric acid (HCl) to remove carbonates. Once the impurities were removed, deionized water was added for three rounds of centrifugation, ensuring that the pH of the centrifuged samples was between 5 and 7. The samples were transferred into a beaker, and deionized water was added for stirring. After standing, the middle portion of the clay particles (particle size < 2 μm) was extracted. The extracted clay underwent titration and was dried to prepare standard, ethanol, and high-temperature slides for XRD analysis, which revealed the clay mineral composition of the red bed mudstone and sandstone.

3.4. XRF Spectrum

To obtain the geochemical characteristics of the red beds, we selected 12 representative red sandstone samples and 2 red mudstone samples. The concentrations of major and trace elements in these samples were measured via wavelength-dispersive X-ray fluorescence spectroscopy (WD-XRF). Additionally, 4 red mudstone samples were selected for separate trace element analysis. The selected samples were ground to a particle size of 200 mesh and then dried in an oven at 105 °C for 12 h. After drying, the samples were cooled in a desiccator for 2 h. Subsequently, 0.6 g of the standard sample was mixed uniformly with 6.0 g of lithium tetraborate/borate (2:1) in a Pt–Au crucible and mold. Four drops (0.18 mL) of a 0.12 g/mL ammonium bromide solution were added, and the mixture was melted via an electric fusion device at 1050 °C for 19 min, followed by cooling for 285 s to form glass discs. The loss on ignition (LOI) value was measured by igniting 0.6 g of the dried sample in a muffle furnace at 1000 °C for 1.5 h. The prepared glass discs were then analyzed via a PANalytical AXIOS Minerals (Malvern Panalytical, Almelo, The Netherlands) instrument equipped with a rhodium anode X-ray tube and an excitation power of 4 kW in an XRF spectrometer. The detection limits (LODs) and quantification limits (LOQs) for each element were automatically determined via SuperQ 4.0 software. On the basis of the measurement results and the LOD/LOQ data, a quantitative analysis of the elemental content in the samples was conducted, yielding geochemical data for the red layers [62].
Observations of unpolished rock fragments and X-ray diffraction analysis of clay minerals were performed at the Chinese Academy of Geological Sciences, whereas the remaining experiments were conducted at the Institute of Geology and Geophysics, Chinese Academy of Sciences.

4. Results

4.1. Petrographic Characteristics of Red Sandstone

Thin section statistical data indicate that the Upper Jurassic red sandstone in the study area predominantly comprises very fine-grained sandstone, followed by medium-grained fine sandstone, very fine sandstone, and fine-grained very fine sandstone. The sandstones exhibit poor to moderate sorting, with rounding primarily characterized as subangular to subrounded, reflecting a low degree of textural maturity. The particle contacts are predominantly point lines, with a minor proportion exhibiting line contacts.
According to the sandstone classification scheme proposed by Folk [63], the red sandstone in the study area is classified primarily as feldspathic litharenite, with a minor presence of litharenite (Figure 4). The content of detrital quartz is relatively stable and consists primarily of monocrystalline quartz, ranging from 12.9% to 23.3%, with an average of 17.9%. Polycrystalline quartz and chert content range from 1.8% to 9.3%, averaging 4.6%. The feldspar is mainly potassium feldspar, with a relatively low content ranging from 6.2% to 12.3% and an average of 9.2%. Plagioclase content ranges from 2.8% to 12.9%, with an average of 7.0%. The rock fragment content is the highest, ranging from 27.5% to 48.3%, with an average of 36.9% (Table 1). The volcanic rock fragments are the most common, ranging from 11.5% to 31.6% and averaging 21.3%; metamorphic rock fragments average 15.2%, and sedimentary rock fragments are the least abundant, averaging 0.4% (Table 1). According to the continental sandstone framework mode triangular diagram of Dickinson [64] (Figure 5), the sandstone data in the study area are mainly plotted in the transitional arc area, with a few data plotted in the dissected arc area. Sandstone framework grain development indicates the source is from a magmatic arc with moderate denudation. The Qm-F-Lt diagram shows that monocrystalline quartz (Qm) content is low and polycrystalline rock fragments (Lt) content is high, indicating plutonic rocks are less eroded, with volcanic rock fragments being the main component.
Thin section study shows that volcanic rock fragments are predominantly extrusive rock fragments, with a minor amount of intrusive rock fragments. Extrusive rock fragments are mainly medium andesite fragments (Figure 6a,c) and acid cryptocrystalline extrusive rock fragments (Figure 6h). Andesite fragments show long tabular plagioclase phenocrysts and their matrix appears black under orthogonal polarization. Acidic extrusive rock fragments mainly show porphyritic texture, with a cryptocrystalline felsic matrix and black spots on the surface. Mafic basalt fragments are common, displaying intergranular textures with triangle interstices filled with dark minerals (Figure 6d,g). Intrusive rock fragments are mainly altered basic diabase fragments (Figure 6f,g) and are less common. Metamorphic rock fragments are mainly metamorphic sandstone, schist, phyllite, and mud slate fragments with medium and low metamorphic degrees. Sedimentary rock fragments are mainly argillaceous rock fragments. The sandstone components have low compositional maturity, indicating a proximal, rapid-accumulation setting.

4.2. Diagenetic Authigenic Mineral Characteristics of Red Sandstone

The authigenic minerals in the Upper Jurassic red sandstone include fine particles of hematite, quartz overgrowth, feldspar overgrowth, authigenic carbonate minerals, analcime, and various authigenic clay minerals.

4.2.1. Hematite

Under microscopic observation of red sandstone thin sections, the surfaces of the detrital grains are generally covered by opaque reddish-brown material, whereas the interstitial spaces are filled with a substantial amount of opaque reddish-brown clay matrix (Figure 7a). Under reflected light, these opaque reddish-brown coatings and the red clay matrix filling the pores exhibit vibrant red and orange-red luster (Figure 7b), which is attributed to the infiltration of hematite into the clay matrix. Additionally, some grain surfaces develop black opaque spots (Figure 7a), which also display a red luster under reflected light (Figure 7b), indicating that these grains have undergone a process of alteration associated with hematitization.

4.2.2. Authigenic Carbonate

Microscopic observations reveal that three main types of carbonate cements have developed in the red sandstone of the study area. The first type is a substrate-type cement of sparry calcite, which appears pink after being stained with a mixture of alizarin red S and potassium ferricyanide. This sparry calcite occupies nearly all the intergranular pore spaces, with only a reddish-brown hematite film developed between the sparry calcite cement and the grains (Figure 7d). This type of calcite is usually considered to have formed during early diagenetic processes that occurred simultaneously with sedimentation. The second category comprises iron calcite, which is embedded in the pores in a granular mosaic texture and appears purple under plane-polarized light. This ferroan calcite fills the residual intergranular pores formed during the growth of authigenic feldspar and quartz overgrowth (Figure 7e,g), often engaging in replacement interactions with the surrounding grains (Figure 7g). The third category comprises ferroan dolomite cements that exhibit rhombohedral grain shapes, typically filling larger intergranular voids, and a reddish-brown hematite film grows at the contact points with the grains (Figure 7h).

4.2.3. Secondary Quartz Overgrowth

The phenomenon of quartz overgrowth is commonly observed in the red sandstone of the study area, primarily exhibiting a rimmed growth pattern, which is predominantly produced in a single generation.
Microscopic observations revealed that the edges of the quartz overgrowth particles were uniformly surrounded or partially encircled by growth. Additionally, a thin film of hematite is observed to grow between the quartz overgrowth and the grains, followed by the external growth of the calcite cement (Figure 7e). This petrological relationship indicates that quartz overgrowth formed after the deposition of the hematite film and prior to calcite cementation, suggesting an earlier formation time.

4.2.4. Authigenic Feldspar

In the red sandstone of the study area, feldspathic cement is also widely developed, primarily in two forms: relatively overgrowth (Figure 7e) and authigenic feldspar microcrystals (Figure 7f,g). The authigenic feldspar microcrystals are more prevalent than the feldspar overgrowth, predominantly exhibiting a plate-columnar morphology on the grain surfaces. Both the feldspar overgrowth and the authigenic feldspar microcrystals evidently grow outside the hematite film.

4.2.5. Analcime

Under microscopic observation of red sandstone, analcime cement is occasionally observed, typically coexisting with iron calcite in larger intergranular pores. Some analcime surfaces exhibit the development of dissolution pores (see Figure 7i), which formed later than the hematite.

4.2.6. Clay Minerals

On the basis of detailed observation of polished thin sections of red sandstone via scanning electron microscopy, we found that the pores and particle edges predominantly developed amorphous mixed-layer illite/smectite clays and fibrous illite clays. The interstitial spaces between the particles were filled with a considerable amount of clay matrix rich in quartz and feldspar fragments (see Figure 8a,b). Some bright, fine crystal fragments are dispersed within the clay matrix (see Figure 8a,c), and energy dispersive spectroscopy confirms that these fragments are primarily composed of hematite crystals (see Figure 8f). Within the intergranular pores, considerable amounts of fibrous illite, flaky chlorite cement, and amorphous mixed-layer illite/smectite clay minerals are present, interspersed with fragments of quartz and feldspar (see Figure 8c). Among the micropores of the clay matrix, substantial growth of isolated microcrystalline hematite particles in a flaky, clastic form can be observed (see Figure 8d). This distribution imparts characteristics to the clay matrix within the pores that are analogous to those of the clay matrix on the surface of the particles. X-ray energy dispersive spectroscopy revealed the partial substitution of iron (Fe) by titanium (Ti) in hematite (see Figure 8f). Additionally, the development of fragmented titanium oxide particles was observed within the clay matrix (see Figure 8c).
Observations of unpolished red sandstone samples via scanning electron microscopy yielded results consistent with those obtained from polished thin sections. On the surface of the detrital grains, we identified flaky chlorite clay and abundant fibrous illite clay (see Figure 9a). The particle surfaces exhibited extensive development of a clay matrix, which contained numerous fine granular quartz grains, feldspar fragments, and amorphous mixed-layer illite/smectite clays. Within these clay matrices, small hematite crystal fragments were also widely distributed (see Figure 9b).

4.3. X-Ray Diffraction Patterns of Clay Minerals in Red Sandstone and Mudstone

X-ray diffraction analysis was conducted on red bed samples from the study area, and the results are presented in Table 2. The findings indicate that the Upper Jurassic red bed mudstone in the Yongjin area predominantly contains four types of clay minerals: mixed-layer illite/smectite, illite, chlorite, and kaolinite. Among the clay minerals, the illite/smectite mixed layer has the highest content, with an average of 42.0%. This component is followed by illite, which has an average relative content of 40.2%. The chlorite content ranges from 11.0% to 16.0%, with an average content of 12.2%, indicating a relatively low abundance. Kaolinite is not well developed, with a distribution range of 5.0% to 7.0%. The clay mineral development in the red sandstone is analogous to that in the mudstone, which is characterized predominantly by the presence of mixed-layer illite/smectite clay minerals, with an average relative content of 37.2%. The contents of illite and chlorite are lower than those of illite/smectite, averaging 31.4% and 25.4%, respectively. Kaolinite is relatively rare, with an average content of 6.0%. The distribution of the clay mineral content obtained via X-ray diffraction (XRD) analysis is consistent with that observed via scanning electron microscopy (SEM).

4.4. Geochemical Characteristics

The major and trace element characteristics of red mudstone and sandstone are presented in Table 3 and Table 4. The major element analysis indicates that red sandstone has a high SiO2 content, ranging from 66.45 wt.% to 78.14 wt.%, with an average of 73.0 wt.%. The next most abundant element after SiO2 is Al2O3, which ranges from 10.15 wt.% to 15.5 wt.%, with an average concentration of 11.0 wt.%. The contents of Na2O (2.93 wt.% to 3.72 wt.%), K2O (1.97 wt.% to 3.37 wt.%), and MgO (0.65 wt.% to 2.51 wt.%) decrease in decreasing order. The variations in CaO and loss on ignition (LOI) are considerable, ranging from 0.51 to 4.21 wt.% and from 1.55 to 6.37 wt.%, respectively. These variations are attributed primarily to the differing carbonate contents within the red sandstone samples. The loss of carbonates during calcination contributes to the broad range of LOI values. Additionally, the variability in the Fe2O3 content (2.04 to 4.82 wt.%) is influenced by the heterogeneity of the red sandstone samples, with some samples exhibiting higher iron-rich clay contents in their pores, whereas others only display a thin coating of hematite on the particle surfaces, resulting in substantial intergranular porosity. The average Fe2O3 content is 2.80 wt.%. Furthermore, the samples contain trace amounts of TiO2 (0.29 to 0.71 wt.%), MnO (0.03 to 0.29 wt.%), and P2O5 (0.09 to 0.13 wt.%).
Compared with the red sandstone samples, the red mudstone samples present lower SiO2 (average content of 57.4 wt.%) and CaO (average content of 1.4 wt.%) levels, indicating a lower carbonate content in the red mudstone. Conversely, the Fe2O3 and TiO2 contents are considerably higher than those in the sandstone samples, with average values of 7.3 wt.% and 0.7 wt.%, respectively. This comparison suggests that the clay contains a greater proportion of hematite fragments. The relatively high contents of Al2O3 (average 17.5 wt.%), K2O (average 3.6 wt.%), and MgO (average 2.5 wt.%) can be attributed to the greater development of clay minerals within the mudstone. In contrast, the levels of Na2O (average 3.4 wt.%), MnO (average 0.1 wt.%), and P2O5 (average 0.2 wt.%) do not substantially vary.
The trace element contents of the tested samples exhibit notable correlations and regularity. The red sandstone is characterized by a high concentration of Ba (average of 481.3 ppm), moderate to high levels of Sr (123.1 ppm), Cr (59.7 ppm), and V (32.3 ppm), and low concentrations of Co, Ni, and Zn. In contrast, the mudstone samples display higher concentrations of Ba (average of 695.3 ppm), V (average of 85.6 ppm), and Zn (average of 65.3 ppm) than the sandstone samples do. This phenomenon is attributed primarily to the effects of diagenesis, with insignificant differences observed in the concentrations of other elements.
To better understand the concentrations of major and trace elements in the red bed samples, the average elemental data were compared with the upper continental crust (UCC) [65,66] data. The results (Figure 10a,b) indicate that, with a few exceptions, the samples exhibit a similar pattern to UCC normalization.
For the sandstone samples, the concentrations of major elements such as Ti, Al, Fe, Na, K, and P tend to be depleted compared with those in the UCC, while Mn and Ca exhibit significant variability, likely due to carbonate cement in some red sandstone samples. Compared to UCC, trace elements like Co, Sr, V, and Zn show depletion, while Cr and Ni exhibit significant variability. This may be due to the contact with gray sandstone in some samples. For the mudstone samples, Fe, Al, and Mg are relatively enriched, and Ca and P are depleted. Ba, Cr, Ni, and Zn show enrichment trends compared to the UCC, while Sr and V are depleted.

5. Discussion

5.1. Formation Time of Red Beds

The timing of the formation of red beds is crucial for understanding the genesis of these deposits. Various explanations have been proposed for when sedimentary rocks and soils become red. Krynine [24] and Folk [25], among others, suggested that the coloration of the rocks occurred before deposition, indicating that the sediments were already red at the time of deposition. In 1976, Walker [21] proposed a new perspective, suggesting that typical red beds are products of diagenesis. According to his view, the sediments were not red at the time of deposition. Over time, the iron-bearing particles within the sediments came into contact with the groundwater in an unbalanced state. This interaction led to the reddening of the rocks. This process occurs during the very early stages of diagenesis. Researchers in the academic community have subsequently proposed that red beds form during various stages of diagenesis. For example, Eren et al. [4] were the first to reveal the phenomenon of microbial-induced precipitation of hematite during the early diagenetic stage. Scholars such as Zhang et al. [29] proposed that red beds undergo reddening during the weathering process in the epigenetic diagenetic stage. These authors suggested that the uplift of the basin caused the strata to enter the phreatic oxidation zone, thereby triggering the weathering process of hematitization. In addition, Lian [28] conducted laboratory heating simulation experiments and proposed that red beds may experience thermal metamorphism and reddening due to high-temperature dehydration during the middle to late diagenetic stages. However, the recent study by Shahrokhi et al. [30] challenges the idea of red bed formation during the late diagenesis stage. They argue that post-sedimentary rock reddening mainly occurs during the early diagenesis stage. These studies enhance our understanding of red bed formation timing and provide a solid theoretical basis for further research in related fields.
The formation time of Upper Jurassic red beds in the study area has been heavily researched. It remains unclear whether the formation of red beds was primarily due to oxidation during the sedimentary process, the result of atmospheric water leaching and erosion following the uplift of the strata at the end of the Jurassic, or a combination of both factors. Currently, no consensus exists in the academic community. Jiao et al. [40] observed under a microscope that no hematite was found at the contact points between particles in the red beds; instead, hematite alteration was more prevalent in the intergranular pores. These authors suggested that hematitization occurred later than the initial compaction and that the red beds were oxidation products from a noncontemporaneous sedimentary period. However, our study revealed hematite films at clastic particle points and point-line contacts in red beds (Figure 7c). This petrological relationship clearly indicates that the hematite films formed prior to considerable compaction [20,67]. Microscopic observations of diagenetic authigenic minerals indicate that some rock fragments undergo hematitization alteration on their surfaces. Furthermore, before the extensive development of early quartz overgrowth, the reddish-brown hematite uniformly coated the surfaces of the clastic particles. This phenomenon is evidenced by the occurrence of quartz overgrowths encapsulating hematite films. (Figure 7e). Moreover, within the interparticle pores, authigenic feldspar microcrystalline, analcime, ferroan–calcite, and ferroan dolomite are observed to grow extensively on top of the hematite films (Figure 7f–i). Even in dense calcite-cemented sandstones, hematite films are visible between particles and pore calcite (Figure 7d). This calcite is typically considered to have formed during the early stages of diagenesis, concurrently with the deposition of clastic particles.
These observations indicate that hematite precipitated very early, with growth on the surfaces of clastic particles preceding cement formation. The subsequent diagenetic alteration does not cause the red coloration of the rock. Therefore, we posit that the sediments acquired their red characteristics during deposition or shortly after, in the early diagenetic stage. Research by Jiang [32] on the Middle Jurassic continental red beds in the Sichuan Basin supports this hypothesis, showing that an oxidizing environment during or soon after deposition can cause hematite precipitation. Hu’s work [68] on Cretaceous red beds also confirms that hematite can form during syn-sedimentary to early diagenetic stages.
Iron is commonly present in the layered silicate structure as divalent or trivalent forms and may also occur as exchangeable cations in the interlayers of layered clay minerals [31,32]. Sediment provenance analysis shows that the red sandstones in the study area contain abundant volcanic rock fragments, mainly andesite, basalt, and acidic cryptocrystalline extrusive rock fragments, which underwent extensive hematitization (Figure 6b,e,i). These volcanic rock fragments, especially andesite and basalt, are difficult to identify as to their original minerals due to extensive hematitization. It is inferred that they were altered from olivine, pyroxene, ilmenite, and magnetite. The hematitization of ilmenite (Figure 6i) further confirms that in situ alteration of iron-rich minerals is the main source of iron ions for hematite, a conclusion supported by many studies [6,20,21]. Scanning electron microscopy observations revealed that the surfaces and interstitial spaces of the particles in the red sandstone were primarily occupied by amorphous clay minerals. These clay minerals infiltrate particle surfaces and interparticle pores during sedimentation, and hematite crystals form as fine dispersions within clay micropores (Figure 8). As hematite has a fine grain size, we cannot observe its surface morphology, which hinders the analysis of its formation and relationship with clay minerals. Liu et al. [69] used high-resolution transmission electron microscopy (HRTEM) to observe the relationship between hematite and clay minerals, a method worth our reference. Their research reveals that the interplanar spacing of hematite matches the lattice stripe spacing of kaolinite. This is mainly due to the release of iron ions during the conversion of montmorillonite to kaolinite, which are adsorbed on the kaolinite surface and precipitate to form hematite. In this regard, we infer that during the sedimentation or early diagenesis of the red beds in the study area, the transformation of smectite to illite would release Fe and Mg, thus providing an iron source for hematite in the clay matrix. The widespread occurrence of mixed-layer illite/smectite clay minerals provides strong evidence for this inference, indicating that smectite was once extensively converted to illite. The iron in clay mineral lattices is limited, making them a minor iron source in red beds [6,20,21,25]. Overall, the iron in hematite mainly comes from the transformation of volcanic rock fragments during the syn-sedimentary and early diagenetic stages, especially those unstable iron-bearing silicates (ilmenite, magnetite, olivine, pyroxene). The conversion of montmorillonite to illite is a secondary source of iron. These minerals release iron ions in situ under suitable geological conditions, leading to hematite formation.

5.2. Paleoenvironment and Paleoclimatology of Red Beds

The traditional view holds that sedimentary rock layers with a distinctive red color are formed under arid to semi-arid climatic conditions, with dry and oxidizing conditions that help preserve hematite. Early studies on modern red deserts also support the view that red beds form under hot and arid climatic conditions [20,67]. However, with more research and the discovery of red beds in different regions, this traditional view has been challenged. Some studies suggest that red beds do not necessarily indicate specific paleoclimate conditions [70] and that red beds can also form in nonarid environments. For example, under warm and humid climatic conditions, red sediments can form through good drainage conditions and specific soil chemical processes [21]. As shown in Table 5, the formation of red beds involves various complex environmental factors and can occur in different settings. It should not be reduced to one model but should be interpreted in regional studies [70].
The content and composition of clay minerals are often used to reconstruct paleoenvironments [71,72,73,74,75]. Li et al. [41] and He [38] observed the clay mineral assemblages and their content variations in the red bed section of the study area. They reported that the clay minerals exhibited a distribution characterized by high illite and relatively high mixed-layer illite/smectite contents, low amounts of chlorite and montmorillonite, and a lack of kaolinite. They concluded that this assemblage indicates that the sedimentation period of the red beds in the Yongjin area was a hot and arid oxidizing environment, which experienced weak alkaline and weak chemical weathering processes. The X-ray diffraction and scanning electron microscopy results of the clay minerals from the red bed samples indicate that the primary clay minerals are amorphous mixed-layer illite/smectite and fibrous illite, with average contents of 42% and 40%, respectively. Additionally, a small amount of flaky chlorite clay has developed, while montmorillonite and kaolinite are absent. These findings are consistent with previous research results. However, beyond 3000 m depth and 100 °C, considerable diagenetic changes in clay minerals occur [73,75]. In the Yongjin area, the deep burial of red beds leads to considerable diagenetic effects on clay mineral contents and types, which are often ignored when interpreting the paleoenvironments of clay minerals. Although the findings of our study regarding clay minerals are generally consistent with those of previous research, we contend that clay minerals alone cannot be used to definitively reconstruct paleoenvironments and that their environmental indicators may be inaccurate.
Table 5. Geochemical indicators for red bed formation in different regions globally.
Table 5. Geochemical indicators for red bed formation in different regions globally.
RegionSystemGeochemical IndexIndex Range CharacteristicsRed Bed GenesisReferences
Songliao Basin, ChinaCretaceousCIA56.2~67.8The early weathering is weak, and it is formed in an oxygen-rich water environment in an arid climateSong et al., 2015 [76]
V/cr1.0~1.7
Ni/Co1.9~2.5
U/Th0.1~0.2
Sr/Cu1.0~10.0
Sichuan Basin, ChinaJurassicCIA66.2–75.3Weathering of source rocksJiang et al., 2023 [32]
CIW77.5–86.6
Lufengchuan Street Basin, ChinaJurassicCIA62.5–78.0The weathering conditions are moderate, and the red beds are mainly formed in the post-sedimentary oxygen-rich sedimentary waterWang et al., 2024 [77]
U/Th0.16–1.03
V/Cr0.50–1.85
Cala Viola, ItalyPermianCIA86.8–98.5Non-weathering formation, good drainage, and oxidative sedimentary water conditions promote the formation of red bedsSheldon et al., 2005 [70]
Al2O34.99%–20.65%
CaO0.03%–0.18%
Sr36–272 ppm
Mersin area, TurkeyMioceneMn/Fe0.003–0.023It is formed in an alternating wet and dry climate environment, in an oxidizing water sedimentary environmentEren et al., 2015 [67]
Ni/Co2.3–9.2
Sr/Ba0.3–2.2
Transvaal and Olifantshoek Basin, South AfricaPaleoproterozoicCIA72–78Formed by early weathering and post-deposition oxidationLand et al., 2017 [78]
Al2O312.9–20.3
CaO0.01–1.5
MgO0.2–2.3
The elemental composition of clastic sedimentary rocks, particularly mudstones and siltstone–mudstone, is often used to reflect paleoclimatic conditions during deposition [64,79]. The use of sandstone or its combination with mudstone to determine the paleoclimatic environment during the sedimentary period has been successfully applied in many studies [80,81,82]. Commonly used elemental ratio indicators like Fe/Mn, CIA, and CIW are of great significance for analyzing sedimentary paleoclimate environments. Thus, the precision of elemental measurements is crucial for the accuracy of inference results. However, conventional X-ray fluorescence spectrometry cannot distinguish iron oxides of different valence states and is prone to errors during measurement. In contrast, diffuse reflectance spectrometry can identify various iron oxide minerals and their contents, and micro area in situ geochemical testing offers more precise elemental content data [30,69]. Notably, the elemental compositions of red bed sandstones and mudstones are highly susceptible to diagenetic alteration during burial, resulting in significant discrepancies between their current elemental compositions and those during deposition. Consequently, indices calculated based on major and trace element measurements may not accurately reflect the paleoclimate during deposition. In contrast, isotope analysis techniques, such as δ13C, δ18O, δ7Li, and δ⁹⁸/⁹⁵Mo isotope tracking, are less affected by post-depositional diagenetic evolution. They can more precisely indicate the depositional conditions and better identify the paleoclimate during red bed sedimentation [77,83]. Moreover, utilizing spore–pollen assemblages can help reduce the uncertainty of elemental ratios when reconstructing paleoclimate.
Owing to experimental limitations, this study primarily employs X-ray fluorescence spectrometry to measure major and trace elements in red bed sandstones and mudstones. The paleoclimate of the Jurassic top red bed sediment in the study area is reconstructed based on elemental distribution patterns and characteristics of related elemental ratios. Advanced methods will be incorporated in future research for validation.
During the sedimentary process, with increasing weathering intensity, major elements such as Al, Fe, and Mn typically increase in concentration as oxides, whereas the concentrations of elements such as Ca, Mg, Na, and K tend to decrease gradually [66]. Therefore, previous studies often used major element oxide contents to infer the climate during sedimentary processes. The UCC normalization test results of major element oxides in the mudstone samples indicate that the average contents of Al2O3, Fe2O3, and MnO have increased compared with the standard values of the upper crust, although the changes are insignificant. Conversely, the average contents of CaO and Na2O decreased, whereas the K2O content remained relatively stable. The UCC normalization results for the sandstone samples show that the contents of CaO, Na2O, MgO, and K2O have decreased, but again, the reductions are unsubstantial. This phenomenon suggests that the red beds in the study area have undergone slight weathering. In the study area, the red bed samples exhibited relative enrichment of Fe and stable levels of Mn. Fe is prone to oxidation and subsequent precipitation under arid conditions, leading to its accumulation. The chemically stable Mn element is present at relatively constant concentrations and precipitates in large amounts only under conditions of intense evaporation. Therefore, the Fe/Mn ratio can be used as an indicator of paleoclimatic conditions. In the study area, the Fe/Mn ratio is relatively high, with average values of 47.4 for sandstone and 94.9 for mudstone, indicating an arid sedimentary environment.
The chemical index of alteration (CIA) and the chemical weathering index (CIW) are commonly used indicators for assessing the intensity of weathering source rocks [84]. The CIA and CIW values are derived from the relative ratios of oxides, with their calculation formulas provided in (1) and (2). The combined use of these indices allows for a more accurate assessment of the extent of chemical weathering experienced by sandstone and mudstone. CaO* refers to the calcium oxide content in silicates, and the method proposed by McLennan [66] in 1993 has been applied for its correction. The CIW index is calculated via the method introduced by Harnois et al. in 1988 [85], which aims to eliminate the influence of potassium exchange during diagenesis on the results:
C I A = m o l a r ( A l 2 O 3 ) m o l a r A l 2 O 3 + C a O * + N a 2 O + K 2 O × 100  
C I W = m o l a r ( A l 2 O 3 ) m o l a r A l 2 O 3 + C a O * + N a 2 O × 100  
The CIA values for the red sandstone samples range from 49.4 to 56.9, whereas the mudstone values range from 59.2 to 59.7, with average values of 52.2 and 59.5, respectively. These results are consistent with those of previous studies [38] and fall within the range of weak chemical weathering under arid climatic conditions. In addition, this study utilizes the A-CN-K ternary diagram (where A, CN, and K represent the molar contents of Al2O3, CaO* + Na2O, and K2O, respectively) to analyze the effects of potassium substitution during the diagenesis process [66,86,87]. As shown in Figure 11, the CIA values of the sample points range between 50 and 60, indicating that they have not experienced substantial weathering compared with the upper crust CIA standard value of 47. However, the sample points deviate from the weathering trend line parallel to A–CN, indicating that the red beds in the study area have undergone slight potassium alteration. To address this finding, the study employs a combined approach using the CIA and CIW indices. The CIW index, which is insensitive to potassium alteration (Table 3), is used to mitigate the impact of potassium alteration on the calculations during diagenesis. The average CIW value is 59 for the red sandstone samples but 69 for the mudstone samples, indicating a transition from weak chemical weathering under arid conditions to moderate chemical weathering under warm, humid conditions. Therefore, the comprehensive analysis suggests that the red beds experienced weak chemical weathering during the depositional period of the study area under arid climate conditions.
The distribution patterns of trace elements in rocks are important tools for inferring paleoclimate and paleoenvironmental conditions during sedimentation. The trace element analysis reveals that the average concentrations of Ba in the red sandstone and mudstone are 481.3 ppm and 695.3 ppm, respectively, whereas the average concentrations of Sr are 123.1 ppm and 210.8 ppm, respectively. This comparison indicates that Ba is more enriched than Sr is, with a ratio of approximately 1:4. Owing to the greater mobility of Sr than Ba, Ba is more likely to precipitate in high-salinity waters. This phenomenon suggests that the sedimentary water body had elevated salinity, indicating a relatively arid climate at that time.
Additionally, the contents of trace elements such as V, Cr, Co, Ni, and Zn in mudstone are sensitive indicators of the redox conditions of the sedimentary water body. Notably, the ratios of V/Cr, Ni/Co, and V/(V+Ni) have been widely used to determine the paleo-redox conditions of sedimentary environments [88,89]. The test data of the red beds indicate that V and Cr are more enriched than Co, Ni, and Zn. The ratio characteristics (Figure 12) reveal that in the Upper Jurassic formations, the distribution of the trace element ratio V/(V+Ni) ranges from 0.61 to 0.81, with an average value of 0.69. This value is close to the boundary between weakly oxidizing and oxidizing environments. The V/Cr and Ni/Co ratios range from 0.61 to 2.98 and 2.15 to 3.09, with average values of 1.32 and 2.74, respectively (Table 6). These values indicate an oxidizing environment. This finding suggests that during the deposition of the Upper Jurassic, the basin waters were predominantly oxidizing, leading to the oxidation of magnesium–iron silicate minerals and iron-bearing clay minerals, resulting in the release of Fe3⁺ and reddening of the sediments.
In paleoclimate interpretation, this study uses the chemical indices (CIA and CIW) to evaluate chemical weathering intensity during deposition and combines the Fe/Mn ratio for climate inference. The V/Cr, Ni/Co, and V/(V+Ni) ratios help deduce paleo-redox conditions, while the Sr/Ba ratio is used for paleosalinity assessment. Integrated analyses reveal an arid to semi-arid depositional environment with oxidizing water conditions in the study area. However, the study has limitations, such as a lack of analysis of paleobathymetry and paleotemperature indicators. In order to compensate for these shortcomings, the depth and temperature of ancient water bodies can be reconstructed using the carbon isotope (δ13C) and oxygen isotope (δ18O) values of the sedimentary water, as well as the mass fraction of the trace element strontium (Sr). In addition, the U/Th ratio and isotopes like δ7Li and δ98/95Mo are used to reconstruct redox conditions. The Sr/Cu ratio, a common indicator for climate humidity, can cross-validate weathering intensity indicators. Due to experimental limitations, this study lacks advanced geochemical indicators for verification, potentially causing some result deviations. In recent years, Tian [90] used spore–pollen assemblages to reconstruct the Jurassic paleoclimate of the Junggar Basin. Results show the climate shifted from warm/wet in the early Jurassic to hot/arid in the Late Jurassic. During Late Jurassic sedimentation, the climate was already relatively arid. Liu [91] used geochemical proxies like Rb/Sr, Sr/Cu, and U/Th to reconstruct the paleoclimate of the central area in the Junggar Basin during the Late Jurassic. Results show that the climate gradually turned from semi-arid to arid. In addition, Luo et al. [92] integrated geochemical proxies (e.g., Sr/Cu, U/Th), rare earth elements, and heavy mineral assemblages to reconstruct the Middle–Late Jurassic climate sequence in the Junggar Basin. Their reconstruction confirmed the shift from a warm/wet to an arid climate, and indicated that the Late Jurassic climate became dry. The paleoclimate reconstruction results in this study align with recent findings by other scholars on the Middle–Late Jurassic paleoclimate in the Junggar Basin. This consistency demonstrates the reliability of using geochemical indicators like CIA, CIW, V/Cr, and Ni/Co for paleoclimate reconstruction in this study.
Additionally, Liu et al. [93] found red sediments in their study of the modern Loess Plateau and suggested that this red sediment mainly deposits in arid to semi-arid regions. As rainfall increases, loess layers appear redder due to more hematite formation, creating red soils. The classic Luochuan section shows this. We compared geochemical indicators of modern Luochuan loess deposits and red beds in our study area, finding similarities [94,95]. This suggests that red bed sediment in the study area is consistent with modern red soil sediment in northern areas, both forming in semi-arid to arid oxidative aquatic environments.
The arid paleoclimate indicated by the Upper Jurassic red beds in the study area aligns with regional paleoclimate conditions. Global studies have shown that the Jurassic climate experienced significant fluctuations, with a sudden temperature drop in the early Middle Jurassic that persisted until the Late–Middle Jurassic, followed by a gradual rise, leading to the formation of a high-temperature platform in the Late Jurassic [96]. During the Early Jurassic, northern China experienced a lush, humid climate, which transitioned to a tropical semi-arid to arid climate by the Late Jurassic [97]. Therefore, the terrestrial red beds at the top of the Jurassic in the Yongjin area of the Junggar Basin directly reflect the Late Jurassic aridification event, allowing comparison with global climate changes.

5.3. Insights from Red Bed Sedimentary Facies

Previous studies on the Jurassic red bed sedimentary facies in the Yongjin area of the Junggar Basin have been relatively systematic. Fang et al. [60] and Fan [61] analyzed sediment core colors, sedimentary structures, and sequences and concluded that red beds primarily comprise lacustrine facies characterized by red siltstone and mudstone, as well as shallow lake facies dominated by red mudstone and gray-green sandstone. The main sedimentary environment is semi-arid to semi-humid lacustrine. The alternation of reddish-brown and gray sandstone reflects changes in paleoclimate between humid and arid conditions. Huang [53] and Song et al. [58] analyzed lithology characteristics, sedimentary structures, logging curves, and sedimentary numerical simulations. They concluded that the lower part of the red beds is characterized by a shallow-water braided river delta face, whereas the upper part, due to a reduced sediment supply, features a shallow-water meandering river delta face. These authors suggested that the sedimentation of red beds occurred during a period of persistent aridity, with water turbulence being the reason for the alternation of red and gray sandstones. When the water level was low, the sediments were exposed to the surface and underwent oxidation, resulting in the formation of red sand bodies. The findings of this study on element distribution patterns suggest that the sedimentation environment of the red layer in the study area was oxidizing and arid, which was conducive to the formation of hematite but not gray sandstone. The CIA and CIW calculations revealed that the sedimentary rocks experienced mild chemical weathering due to arid conditions, indicating a persistently dry climate during sedimentation, aligning with regional climatic changes. Additionally, macrocore observations reveal that the red beds are primarily composed of red sandstone and mudstone, with no other colored mudstones. The interbedded gray sandstone contains no plant fossils. The sedimentary structures of gray and red sandstones are similar and are characterized primarily by parallel lamination, planar cross-lamination, and ripple lamination. These features indicate shallow water conditions and high flow velocities during deposition, suggesting a fluvial depositional environment. These characteristics suggest an arid climate and a predominantly terrestrial environment with slight water depth variation during red bed sedimentation. Moreover, we observed widespread indications of oil in the gray sandstone, which we speculate is a result of petroleum intrusion and bleaching processes [3,98], rather than variations in sedimentary water depth. In summary, our analysis inferred that the top of the Jurassic red beds in the study area was characterized by gentle paleotopography, a persistently arid paleoclimate, shallow sedimentary water primarily located on land, and a proximal source, indicative of an arid deltaic plain sedimentary environment.

6. Conclusions

This study utilizes core samples collected from boreholes and employs thin section analysis, scanning electron microscopy, clay mineral X-ray diffraction, and major–trace element analysis to investigate the formation period, paleoclimate, and sedimentary facies of the Upper Jurassic red beds in the Yongjin area, as well as the genesis of the red beds. The results indicate the following.
(1)
Red beds formed during the earliest diagenetic stage, either during sedimentation or shortly after, with the growth of authigenic cement occurring later than the development of hematite.
(2)
The distribution patterns of major and trace elements, along with core observations, indicate that the paleostructure was gentle during the deposition of the red beds. The paleoclimate remained persistently arid, with shallow, oxygen-rich depositional waters, characterizing the overall sedimentation as belonging to an arid deltaic plain.
(3)
Petrographic analysis revealed that the iron in hematite mainly comes from the transformation of volcanic rock fragments during the syn-sedimentary and early diagenetic stages, especially those unstable iron-bearing silicates (ilmenite, magnetite, olivine, pyroxene). The conversion of montmorillonite to illite is a secondary source of iron. This study can provide new insights for understanding global red bed sedimentary mechanisms and late Jurassic paleoclimate reconstruction. Notably, the grey organic-rich sandstone interspersed in the red bed sequence may pose a potential interfering factor. Lithofacies and geochemical evidence show that these grey sandstones and adjacent red beds are not significantly different in initial sedimentary environment conditions. We propose a new genetic model: the anomalous coloration of gray sandstone may be due to secondary bleaching caused by hydrocarbon fluid migration in the later diagenetic stage. In order to test this hypothesis, follow-up research plans focus on the geochemical mechanism of the bleaching process and its spatiotemporal evolution.

Author Contributions

Conceptualization, Y.G. and L.Z.; Methodology, Y.G., L.Z., and Y.L. (Yuhong Lei); Validation, L.Z. and C.H.; Formal analysis, Y.G., Z.Z. and B.X.; Investigation, L.Y. (Yan Li), B.X. and N.L.; Resources, Y.G. and N.L.; Data curation, Y.G. and C.H.; Writing—original draft preparation, Y.G. and L.Z.; Writing—review and editing, L.Z. and C.L.; Visualization, Y.G., Y.J., and Y.L. (Yan Li); Supervision, L.Z. and Y.L. (Yuhong Lei); Funding Acquisition, L.Z. and C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the National Natural Science Foundation of China (42030808) and the SINOPEC Ministry of Science and Technology Project (P24167, 30200018-22-ZC0613-0079).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Authors Chao Li and Lan Yu are employed by the company Petroleum Exploration and Production Research Institute, China Petroleum & Chemical Corporation. The authors declare no conflicts of interest.

References

  1. Parry, W.T.; Chan, M.A.; Nash, B.P. Diagenetic characteristics of the Jurassic Navajo Sandstone in the Covenant oil field, central Utah thrust belt. AAPG Bull. 2009, 93, 1039–1061. [Google Scholar] [CrossRef]
  2. Leanza, H.A.; Hugo, C.A. Cretaceous red beds from southern NeuquénBasin (Argentina): Age, distribution and stratigraphic discontinuities. In VII International Symposium on Mesozoic Terrestrial Ecosystems; 7º International Symposium on Mesozoic Terrestrial Ecosystems; Asociación Paleontológica Argentina: Buenos Aires, Argentina, 2001; Publicación Especial 7; pp. 117–122, (Abstract). [Google Scholar]
  3. Aehnelt, M.; Hilse, U.; Pudlo, D.; Heide, K.; Gaupp, R. On the origin of bleaching phenomena in red bed sediments of Triassic Buntsandstein deposits in Central Germany. Geochemistry 2021, 81, 125736. [Google Scholar] [CrossRef]
  4. Eren, M.; Kadir, S. Colour origin of red sandstone beds within the Hüdai Formation (Early Cambrian), Aydıncık (Mersin), southern Turkey. Turk. J. Earth Sci. 2013, 22, 563–573. [Google Scholar] [CrossRef]
  5. Xie, D.; Yao, S.; Cao, J.; Hu, W.; Wang, X.; Zhu, N. Diagenetic alteration and geochemical evolution during sandstones bleaching of deep red-bed induced by methane migration in petroliferous basins. Mar. Pet. Geol. 2021, 127, 104940. [Google Scholar] [CrossRef]
  6. Houten, F.B.V. Origin of Red Beds A Review-1961–1972. Annu. Rev. Earth Planet. Sci. 1973, 1, 39–61. [Google Scholar] [CrossRef]
  7. Tsirambides, A. Genesis and physical characteristics of the neogene red beds from the cedarhils of thessaloniki, macedonia, Greece. Bull. Geol. Soc. Greece 2018, 36, 708–716. [Google Scholar] [CrossRef]
  8. Eren, M.; Kadir, S. Colour origin of upper cretaceous pelagic red sediments within the Eastern Pontides, northeast Turkey. Int. J. Earth Sci. 1999, 88, 593–595. [Google Scholar] [CrossRef]
  9. Turner, P. Diagenetic origin of Cambrian marine red beds: Caerfai Bay Shales, Dyfed, Wales. Sediment. Geol. 1979, 24, 269–281. [Google Scholar] [CrossRef]
  10. Frostick, L.E.; Reid, I. Tectonic control of desert sediments in rift basins ancient and modern. Geol. Soc. Lond. Spec. Publ. 1987, 35, 53–68. [Google Scholar] [CrossRef]
  11. Metcalfe, R.; Rochelle, C.A.; Savage, D.; Higgo, J.W. Fluid-rock interactions during continental red bed diagenesis: Implications for theoretical models of mineralization in sedimentary basins. Geol. Soc. Lond. Spec. Publ. 1994, 78, 301–324. [Google Scholar] [CrossRef]
  12. Lai, J.; Li, D.; Wang, G.; Xiao, C.; Cao, J.; Wu, C.; Han, C.; Zhao, X.; Qin, Z. Can carbonate cementation be inhibited in continental red bed sandstone? J. Pet. Sci. Eng. 2019, 179, 1123–1135. [Google Scholar] [CrossRef]
  13. He, W.; Yang, Z.; Du, H.; Hu, J.; Zhang, K.; Hou, W.; Li, H. Micro-Mechanisms and Implications of Continental Red Beds. Minerals 2022, 12, 934. [Google Scholar] [CrossRef]
  14. Xu, Q.; Wang, Y.; Ma, L.; Yue, Y.; Meng, T.; Bi, J.; Jiang, C. Paleoclimate quantitative reconstruction and characteristics of continental red beds: A case study of the lower fourth sub-member of Shahejie Formation in the Bonan Sag. J. Pet. Explor. Prod. Technol. 2023, 13, 1993–2014. [Google Scholar] [CrossRef]
  15. Kirkham, R. Distribution, settings, and genesis of sediment-hosted stratiform copper deposits. Geol. Assoc. Can. Spec. Pap. 1989, 36, 3–38. [Google Scholar]
  16. Hölting, B. Geogene Grundwasserbeschaffenheit und ihre regionale Verbreitung in der Bundesrepublik Deutschland. Handb. Bodenschutz 1991, 1, 1300. [Google Scholar]
  17. Tesmer, M.; Möller, P.; Wieland, S.; Jahnke, C.; Voigt, H.; Pekdeger, A. Deep reaching fluid flow in the North East German Basin: Origin and processes of groundwater salinisation. Hydrogeol. J. 2007, 15, 1291–1306. [Google Scholar] [CrossRef]
  18. Rödiger, T.; Sauter, M.; Büchel, G. Infiltration und Grundwasserströmung in geklüftet-porösen Buntsandsteingrundwasserleitern im Osten des Thüringer Beckens. Grundwasser 2009, 14, 21–32. [Google Scholar] [CrossRef]
  19. Tomlinson, C.W. The Origin of Red Beds: A Study of the Conditions of Origin of the Permo-Carboniferous and Triassic Red Beds of the Western United States. J. Geol. 1916, 24, 153–179. [Google Scholar] [CrossRef]
  20. Walker, T.R. Formation of Red Beds in Modern and Ancient Deserts. GSA Bull. 1967, 78, 353–368. [Google Scholar] [CrossRef]
  21. Walker, T.R. Diagenetic origin of continental red beds. In The Continental Permain in Central, West, and South Europe: Proceedings of the NATO Advanced Study Institute held at the Johannes Gutenberg University, Mainz, FRG, 23 September–4 October 1975; Springer: Berlin/Heidelberg, Germany, 1976; pp. 240–282. [Google Scholar]
  22. Franke, W.; Paul, J. Pelagic redbeds in the Devonian of Germany—deposition and diagenesis. Sediment. Geol. 1980, 25, 231–256. [Google Scholar] [CrossRef]
  23. Goudie, A.S.; Pye, K. (Eds.) Chemical Sediments and Geomorphology: Precipitates and Residua in the Nearsurface Environment; Academic Press: London, UK; New York, NY, USA, 1983; 439p, ISBN 0122934806. [Google Scholar]
  24. Krynine, P.D. Section of geology and mineralogy: The origin of red beds. Trans. N. Y. Acad. Sci. 1949, 11, 60–68. [Google Scholar] [CrossRef]
  25. Folk, R.L. Reddening of desert sands; Simpson Desert, NT, Australia. J. Sediment. Res. 1976, 46, 604–615. [Google Scholar] [CrossRef]
  26. Besly, B.; Turner, P. Origin of red beds in a moist tropical climate (Etruria Formation, Upper Carboniferous, UK). Geol. Soc. Lond. Spec. Publ. 1983, 11, 131–147. [Google Scholar] [CrossRef]
  27. Miot, J.; Benzerara, K.; Morin, G.; Kappler, A.; Bernard, S.; Obst, M.; Férard, C.; Skouri-Panet, F.; Guigner, J.-M.; Posth, N.; et al. Iron biomineralization by anaerobic neutrophilic iron-oxidizing bacteria. Geochim. Cosmochim. Acta 2009, 73, 696–711. [Google Scholar] [CrossRef]
  28. Jiang, L.; Chen, G.; Grapes, R.; Peng, Z. Thermal origin of continental red beds in SE China: An experiment study. J. Asian Earth Sci. 2015, 101, 14–19. [Google Scholar] [CrossRef]
  29. Zhang, L.; Liu, C.; Zhang, S.; Fayek, M.; Lei, K.; Quan, X. Unconformity-controlled bleaching of Jurassic-Triassic sandstones in the Ordos Basin, China. J. Pet. Sci. Eng. 2022, 211, 110154. [Google Scholar] [CrossRef]
  30. Shahrokhi, S.; Bábek, O.; Mansurbeg, H.; Ackerman, L.; Šimíček, D.; Gregar, F.; Kořenek, M. Investigating the impact of paleoclimatic conditions and diagenesis on the genesis of Permian Continental Red Beds: A case study from the Bohemian Massif, Czechia. Sediment. Geol. 2024, 470, 106723. [Google Scholar] [CrossRef]
  31. Liu, M.; Ji, C.; Huang, Y.; Ding, Y.; Chen, R.; Chen, M.; Yang, B.; Chen, D. Coloration and Environmental Significance of the Marine Red Bed from the Sowa Formation Carbonate in the Qiangtang Basin. Acta Sedimentol. Sin. 2024, 42, 812–822. [Google Scholar]
  32. Jiang, H.; Xia, Q.; Liu, S.; Zhang, X.; Li, J.; Wang, Y. Investigation of Weathering Intensity and Color Genesis of Terrestrial Sediments: A Case Study of the Upper Shaximiao Formation in the Middle Jurassic Sichuan Basin. J. Sedimentol. 2023, 41, 1025–1039. [Google Scholar] [CrossRef]
  33. Guo, X.; Hu, D.; Li, Y.; Duan, J.; Zhang, X.; Fan, X.; Duan, H.; Li, W. Theoretical Advancements and Key Technologies in Onshore Ultra-Deep Oil and Gas Exploration. Engineering 2019, 5, 233–258. [Google Scholar] [CrossRef]
  34. Guan, X.; Guo, L. New Advances and Prospects in Research on Reservoir Formation of Deep and Ultra-Deep Oil and Gas. Pet. Exp. Geol. 2023, 45, 203–209. [Google Scholar]
  35. Cao, Y.; Yuan, G.; Yang, H.; Wang, Y.; Liu, K.; Zan, N.; Xi, K.; Wang, J. Current Status of Oil and Gas Exploration in Deep and Ultra-Deep Clastic Rock Basins and Research Progress on the Genesis of Quality Reservoirs. J. Petroleum. 2022, 43, 112–140. [Google Scholar]
  36. Ma, Y.; Cai, X.; Zhao, P. Characteristics and Formation Mechanisms of Reef Beach Facies Reservoirs in the Changxing and Feixiguan Formations of the Yuanba Gas Field. J. Petroleum. 2014, 35, 1001–1011. [Google Scholar]
  37. Jinag, X.; Guan, H.; Zheng, H.; Wang, S.; Fan, M.; Bao, Y.; Li, F.; Han, Y. Discussion on the Genesis of Dolomite Reservoirs in the Feixianguan Formation of the Puguang Gas Field in the Sichuan Basin. J. Pet. Exp. Geol. 2014, 36, 332–336+345. [Google Scholar]
  38. He, Z. The Genesis of the Jurassic Red Beds in the Yongjin Area of the Junggar Basin. Xinjiang Pet. Geol. 2009, 30, 133–135. [Google Scholar]
  39. Ding, X.; Shen, Y.; Zhao, Y.; Zhang, Y. The Genesis of the Jurassic Red Beds in the Yongjin Area of the Junggar Basin. Pet. Geol. Recovery Effic. 2009, 16, 12–14+111. [Google Scholar]
  40. Jiao, Y.; Wu, L.; Lu, Y.; Ren, J. The Diagenetic Processes of the Red Beds at the Top of the Jurassic System in the Central Part of the Junggar Basin and the Evolutionary Information of the Chepaizi-Mosuowan Paleouplift. Earth Sci. (J. China Univ. Geosci.) 2008, 33, 219–226. [Google Scholar]
  41. Li, Z.; Yang, M.; Tian, Z. The Origin and Geological Significance of the Continental Red Beds of the Qi Gu Formation in the Yongjin Area, Southern Margin of the Junggar Basin. Xinjiang Geol. 2023, 41, 577–585. [Google Scholar]
  42. Zhao, B. The formation and evolution of the Junggar Basin. Xinjiang Pet. Geol. 1992, 13, 191–196. [Google Scholar]
  43. Zhao, B. The Structural Characteristics and Tectonic Division of the Junggar Basin. Xinjiang Pet. Geol. 1993, 14, 209–216. [Google Scholar]
  44. Wang, H. The Structural Characteristics of the Zhungarian Oil and Gas Basin in Xinjiang. Xinjiang Pet. Geol. 1983, 4, 1–17. [Google Scholar]
  45. Zhang, L.; Li, C.; Luo, X.; Zhang, Z.; Zeng, Z.; Ren, X.; Lei, Y.; Zhang, M.; Xie, J.; Cheng, M.; et al. Vertically transferred overpressures along faults in Mesozoic reservoirs in the central Junggar Basin, northwestern China: Implications for hydrocarbon accumulation and preservation. Mar. Pet. Geol. 2023, 150, 106152. [Google Scholar] [CrossRef]
  46. Zhou, J.; Li, C.; Song, Z.; Zhang, X. Organic Geochemical Characteristics and Hydrocarbon Significance of the Permian System Around the Bogda Mountain, Junggar Basin, Northwest China. Sustainability 2025, 17, 347. [Google Scholar] [CrossRef]
  47. Li, C.; Zhang, L.; Luo, X.; Zeng, Z.; Xiu, J.; Lei, Y.; Cheng, M.; Hu, C.; Zhang, M.; He, W. Clay mineral transformations of mesozoic mudstones in the central Junggar Basin, northwestern China: Implications for compaction properties and pore pressure responses. Mar. Pet. Geol. 2022, 144, 105847. [Google Scholar] [CrossRef]
  48. Zhang, Z.; Zhang, L.; Yan, Y.; Zhang, L. Application of Sedimentary Forward Simulation in Predicting the Distribution of Braided River Delta Sand Bodies in the Xishanyao Formation of the Yongjin Area, Junggar Basin. Nat. Gas Geosci. 2024, 35, 2064–2077. [Google Scholar]
  49. Wu, X.; Zhang, N.; Shi, X.; Zhou, L.; Zhang, Y.; Yang, G.; Guan, S. Structural Characteristics and Hydrocarbon Accumulation Model of the Chemo Uplift in the Junggar Basin. China Pet. Explor. 2006; 11, 65–68+84. [Google Scholar]
  50. Zhao, H. Evolution of the Chemo Uplift and Its Hydrocarbon Accumulation Control in the Junggar Basin. Xinjiang Pet. Geol. 2006, 160–162. [Google Scholar]
  51. Qiao, Y.; Sui, F.; Lin, H.; Song, C. The Control of the Che-Mo Paleouplift on Oil and Gas in the Central Part of the Junggar Basin. J. Pet. Nat. Gas Geol. 2013, 35, 56–61. [Google Scholar]
  52. Jin, X. Sedimentary Reservoir Study of the Jurassic Qigu Formation in the Yongjin Area of the Junggar Basin. Master’s Thesis, China University of Petroleum (Beijing), Beijing, China, 2023. [Google Scholar] [CrossRef]
  53. Huang, C.; Zhu, X.; Jin, X.; Hu, X.; Xiu, J.; Ren, X.; Cheng, C. Quantitative Characterization of the Impact of Diagenesis on Reservoir Quality of Deep-Buried Sandstones in the Qigu Formation in the Yongjin Area of the Junggar Basin. J. Palaeogeogr. 2024, 26, 683–699. [Google Scholar]
  54. Luo, Y. Characteristics and Geological Significance of Seismic Sedimentary Rocks in the Jurassic Qigu Formation in the Yongjin Area of the Junggar Basin. Low Carbon World. 2024, 14, 121–123. [Google Scholar] [CrossRef]
  55. Xie, Y.; Li, H.; Sun, Z. Discovery of a Weathering Crust between Jurassic and Cretaceous and Its Stratigraphic Significance in the Shinan Area of the Junggar Basin. Geol. Rev. 2006, 52, 137–144. [Google Scholar] [CrossRef]
  56. Yang, Z.; He, S.; Wu, H.; Meng, X. Evidence of Existence and Research Significance of the Toutunhe Formation in the Middle Segment of the Che-Mo Paleouplift Within the Central Part of the Junggar Basin. Acta Geosci. Sin. 2008, 52, 469–477. [Google Scholar]
  57. Liu, M.; Tian, Y.; Yang, M.; Wang, X.; Huang, H.; Li, G. Sedimentary Characteristics and Depositional Models of the Top Part of the Xishanyao Formation in the Yongjin Area, Junggar Basin. J. East China Inst. Technol. (Nat. Sci. Ed.) 2021, 44, 590–601. [Google Scholar]
  58. Song, F.; Kong, Q.; Zhang, X.; Cao, H.; Jiao, G.; Yang, Y. Sedimentary Characteristics and Depositional Models of Arid Shallow Water Deltas: A Case Study of the Jurassic Qigu Formation in the Yongjin Area, Central Part of the Junggar Basin. Pet. Geol. Gas Geol. 2024, 45, 1–14. [Google Scholar]
  59. Qiao, D.; Jin, X.; Shi, G.; Hong, Y.; Shi, Z. Significance of Jurassic Qigu Formation Continental Trace Fossils and Paleoenvironments in the Northern Tianshan Region of Xinjiang. J. Chengdu Univ. Technol. (Nat. Sci. Ed.) 2023, 50, 268–278. [Google Scholar]
  60. Fang, S.; Guo, Z.; Song, Y.; Wu, C.; Zhang, Z.; Wang, M.; Fan, R. Jurassic Sedimentary Facies Evolution and Basin Configuration in the Southern Margin of the Junggar Basin. J. Palaeogeogr. 2005, 7, 347–356. [Google Scholar]
  61. Fan, Y. Sedimentary Facies and Depositional Environment of the Upper Jurassic Qigu Formation in the Southern Margin of the Junggar Basin, Xinjiang. Master’s Thesis, Chang’an University, Xi’an, China, 2021. [Google Scholar] [CrossRef]
  62. Zhang, D.-P.; Xue, D.-S.; Liu, Y.-H.; Wan, B.; Guo, Q.; Guo, J.-J. Comparative Study of Three Mixing Methods in Fusion Technique for Determining Major and Minor Elements Using Wavelength Dispersive X-ray Fluorescence Spectroscopy. Sensors 2020, 20, 5325. [Google Scholar] [CrossRef]
  63. Folk, R.L. Petrology of Sedimentary Rocks; Hemphill Pub. Co., Hemphill Publishing Company: Austin, TX, USA, 1980. [Google Scholar]
  64. Dickinson, W.R.; Suczek, C.A. Plate tectonics and sandstone compositions. Aapg Bull. 1979, 63, 2164–2182. [Google Scholar]
  65. Taylor, S.R.; Mclennan, S.M. The continental crust: Its composition and evolution. J. Geol. 1985, 94, 57–72. [Google Scholar]
  66. McLennan, S.M. Weathering and Global Denudation. J. Geol. 1993, 101, 295–303. [Google Scholar] [CrossRef]
  67. Eren, M.; Kadir, S.; Kapur, S.; Huggett, J.; Zucca, C. Colour origin of Tortonian red mudstones within the Mersin area, southern Turkey. Sediment. Geol. 2015, 318, 10–19. [Google Scholar] [CrossRef]
  68. Hu, X.; Wang, S.; Li, X.; Jansa, L. The Paleogene Ocean Red Beds in Southern Tibet: Rock Types, Sedimentary Environments, and Color Genesis. Sci. China: Earth Sci. 2006, 36, 811–821. [Google Scholar]
  69. Liu, X.; Meng, J.; Xie, Y.; Li, B.; Chi, Y.; Kang, C.; Sun, L.; Wei, Z.; Liu, H.; Wang, Y.; et al. Color Mechanisms of Continental Red Beds in the Luojiawopeng Formation at the Southeastern Edge of the Songnen Plain. Acta Geosci. Sin. 2024, 1–12. Available online: http://kns.cnki.net/kcms/detail/11.3474.P.20250205.1537.002.html (accessed on 10 February 2025).
  70. Sheldon, N.D. Do red beds indicate paleoclimatic conditions?: A Permian case study. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 228, 305–319. [Google Scholar] [CrossRef]
  71. Zhu, Y.; Wang, L.; Zhang, F. Main Controlling Factors of Hidden Hydrocarbon Accumulations in the Yongjin Area. Nat. Gas Ind. 2008, 28, 28–31+138–139. [Google Scholar]
  72. Xiu, J.; Wang, X.; Liu, S.; Miao, H.; Ding, J. Reservoir Characteristics and Influencing Factors of the Xishanyao Formation in the Yongjin Area, Junggar Basin. Pet. Geol. Eng. 2015, 29, 28–30+145–146. [Google Scholar]
  73. Lan, X. Discussion on Clay Minerals as Paleoclimatic Indicator Minerals. Geol. Sci. Technol. Inf. 1990, 9, 31–35. [Google Scholar]
  74. Tang, Y.; Jia, J.; Xie, X. The Environmental Significance of Clay Minerals. Earth Sci. Front. 2002, 9, 94–101. [Google Scholar]
  75. Dong, H.; Song, Y. Application of Clay Minerals in Paleoenvironmental Reconstruction. Mar. Geol. Quat. Geol. 2009, 29, 119–130. [Google Scholar] [CrossRef]
  76. Song, S.; Liu, Z.; Sun, P.; Wang, Y.; Xie, W.; Song, Q. Geochemical Characteristics and Provenance Analysis of Red Mudstone Sequences: A Case Study of the Upper Cretaceous Yaojia Formation in the Southeastern Uplift Area of the Songliao Basin. World Geol. 2015, 34, 774–785. [Google Scholar]
  77. Wang, W.; Li, S.; Zhang, Y.; Yuan, Y.; Pang, Z.; Zhao, J. Characteristics of Greenhouse Climate and Environmental Changes during the Middle to Late Jurassic in the Lufeng-Chuanjie Basin, Central Yunnan, and Their Impact on Biological Preservation. J. Jilin Univ. (Earth Sci. Ed.) 2024, 54, 1530–1543. [Google Scholar] [CrossRef]
  78. Land, J.S.; Tsikos, H.; Cousins, D.; Luvizotto, G.; Zack, T. Origin of red beds and paleosols in the Palaeoproterozoic Transvaal and Olifansthoek Supergroups of South Africa: Provenance versus metasomatic controls. Geol. J. 2018, 53, 191–202. [Google Scholar] [CrossRef]
  79. Garzanti, E.; Andò, S.; Padoan, M.; Vezzoli, G.; El Kammar, A. The modern Nile sediment system: Processes and products. Quat. Sci. Rev. 2015, 130, 9–56. [Google Scholar] [CrossRef]
  80. Boboye, O.A.; Nwosu, O.R. Petrography and geochemical indices of the Lagos lagoon coastal sediments, Dahomey Basin (southwestern Nigeria): Sea level change implications. Quat. Int. 2014, 338, 14–27. [Google Scholar] [CrossRef]
  81. Sun, J.; Fu, Y.; Lan, C. Trace Element Geochemical Characteristics of Clastic Rocks in Baxiandun, Qingdao, and Their Implications for Sedimentary Environment. Mar. Sci. 2014, 38, 75–81. [Google Scholar]
  82. Song, L.; Liu, C.; Zhao, H.; Wang, J.; Zhang, X. Geochemical Characteristics of the Huangqikou Formation and Their Implications for Depositional Environment and Tectonic Background in the Ordos Region. Earth Sci. 2016, 41, 1295–1308+1321. [Google Scholar]
  83. Chen, G.; Li, C.; Li, H.; Cha, K. Geochemical Characteristics of Continental Red Beds from the Upper Cretaceous in the Northern Guangdong Danxia Basin and Their Implications for Paleoclimate. Publ. J. Sedimentol. 2025, 1–32. [Google Scholar] [CrossRef]
  84. Price, J.R.; Velbel, M.A. Chemical weathering indices applied to weathering profiles developed on heterogeneous felsic metamorphic parent rocks. Chem. Geol. 2003, 202, 397–416. [Google Scholar] [CrossRef]
  85. Harnois, L. The CIW index: A new chemical index of weathering. Sediment. Geol. 1988, 55, 319–322. [Google Scholar] [CrossRef]
  86. Feng, L.; Chu, L.; Zhang, Q.; Zhang, T. Chemical Index of Alteration (CIA) and Its Application in Neoproterozoic Clastic Rocks. Earth Sci. Front. 2003, 10, 539–544. [Google Scholar]
  87. Xu, X.; Shao, L. Limiting Factors in Analyzing the Degree of Weathering in Source Areas Using Clay Mineral Chemical Index of Alteration. J. Palaeogeogr. 2018, 20, 515–522. [Google Scholar]
  88. Hatch, J.R.; Leventhal, J.S. Relationship between inferred redox potential of the depositional environment and geochemistry of the Upper Pennsylvanian (Missourian) Stark Shale Member of the Dennis Limestone, Wabaunsee County, Kansas, USA. Chem. Geol. 1992, 99, 65–82. [Google Scholar] [CrossRef]
  89. Jones, B.; Manning, D.A.C. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chem. Geol. 1994, 111, 111–129. [Google Scholar] [CrossRef]
  90. Tian, Y. Palynological Assemblages and Paleoclimate Studies of the Jurassic in the Central and Western Regions of the Junggar Basin. Master’s Thesis, China University of Geosciences (Beijing), Beijing, China, 2017. [Google Scholar]
  91. Liu, D. Paleoclimate Analysis of the Jurassic Shihuogou Formation in the Fukang Depression of the Junggar Basin. Sci. Technol. Inf. 2019, 17, 75–76+78. [Google Scholar] [CrossRef]
  92. Luo, L.; Gao, X.; Tan, X.; Gluyas, J.; Wang, J.; Kong, X.; Huang, J.; Shao, H.; Qu, F. Paleo-environment and provenance in a lacustrine shallow-water delta-meandering river sedimentary system: Insights from the Middle–Upper Jurassic formations of the Fukang Sag of Junggar Basin, NW China. Aust. J. Earth Sci. 2019, 66, 699–722. [Google Scholar] [CrossRef]
  93. Liu, X.; Lv, B.; Mao, X.; Wen, C.; Yu, M.; Guo, X.; Chen, J.; Wang, T. Variations of Iron Minerals in Eolian Deposits with Environmental Changes and Their Implications. Quat. Res. 2014, 34, 443–457. [Google Scholar]
  94. Yang, L. A Study on the Sedimentary Characteristics and Formation Mechanisms of Quaternary Red Soils in the Middle and Lower Reaches of the Yangtze River. Ph.D. Thesis, East China Normal University, Shanghai, China, 2017. [Google Scholar]
  95. Peng, S.; Guo, Z. Similarities and differences in clay mineral composition between aeolian three-toed horse laterite and Quaternary loess and their environmental significance. In Proceedings of the 2007 Annual Conference of Institute of Geology and Geophysics; Chinese Academy of Sciences: Beijing, China, 2008; p. 9. [Google Scholar]
  96. Dera, G.; Brigaud, B.; Monna, F.; Laffont, R.; Pucéat, E.; Deconinck, J.-F.; Pellenard, P.; Joachimski, M.M.; Durlet, C. Climatic ups and downs in a disturbed Jurassic world. Geology 2011, 39, 215–218. [Google Scholar] [CrossRef]
  97. Deng, S.; Lu, Y.; Zhao, Y.; Fan, R.; Wang, Y.; Yang, X.; Li, X.; Sun, B. Paleoclimatic Zonation and Evolution of the Jurassic in China. Earth Sci. Front. 2017, 24, 106–142. [Google Scholar] [CrossRef]
  98. Beitler, B.; Chan, M.A.; Parry, W.T. Bleaching of Jurassic Navajo Sandstone on Colorado Plateau Laramide highs: Evidence of exhumed hydrocarbon supergiants? Geology 2003, 31, 1041–1044. [Google Scholar] [CrossRef]
Figure 1. Structural location map of the Yongjin area in the Junggar Basin (a); well location map of the study area (b); and stratigraphic development characteristics map of the Yongjin area (c).
Figure 1. Structural location map of the Yongjin area in the Junggar Basin (a); well location map of the study area (b); and stratigraphic development characteristics map of the Yongjin area (c).
Minerals 15 00347 g001
Figure 2. Comparison and correlation of Jurassic strata in the Yongjin area of the Junggar Basin. (The affiliation of the red beds is currently uncertain; therefore, we denote it as J2t?)
Figure 2. Comparison and correlation of Jurassic strata in the Yongjin area of the Junggar Basin. (The affiliation of the red beds is currently uncertain; therefore, we denote it as J2t?)
Minerals 15 00347 g002
Figure 3. Core column diagram of the Upper Jurassic strata from well Y6 in the Yongjin area, Junggar Basin.
Figure 3. Core column diagram of the Upper Jurassic strata from well Y6 in the Yongjin area, Junggar Basin.
Minerals 15 00347 g003
Figure 4. Ternary diagram of rock types in Upper Jurassic red beds in the Yongjin area of the Junggar Basin. Qtz—Quartz; Fsp—Feldspar; Rf—Rock Fragments; Srf—Sedimentary Rock Fragments; Vrf—Volcanic Rock Fragments; Mrf—Metamorphic Rock Fragments.
Figure 4. Ternary diagram of rock types in Upper Jurassic red beds in the Yongjin area of the Junggar Basin. Qtz—Quartz; Fsp—Feldspar; Rf—Rock Fragments; Srf—Sedimentary Rock Fragments; Vrf—Volcanic Rock Fragments; Mrf—Metamorphic Rock Fragments.
Minerals 15 00347 g004
Figure 5. QFL and QmFLt diagrams for the provenance of quartzose sandstones (after Dickinson et al. [64]). Q—Total quartzose grains; F—Monocrystalline feldspar grains; L—Unstable polycrystalline lithic fragments of either igneous or sedimentary parentage, including metamorphic varieties; Qm—Monocrystalline quartz; Lt—Total polycrystalline lithic fragments.
Figure 5. QFL and QmFLt diagrams for the provenance of quartzose sandstones (after Dickinson et al. [64]). Q—Total quartzose grains; F—Monocrystalline feldspar grains; L—Unstable polycrystalline lithic fragments of either igneous or sedimentary parentage, including metamorphic varieties; Qm—Monocrystalline quartz; Lt—Total polycrystalline lithic fragments.
Minerals 15 00347 g005
Figure 6. Microscopic lithic development characteristics of red beds sandstone in the Yongjin area of the Junggar Basin. (a) Medium andesite fragments, porphyry is plagioclase (+) from well YJ1-3 at 5977.99 m; (b) extensive hematite mineralization alteration occurred in the andesite fragments dark matrix (R) from well YJ1-3 at 5977.99 m; (c) medium andesite fragments feldspar porphyry directional arrangement (+) from well YJ1-3 at 5977.99 m; (d) mafic basalt fragments develop a large number of long tabular plagioclase porphyry (+) from well YJ1-1 at 5832.92 m; (e) large-scale hematite mineralization alteration occurs in the dark matrix in the basalt fragments feldspar triangular pores (R) from well YJ1-1 at 5832.92 m; (f) altered mafic diabase fragments (+) from well YJ1-1 at 5832.92 m; (g) medium andesite fragments on the left and altered mafic diabase fragments on the right (+) from well YJ1-3 at 5977.99 m; (h) plate-columnar ilmenite and acidic extrusive rock fragments are developed, and the porphyry crystals are felsic, and some metamorphic rock fragments can be seen (+) from well YJ1-1 at 5832.92 m; (i) hematite alteration occurs in ilmenite parts, and hematite mineralization spots develop on the surface of acidic extrusive rock fragments (R) from well YJ1-1 at 5832.92 m. Bf—Basalt fragments; Af—Andesite fragments; Aerf—Acidic extrusive rock fragments; Dbf—Diabase fragments; Ilm—Ilmenite; Hem—Hematite.
Figure 6. Microscopic lithic development characteristics of red beds sandstone in the Yongjin area of the Junggar Basin. (a) Medium andesite fragments, porphyry is plagioclase (+) from well YJ1-3 at 5977.99 m; (b) extensive hematite mineralization alteration occurred in the andesite fragments dark matrix (R) from well YJ1-3 at 5977.99 m; (c) medium andesite fragments feldspar porphyry directional arrangement (+) from well YJ1-3 at 5977.99 m; (d) mafic basalt fragments develop a large number of long tabular plagioclase porphyry (+) from well YJ1-1 at 5832.92 m; (e) large-scale hematite mineralization alteration occurs in the dark matrix in the basalt fragments feldspar triangular pores (R) from well YJ1-1 at 5832.92 m; (f) altered mafic diabase fragments (+) from well YJ1-1 at 5832.92 m; (g) medium andesite fragments on the left and altered mafic diabase fragments on the right (+) from well YJ1-3 at 5977.99 m; (h) plate-columnar ilmenite and acidic extrusive rock fragments are developed, and the porphyry crystals are felsic, and some metamorphic rock fragments can be seen (+) from well YJ1-1 at 5832.92 m; (i) hematite alteration occurs in ilmenite parts, and hematite mineralization spots develop on the surface of acidic extrusive rock fragments (R) from well YJ1-1 at 5832.92 m. Bf—Basalt fragments; Af—Andesite fragments; Aerf—Acidic extrusive rock fragments; Dbf—Diabase fragments; Ilm—Ilmenite; Hem—Hematite.
Minerals 15 00347 g006
Figure 7. Microscopic characteristics of the top red beds sandstone of the Jurassic in the Yongjin area of the Junggar Basin. (a) Development of abundant hematite films and iron-rich clay matrices (−) from well YJ12 at 5783.06 m; (b) hematite appears red under light reflected by the microscope, and the hematite alteration on the particle surfaces originates (R) from well YJ12 at 5783.06 m; (c) hematite developed at the contact between particles (R) from well Y12 at 5781.51 m; (d) porous cemented calcite outside of hematite (−) from well YJ1-1 at 5822.77 m; (e) hematite developed within quartz overgrowth (−) from well YJ1-1 at 5832.31 m; (f) hematite development at the contact between columnar autogenic microcrystals (−) from well YJ12 at 5786.96 m; (g) siderite fills the remaining pores and replaces the particles (−) from well YJ12 at 5786.96 m; (h) ferroan dolomite developed externally on hematite (−) from well YJ3-X17 at 5741.26 m; (i) hematite developed externally with zeolite and has been dissolved (−) from well YJ3-X17 at 5733.86 m. Qtz—quartz; Fsp—feldspar; Qo—quartz overgrowth; Fo—feldspar overgrowth; Hem—hematite; Cal—calcite; Fe–cal—ferroan calcite; Dol—ferroan dolomite; Anl—analcime.
Figure 7. Microscopic characteristics of the top red beds sandstone of the Jurassic in the Yongjin area of the Junggar Basin. (a) Development of abundant hematite films and iron-rich clay matrices (−) from well YJ12 at 5783.06 m; (b) hematite appears red under light reflected by the microscope, and the hematite alteration on the particle surfaces originates (R) from well YJ12 at 5783.06 m; (c) hematite developed at the contact between particles (R) from well Y12 at 5781.51 m; (d) porous cemented calcite outside of hematite (−) from well YJ1-1 at 5822.77 m; (e) hematite developed within quartz overgrowth (−) from well YJ1-1 at 5832.31 m; (f) hematite development at the contact between columnar autogenic microcrystals (−) from well YJ12 at 5786.96 m; (g) siderite fills the remaining pores and replaces the particles (−) from well YJ12 at 5786.96 m; (h) ferroan dolomite developed externally on hematite (−) from well YJ3-X17 at 5741.26 m; (i) hematite developed externally with zeolite and has been dissolved (−) from well YJ3-X17 at 5733.86 m. Qtz—quartz; Fsp—feldspar; Qo—quartz overgrowth; Fo—feldspar overgrowth; Hem—hematite; Cal—calcite; Fe–cal—ferroan calcite; Dol—ferroan dolomite; Anl—analcime.
Minerals 15 00347 g007
Figure 8. SEM and EDS characteristics of polished thin sections of Upper Jurassic red beds in the Yongjin area of the Junggar Basin. (a) The matrix, which is rich in illite/smectite mixed-layer clay, quartz, and feldspar fragments, develops between the particle contacts, with hematite fragments interspersed throughout (BSE) from well YJ15 at 5791.3 m. (b) Sheet-like chlorite developed between the particles, interspersed with a matrix of detrital fragments (BSE) from well YJ15 at 5791.3 m. (c) The pores are filled with a large amount of fibrous illite, amorphous illite/smectite mixed-layer clay, and sheet-like chlorite clay. Numerous fine hematite fragments grow within the micropores of the clay matrix, with an occasional occurrence of titanium oxide fragments (BSE) from well YJ12 at 5774.5 m. (d) Numerous hematite fragments develop within the pores (BSE) from well YJ15 at 5791.3 m. (e) Spectroscopic characteristics of illite/smectite mixed-layer clay (EDS). (f) Spectroscopic characteristics of hematite partially substituted by Ti (EDS). Ill—illite; Chl—chlorite; I/S—illite/smectite mixed-layer; Ti oxide—titanium dioxide.
Figure 8. SEM and EDS characteristics of polished thin sections of Upper Jurassic red beds in the Yongjin area of the Junggar Basin. (a) The matrix, which is rich in illite/smectite mixed-layer clay, quartz, and feldspar fragments, develops between the particle contacts, with hematite fragments interspersed throughout (BSE) from well YJ15 at 5791.3 m. (b) Sheet-like chlorite developed between the particles, interspersed with a matrix of detrital fragments (BSE) from well YJ15 at 5791.3 m. (c) The pores are filled with a large amount of fibrous illite, amorphous illite/smectite mixed-layer clay, and sheet-like chlorite clay. Numerous fine hematite fragments grow within the micropores of the clay matrix, with an occasional occurrence of titanium oxide fragments (BSE) from well YJ12 at 5774.5 m. (d) Numerous hematite fragments develop within the pores (BSE) from well YJ15 at 5791.3 m. (e) Spectroscopic characteristics of illite/smectite mixed-layer clay (EDS). (f) Spectroscopic characteristics of hematite partially substituted by Ti (EDS). Ill—illite; Chl—chlorite; I/S—illite/smectite mixed-layer; Ti oxide—titanium dioxide.
Minerals 15 00347 g008
Figure 9. SEM characteristics of Jurassic upper red bed clasts from the Yongjin area, Junggar Basin. (a) The particle surfaces are developed with a clastic matrix, amorphous mixed-layer illite/smectite clay, flaky chlorite clay, and fibrous illite clay (ETD) from well YJ1-1 at 5832.3 m. (b) High-brightness hematite clasts are mixed within the clay matrix in the pores (ETD) from well YJ1-1 at 5822.7 m.
Figure 9. SEM characteristics of Jurassic upper red bed clasts from the Yongjin area, Junggar Basin. (a) The particle surfaces are developed with a clastic matrix, amorphous mixed-layer illite/smectite clay, flaky chlorite clay, and fibrous illite clay (ETD) from well YJ1-1 at 5832.3 m. (b) High-brightness hematite clasts are mixed within the clay matrix in the pores (ETD) from well YJ1-1 at 5822.7 m.
Minerals 15 00347 g009
Figure 10. UCC-normalized plots of average major and trace element contents for Upper Jurassic red bed samples from the Yongjin area of the Junggar Basin. (a) UCC-normalized plot of major elements; (b) UCC-normalized plot of trace elements.
Figure 10. UCC-normalized plots of average major and trace element contents for Upper Jurassic red bed samples from the Yongjin area of the Junggar Basin. (a) UCC-normalized plot of major elements; (b) UCC-normalized plot of trace elements.
Minerals 15 00347 g010
Figure 11. A-CN-K ternary diagram of the Jurassic red beds in the Yongjin area of the Junggar Basin. A—molar (Al2O3), CN—molar (CaO*) + molar (Na2O), K—molar (K2O).
Figure 11. A-CN-K ternary diagram of the Jurassic red beds in the Yongjin area of the Junggar Basin. A—molar (Al2O3), CN—molar (CaO*) + molar (Na2O), K—molar (K2O).
Minerals 15 00347 g011
Figure 12. Oxidation-reduction environment discrimination diagrams of mudstone from the Jurassic top red beds in the Yongjin area, Junggar Basin: V/Cr (a), Ni/Co (b), and V/(V+Ni) (c).
Figure 12. Oxidation-reduction environment discrimination diagrams of mudstone from the Jurassic top red beds in the Yongjin area, Junggar Basin: V/Cr (a), Ni/Co (b), and V/(V+Ni) (c).
Minerals 15 00347 g012
Table 1. Composition of rock components in the Upper Jurassic red sandstone of the Yongjin area in the Junggar Basin.
Table 1. Composition of rock components in the Upper Jurassic red sandstone of the Yongjin area in the Junggar Basin.
WellDepthQtzFspRf
/mSCQ/%PCQ/%Kfs/%Plg/%Vrf/%Srf/%Mrf/%
Y25961.121.04.410.611.818.50.013.3
Y25963.921.36.17.07.026.90.416.5
Y25966.817.96.89.56.314.40.025.1
Y25967.516.84.49.87.128.00.012.4
Y2596819.24.511.36.325.30.08.9
Y25969.416.82.59.17.722.40.014.4
Y25974.217.91.811.94.317.60.516.6
Y66026.114.24.310.010.821.00.017.9
Y66026.814.85.56.58.311.50.016.0
YJ1-15832.917.43.08.212.912.60.219.9
YJ1-35978.016.74.18.15.623.80.018.8
YJ125774.520.94.511.17.714.51.915.1
YJ125779.417.22.48.87.324.70.812.8
YJ125781.523.32.911.87.914.90.512.5
YJ125783.116.15.57.25.015.10.017.6
YJ12578715.55.76.27.429.30.714.8
YJ145504.812.97.29.73.823.50.717.2
YJ3-X175733.921.53.810.66.923.70.911.8
YJ3-X175735.513.94.88.86.626.50.513.8
YJ3-X175741.222.32.912.37.119.80.010.2
YJ3035734.416.35.06.22.831.62.014.7
YJ3035752.420.09.36.93.724.40.213.0
Note: SCQ—Monocrystalline quartz; PCQ—Polycrystalline quartz; Kfs—Potassium Feldspar; Plg—Plagioclase; Srf—Sedimentary Rock Fragments; Vrf—Volcanic Rock Fragments; Mrf—Metamorphic Rock Fragments.
Table 2. Relative abundance of clay minerals in the Upper Jurassic red beds of the Yongjin area, Junggar Basin.
Table 2. Relative abundance of clay minerals in the Upper Jurassic red beds of the Yongjin area, Junggar Basin.
No.WellLithologyDepth (m)S (%)I/S (%)Ill (%)K (%)Chl (%)
1Y2sandstone5963.904328623
2Y2sandstone5967.504328623
3Y2sandstone5968.003633625
4Y6sandstone6026.803138724
5Y6sandstone6027.403330532
6Y6mudstone6013.404435516
7Y6mudstone6014.704539511
8Y6mudstone6015.703942712
9Y6mudstone6017.504043611
10Y6mudstone6020.704242511
Note: Ill: illite; Chl: chlorite; I/S: mixed-layer illite/smectite; S: montmorillonite; K: kaolinite.
Table 3. Relative abundance of major elements in Upper Jurassic red beds in the Yongjin area, Junggar Basin.
Table 3. Relative abundance of major elements in Upper Jurassic red beds in the Yongjin area, Junggar Basin.
NoWellDepth (m)LithologySiO2TiO2Al2O3TFe2O3MnOMgOCaONa2OK2OP2O5LOITotalCIACIW
wt.%wt.%wt.%wt.%wt.%wt.%wt.%wt.%wt.%wt.%wt.%wt.%
1Y25961.1sandstone76.540.2911.32.040.030.80.73.722.270.091.5599.3452.8 59.7
2Y25963.9sandstone780.3110.252.050.060.831.123.521.970.091.96100.1549.4 55.1
3Y25966.8sandstone66.470.4710.763.430.282.514.213.192.170.16.59100.1850.2 56.4
4Y25967.5sandstone78.140.3110.182.220.040.70.963.392.10.11.699.7350.5 56.9
5Y25969.4sandstone75.860.4411.163.10.040.870.753.522.230.121.699.6853.4 60.3
6Y25974.2sandstone72.110.3110.672.250.070.73.613.581.990.14.0399.4255.1 62.0
7YJ1-15832.9sandstone73.080.2910.152.720.130.83.832.932.340.094.14100.550.2 57.3
8YJ125779.4sandstone68.50.3910.192.820.292.314.073.142.170.16.37100.3449.6 56.0
9YJ145504.8sandstone76.910.3610.652.320.040.861.253.292.30.12.1100.1849.7 56.2
10YJ3-X175735.5sandstone73.370.3510.272.510.080.653.283.362.010.093.6899.6556.1 63.6
11YJ3035734.4sandstone70.980.4211.053.130.120.823.883.262.280.114.31100.3652.8 59.8
12YJ3035752.4sandstone66.450.7115.54.820.031.280.513.593.370.133.3799.7759.7 69.5
13YJ3035732mudstone56.520.7317.727.40.092.621.623.593.590.176.29100.3356.9 65.0
14YJ1-35966mudstone58.230.717.257.260.072.461.213.253.670.175.2399.559.2 68.6
Note: CIA and CIW represent the chemical index of alteration and the chemical weathering index, respectively. The specific calculation formulas are provided below.
Table 4. Relative abundance of trace elements in Upper Jurassic red beds in the Yongjin area, Junggar Basin.
Table 4. Relative abundance of trace elements in Upper Jurassic red beds in the Yongjin area, Junggar Basin.
NoWellDepth (m)LithologyBaCoCrNiSrVZn
(ppm)(ppm)(ppm)(ppm)(ppm)(ppm)(ppm)
1Y25961.1sandstone565.206.4026.609.40151.0034.9026.20
2Y25963.9sandstone860.108.00157.90119.50153.6032.6024.50
3Y25966.8sandstone372.2013.6063.8012.90189.3049.5038.60
4Y25967.5sandstone433.808.1037.0011.10146.3034.5027.20
5Y25969.4sandstone628.9011.0051.7015.90156.6056.5035.70
6Y25974.2sandstone1263.0014.0058.7014.20167.1041.0034.90
7YJ1-15832.9sandstone384.3013.40130.6046.10191.8041.6033.40
8YJ125779.4sandstone761.608.70114.9052.80180.0053.4037.30
9YJ145504.8sandstone506.406.8075.0092.40141.3044.1029.40
10YJ3-X175735.5sandstone1017.107.5046.007.40183.3047.8030.30
11YJ3035734.4sandstone396.7017.40297.00229.90210.9065.7039.80
12YJ3035752.4sandstone414.0012.7055.1026.60164.20103.3066.00
13YJ3035732.0mudstone274.4020.7064.0044.60190.80117.00107.00
14YJ1-35966.0mudstone502.3019.4089.0057.30170.70110.5098.10
*15YJ3015514.0mudstone1567.0012.62113.3037.62345.1069.3850.41
*16YJ3015582.0mudstone1975.0012.1675.7437.55383.8059.4569.57
*17YJ1-15821.7mudstone279.6017.68100.6049.37142.6095.36102.50
*18Y66024.6mudstone530.0017.66142.9043.90128.4097.8988.68
Note: *15 to *18 are for the collected data.
Table 6. Trace element results for the Upper Jurassic red bed mudstone in the Yongjin area.
Table 6. Trace element results for the Upper Jurassic red bed mudstone in the Yongjin area.
Trace Element RE WOE–WREOEMin–MaxAverageSources
V/(V+Ni)>0.840.6~0.84<0.60.61–0.810.69Hatch J R et al., 1992 [88]
V/Cr>4.252.0~4.25<2.00.61–2.981.32Jones B et al., 1994 [89]
Ni/Co>7.05.0~7.0<5.02.15–3.092.74Jones B et al., 1994 [89]
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

Guo, Y.; Li, C.; Zhang, L.; Lei, Y.; Hu, C.; Yu, L.; Zheng, Z.; Xu, B.; Liu, N.; Jia, Y.; et al. Genesis of the Upper Jurassic Continental Red Sandstones in the Yongjin Area of the Central Junggar Basin: Evidence from Petrology and Geochemistry. Minerals 2025, 15, 347. https://doi.org/10.3390/min15040347

AMA Style

Guo Y, Li C, Zhang L, Lei Y, Hu C, Yu L, Zheng Z, Xu B, Liu N, Jia Y, et al. Genesis of the Upper Jurassic Continental Red Sandstones in the Yongjin Area of the Central Junggar Basin: Evidence from Petrology and Geochemistry. Minerals. 2025; 15(4):347. https://doi.org/10.3390/min15040347

Chicago/Turabian Style

Guo, Yongming, Chao Li, Likuan Zhang, Yuhong Lei, Caizhi Hu, Lan Yu, Zongyuan Zheng, Bingbing Xu, Naigui Liu, Yuedi Jia, and et al. 2025. "Genesis of the Upper Jurassic Continental Red Sandstones in the Yongjin Area of the Central Junggar Basin: Evidence from Petrology and Geochemistry" Minerals 15, no. 4: 347. https://doi.org/10.3390/min15040347

APA Style

Guo, Y., Li, C., Zhang, L., Lei, Y., Hu, C., Yu, L., Zheng, Z., Xu, B., Liu, N., Jia, Y., & Li, Y. (2025). Genesis of the Upper Jurassic Continental Red Sandstones in the Yongjin Area of the Central Junggar Basin: Evidence from Petrology and Geochemistry. Minerals, 15(4), 347. https://doi.org/10.3390/min15040347

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