Next Article in Journal
Geochronology and Geochemistry of the Neoarchean Metabasalt in the Southern Liaoning Province, North China Craton: Implications on Regional Crustal Evolution
Previous Article in Journal
Iron-Rich Slag-Based Alkali-Activated Materials for Radioactive Waste Management: Characterization and Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Petrology and Geochemistry Features of the Middle Triassic Anisian Shale in Sichuan Basin, South China: Implications for Climatic and Environmental Condition Change

1
College of Geosciences, China University of Petroleum, Beijing 102249, China
2
Shale Oil Project Department, Shengli Oilfield Company, SINOPEC, Dongying 257000, China
3
State Key Laboratory of Petroleum Resources and Engineering, Beijing 102249, China
4
State Key Laboratory of Shale Oil and Gas Enrichment Mechanisms and Effective Development, Dongying 257000, China
5
Institute of Sedimentary Geology, Chengdu University of Technology, Chengdu 610059, China
6
Shengli Oilfield Company, SINOPEC, Dongying 257000, China
7
School of Geoscience and Technology, Southwest Petroleum University, Chengdu 610500, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(12), 1230; https://doi.org/10.3390/min15121230
Submission received: 19 September 2025 / Revised: 7 November 2025 / Accepted: 14 November 2025 / Published: 21 November 2025
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)

Abstract

The Middle Triassic Anisian stage records a critical phase of marine ecosystem recovery after the Permian-Triassic mass extinction. The recent discovery of the Lei3 shale member within the predominantly carbonate Leikoupo Formation in the Sichuan Basin presents a unique opportunity to investigate this recovery period. However, the depositional model, the primary controls on organic matter accumulation, and the hydrocarbon potential of this newly identified shale unit remain poorly understood. This study integrates petrological, elemental geochemical, and isotopic (δ13Corg, δ15Norg) data to reconstruct the depositional environment and to establish a genetic model for organic matter enrichment. Our results reveal a vertical evolution from organic-rich mixed shale at the base to clay shale and finally to calcareous shale at the top. This facies transition corresponded to decreasing chemical weathering (CIA values from ~80 to ~70), a shift from anoxic to dysoxic conditions, and declining paleoproductivity, driven by regional tectonic uplift and a climatic shift from warm-humid to sub-humid. We propose a novel ‘tectonically induced organic matter enrichment model’ where tectonic uplift, rather than climate or productivity alone, was the primary driver by modulating sea level, nutrient supply, and preservation conditions. The mixed shale, with the highest TOC content and type II1 kerogen, represents the most promising hydrocarbon-prone interval. This work not only clarifies the intricate tectonic–climate–productivity interplay during the Anisian but also provides critical insights for the exploration of shale gas in this emerging target.

1. Introduction

The biological mass extinction event, one of the most serious prior to the Phanerozoic, occurred during the Permian–Triassic periods. Restoration of the marine ecosystem was very slow, taking about 5 million years from the early Triassic [1]. However, the recovery of marine diversity and ecosystems suddenly accelerated from the Anisian Stage of the Middle Triassic [2]. Therefore, the Anisian Stage is the key to environmental change after the Permian-Triassic mass extinction.
According to conodont biostratigraphy [3,4] and magnetostratigraphy, in combination with zircon U/Pb ages, it is believed that the age of the Early-Middle Triassic World Line is 247.21 ± 0.1 Ma [5,6]. In recent years, Li et al. (2021) conducted U-Pb geochronology on the claystone at the base of the Leikoupo Formation, yielding a depositional age of 247.6 ± 1.1 Ma [7]. After that, the Anisian Stage began. The transition from the Early Triassic to the Middle Triassic was marked by a layer of altered acid tuff. This can be identified in most outcrops in the Sichuan Basin and its neighboring provinces (e.g., Guizhou) [8]. Shallow water carbonates were generally deposited during the Anisian period; however, a set of shale was suddenly deposited in the 3rd member of the Leikoupo Formation in the Middle Triassic (Abbreviated as Lei3 member). The reason for the deposition of this set of shale is still unclear.
Previously, Sun et al. (1989) proposed that the Yangtze Plate is not only the last marine sedimentary system but also the only potential reservoir that contains potential source rocks (Middle Triassic shales) [8]. Reservoirs in the Leikoupo Formation from the Anisian Period were discovered at the end of the last century, but the shale source proposed by scholars was not found at that time [9,10]. In recent years, there has only been a small amount of evidence that carbonate source rocks existed in this period [11]. It was not until 2021 that shale was discovered in the Lei3 member of the Leikoupo Formation [12], which not only verifies the statement that shale was deposited in the Mesozoic Leikoupo Formation in the Sichuan Basin but also provided new ideas on the source–reservoir relationship and for the next steps in the oil and gas exploration and development process for Leikoupo Formation.
While the recent identification of the Lei3 shale is a significant breakthrough [12], key aspects of this unit are still poorly constrained, creating critical knowledge gaps. First, a detailed depositional model that explains the lateral and vertical heterogeneity of the Lei3 shale is lacking. Second, the relative importance of tectonic activity, paleoclimate, ocean productivity, and redox conditions in controlling organic matter enrichment has not been systematically evaluated. Is the primary novelty the detailed reconstruction of the depositional model for this newly identified unit, the investigation of the tectonic–climate–productivity interplay, or its specific hydrocarbon potential? Third, the hydrocarbon generation potential of the different shale lithofacies (mixed, clay, calcareous) remains to be quantitatively assessed.
To address these gaps, this study employs a multi-proxy dataset—including petrology, major and trace elements, stable isotopes (δ13Corg, δ15Norg), and organic geochemistry—from the Lei3 member. The specific objectives are to: (1) reconstruct the paleoenvironmental conditions (redox, productivity, climate) during the deposition of the Lei3 shale; (2) establish a comprehensive depositional model that explains the observed lithofacies succession and organic matter accumulation; and (3) evaluate the hydrocarbon potential of the different shale lithofacies. By doing so, this work moves beyond a standard geochemical characterization to define the unique sedimentary and tectonic controls on this emerging shale unit, with implications for both understanding the Anisian recovery and for guiding future petroleum exploration in the Sichuan Basin.

2. Geological Setting and Stratigraphy

The Sichuan Basin is located in the northwestern part of the plate, on the eastern margin of the Paleo-Tethys Ocean [13,14,15,16]. Numerous paleomagnetic measurements limit the paleolatitude of the Sichuan Basin in the Middle Triassic to 14.7–23.3° [17,18,19,20,21].
The Sichuan Basin is a multi-cycle superimposed basin, which evolved from a passive continental margin in the Pre–Middle Triassic to a late foreland basin [22,23,24]. It contains the oldest record of marine sediments in China [25,26,27].
In the Middle Triassic, the South China Block and the North China Block collided, and the Mianlüe Ocean gradually closed from east to west [28,29,30,31]. A strong orogenic movement began in southern China [32,33,34,35], which is characterised by NW-trending folds, thrust faults, and post-orogenic granitic intrusions. The formation of the Xuefengshan tectonic belt was caused by the eastward intracontinental subduction of the South China Plate [32,33,36], which was also squeezed to the west to form the Luzhou paleo-uplift [37]. These tectonic events greatly changed the sedimentary environment of the Sichuan Basin [22,23,38].
There was a large-scale transgression in the Lei3 sedimentary period, which was larger than the transgression at the end of the Early Triassic (Olenekian) [10]. During this period of transgression, seawater mainly entered the Upper Yangtze River Basin through the Kaiyuan Strait between West Sichuan Central Yunnan oldland and North Vietnam oldland, then invaded the Sichuan Oceanic Basin from the southwest by crossing the Qiannan barrier reef. The seawater in the secondary direction of transgression passed through the channel in the Longmen Mountain island chain via the Songpan-Garze trough or overflowed the underwater seawall into the Sichuan Oceanic Basin (Figure 1) [18,38]. Subsequently, the Ladinia Regression occurred in the later stage of the Middle Triassic, and the sedimentary facies converted into continental facies [39].

3. Samples and Methods

Thirty-one samples were taken from HC125 well (106°28′35″ E, 30°26′47″ N) of the Lei3 member of the upper Yangtze Platform in southern China. All samples were identified under a microscope (Nikon LV100POL produced by Nikon in the Tokyo, Japan), including by alizarin red and potassium ferricyanide staining and casting thin section analysis. Each sample was plated with gold powders to facilitate electron microscopy. The gold-plated samples were scanned by Quanta 250 FEG scanning electron microscope (SEM) produced by Thermo Fisher in Waltham, MA, USA at 5 kV acceleration voltage. Samples for geochemical analysis were carefully selected under a binocular microscope and were analysed for major and trace elements, total organic carbon (TOC), X-Ray Diffraction (XRD) and δ13Corg isotopes. Only 18 samples were suitable for δ15Norg isotope analysis.

3.1. XRD

To evaluate their mineralogy, thirty-one samples were ground using a mortar and pestle into powder (200 mesh). The samples were analyzed with X-rays using Panalytical X’Pert PRO X instrument produced by Panalytical in the Almelo, The Netherlands. according to the requirements of Chinese National Standard SY/T5163-2010 [41], and step-scanned at a 5–60° 2θ interval.

3.2. Major and Trace Elemental Analysis

Major and trace elements of 31 powder samples (about 200 mesh) from the Leikoupo Formation were analyzed in the Key Laboratory of Crustal Dynamics, Institute of Crustal Dynamics, China Earthquake Administration (CEA), Beijing. According to the glass melting method, X-ray fluorescence (XRF produced by Bruker in Bill Ricard, MA, USA; AXIOS) was used to determine the content of main elements. Prior to the experiments, powdered shale samples were calcined in a high-temperature furnace (700 °C) to completely remove organic matter. Trace elements were measured using Inductively Coupled Plasma Mass Spectrometry (ICP-MS produced by Agilent in Santa Clara, CA, USA; X-series II). Before the experiments, shale samples were placed into a polytetrafluoroethylene (PTFE) vessel with a mixed solution of HClO4, HF and HNO3 to dissolve powdered samples.
Uncertainty reports for individual major elemental oxide analyses were better than 5% (2σ). The major elemental oxides of all the samples were determined according to Chinese National Standard SamplesGSR-3 and GSR-11 (referring to basalts and rhyolites, respectively). The accuracy of trace and rare earth elemental analysis was monitored by Chinese National Standard Samples GSR-1, GSR-2, GSR-10 and GSR-16, representing granites, andesites, gabbros and dolerites, respectively [42].

3.3. TOC

The preparation of samples for TOC analysis followed the steps provided by Chinese National Standard GB/T 19145-2003 [43]. Sample (10 g) was crushed to 80 mesh using a pestle and agate mortar. The sample was digested in HCl (5% concentration) for 2 h to remove carbonate minerals, and the acid was removed by washing with distilled water. The sample was then dried in an oven at 60 °C for 48 h. A LECO CS-230 carbon and sulfur analyzer produced by Leco in St. Joseph, MI, USA was used to measure the TOC of the sample, and the precision was within 0.5%.

3.4. Nitrogen and Organic Carbon Isotopes

To analyze the nitrogen and carbon isotope ratios, fresh parts from Anisian (Middle Triassic) shale were cut to eliminate the influence of the weathered surface of the sample on the experimental results. Rock powder was then prepared from suitable fragments of 1–2 cm in diameter using a polycarbonate tube and agate mortar. The powder (1–2 g) was first digested overnight at 70 °C with HCl (~10 M, 12–18 mL) to eliminate any carbonate minerals. When the reaction was complete, the residue was separated from the acid and rinsed with ethanol. The samples were then dried overnight at 80 °C and stored in glass tubes. Total nitrogen (TN) and total organic carbon (TOC) concentrations and isotope ratios were determined using an elemental analyzer–isotope ratio monitoring mass spectrometer (EA-IRMS; Thermo Finnigan DELTA plus Advantage mass spectrometer combined with EA 1112 series FLASH elemental analyzer, produced by Thermo Fisher in Waltham, MA, USA). The analysis process was as described by Kikumoto et al. [44]. The analytical reproducibility of TN and TOC concentrations was evaluated by replicate analysis of internal standard materials and samples; relative standard deviations were better than ±1.8 and 1.6% when N and C content in rock samples exceeded 1 ppm, respectively. Nitrogen and carbon isotopic compositions (δ15N and δ13C) are reported in ‰ relative to atmospheric N2 and Vienna Peedee Belemnite (V-PDB) standards, respectively, based on (Rsample/Rstandard − 1) × 1000, where Rsample and Rstandard are the isotopic compositions of rock samples and reference materials, respectively. The δ15N and δ13C isotopic compositions of the rock samples and reference materials are denoted 15N/14N and 13C/12C, respectively. The analytical reproducibility of the δ15N and δ13C values determined from the replicate analyses of the internal standard materials was better than ±0.1‰ and ±0.2‰, respectively.

3.5. Proxy Calculations

Before using the trace element substitute index to indicate the deposition environment, Ti levels are usually used to correct the trace elements [45,46]. This enables calculation of the non-detrital or excess elements and the corresponding excess values, which eliminates the interference of terrigenous materials [47,48].
The equation is as follows:
Elementxs = Elementtotal − Titotal × (Element/Ti) PAAS
Elementxs and Elementtotal correspond to the content of elements from non-detrital sources and the total elemental content, respectively. Element/Ti is the content ratio of the element to Ti in Post-Archean Australian Shale (PAAS) [49,50].
To determine the degree of restriction and the redox conditions of the water mass, the element enrichment factor was calculated. First, the concentration of Mo and U needed to be transformed into the form EFX (enrichment factor of X element, X refers to some element), which was calculated according to the following equation [48]:
EFX = (X/Al)sample/(X/Al)PASS
Here, X refers to some element. X and Al are the weight concentrations of X and Al elements, respectively.
Index of compositional variability (ICV) was used to assess the extent to which sediment recycling altered the composition of fine-grained clastic rocks [51]. The IVC was calculated according to the following equation:
ICV = (Fe2O3 + K2O + Na2O + CaO + MgO + MnO + TiO2)/Al2O3
The measured chemical index of alteration (CIA) values were used to estimate the paleoclimate and degree of weathering by excluding the sedimentary alterations after circulation and sorting [52]. The CIA formula was calculated as follows:
CIA = mole [Al2O3/(Al2O3 + CaO* + Na2O + K2O)] × 100
Here, the measured values are expressed as molar proportions, and CaO* represents the CaO in silicate minerals. CaO* can be calculated and corrected as follows:
CaO* = mole CaO − mole P2O5 × 10/3
This method for correcting the CaO* value was proposed by McLennan [53]. If the corrected number of moles is less than that of Na2O, the CaO value is adopted as the CaO* value. Otherwise, the CaO* was assumed to be equivalent to the moles of Na2O.

4. Results

4.1. Lithology and Mineralogy

At present, using siliceous minerals (quartz + feldspar), carbonate minerals (calcite + dolomite) and clay minerals as the three terminal elements, the shale facies of the Lei3 Member can be divided into three types: mixed shale, clay shale and calcareous shale (Figure 2).
There were few organisms in the shale, and lamellation could be seen. Through microscopic observation and XRD analysis, it was found that the shale deposited in the Lei3 Member contained more carbonate, clay and quartz minerals than other minerals. The carbonate minerals were mainly calcite (22.3–74.7 wt%; mean: 47.3 wt%), the quartz content was 9.2–31.2 wt% (mean: 19.3 wt%) and the clay mineral content was 14.2–42.3 wt% (mean: 33.4 wt%).
The calcareous shale was mainly deposited in the upper part of the Lei3 member. It was light-gray to gray, mostly homogeneous and massive, lacking bedding (Figure 3A,D). The content of quartz and clay minerals was low, with an average of 15.4 and 16.8 wt%, respectively, while the content of carbonate was the highest, reaching 67.7 wt% (Figure 4). The clay shale was mainly deposited in the middle of the Lei3 member. It was gray-black, had low quartz content and the highest clay mineral content (Figure 3B,E), with mean values of 18.8 and 51.8 wt%, respectively. The carbonate minerals were mainly calcite (mean: 29.4 wt%). The mixed shale was mainly deposited at the bottom of the Lei3 member. It was gray-black and contained a mean quartz and clay mineral (Figure 3C,F) content of 24.2 and 41.2 wt%, respectively. The carbonate content was relatively low (mean: 34.6 wt%).
Overall, the quartz content showed a decreasing upward trend, the clay mineral content increased and then decreased, the carbonate minerals showed an increasing trend, and the TOC showed a decreasing trend (Figure 4).

4.2. Isotopic and Elemental Geochemistry

The distribution range of δ13Corg values was −27.07‰ to 31.07‰ (mean: −28.92‰). Upward, a weak increasing trend was exhibited. The δ15Norg isotope values ranged from 2.55‰ to 4.85‰ (mean: 3.79‰) and showed a similar trend to δ13Corg, with a slight increase in the upward direction (Figure 5).
The V/Cr, Ni/Co, V/(V + Ni), U/Th, MoXS, UXS and SrXS/BaXS values were used as proxies for water mass redox conditions [55,56,57,58,59]. Upward, the V/Cr, Ni/Co, V/(V + Ni), U/Th, UXS and MoXS profiles exhibited similar decreasing trends (Figure 6).
The V/Cr value distribution range was 2.04–7.56 (mean: 4.21), with the distribution ranges of mixed shale, clay shale and calcareous shale being 4.6–7.56, 3.8–4.51, and 2.04–3.11, respectively (mean values of 6.18, 4.08, and 2.62, respectively). The distribution range of Ni/Co values was 5.06–10.73 (mean: 6.84) with the ranges of mixed shale, clay shale and calcareous shale being 7.05–10.73, 6.08–7.18, and 5.06–6.54 (with mean values of 8.24, 6.59, and 5.81, respectively). The V/(V + Ni) values ranged from 0.66 to 0.83 (mean: 0.74), with the ranges of mixed shale, clay shale, and calcareous shale being 0.78–0.83, 0.74–0.77, and 0.66–0.73, respectively (with mean values of 0.8, 0.75, and 0.69, respectively). The distribution range of U/Th values was 0.79–3.39 (mean: 1.33) with the ranges of mixed shale, clay shale and calcareous shale being 0.79–1.02, 1.02–1.28, and 1.36–3.39 (with mean values of 0.95, 1.11, and 1.94, respectively). The range of MoXS values was from 8.38 to 54.78 (mean: 29.25). The distribution range of mixed shale, clay shale and calcareous shale was 32.12–54.78, 26.7–38.37, and 8.38–29.73, respectively (with mean values of 40.93, 33.13, and 17.29, respectively). The UXS values ranged from 3.84 to 17.74 (mean: 10.66). The distribution range of mixed shale, clay shale and calcareous shale was 9.45–17.74, 9.02–12.77, and 3.84–11.27 (with mean values of 14.09, 11.16, and 7.49, respectively). The range of Sr/Ba values was from 0.54 to 1.52 (mean: 0.88).

4.3. Terrigenous Fluxes

The concentrations of Al and Ti can be used as representative of terrigenous debris input. The distribution range of Al values was 4.07%–8.02% (mean: 5.92%) with the ranges of mixed shale, clay shale and calcareous shale being 4.07%–5.78%, 4.36%–7.16%, and 5.61%–8.02% (with mean values of 4.86%, 6.01%, and 6.76%, respectively). The Ti values ranged from 0.19% to 0.52% (mean: 0.34%), with the ranges of mixed shale, clay shale, and calcareous shale being 0.25%–0.31%, 0.29%–0.35%, and 0.38%–0.52%, respectively (with mean values of 0.25%, 0.34%, and 0.43%, respectively).

4.4. Organic Geochemistry

4.4.1. Organic Matter Abundance and Maturity

Total Organic Carbon (TOC) in Lei3 samples from HC125 ranges from 0.11 to 4.01, with an average of 1.32. Stratigraphically, shale types ranked by TOC from highest to lowest are mixed shale > clay shale > calcareous shale > carbonate rock, with TOC values of 2.77%, 1.54%, 0.76%, respectively. TOC shows a decreasing trend vertically. Carbonate rock TOC is low and not discussed in this paper.
Lei3 shale samples in HC125 exhibit high Tmax values: 424–592 °C (average: 528.25 °C). Tmax shows a slight increase with depth. All shale samples display very low S2 values (0.06–0.60), averaging 0.21, and a Production Index (PI) ranging from 0.14 to 0.35, with an average of 0.20.

4.4.2. Kerogen Components

Through vitrinite reflectance microscopy, microcharacteristics and component contents of kerogen within Lei3 shale have been observed. Based on these microscopic differences, vitrinite, inertinite, exinite, and huminite can be identified in Lei3 samples. Across all samples, huminite predominates with amorphous humic matter (52%–68%). Following this is the exinite (29%–41%). The kerogen type index (TI) is calculated using the formula (TI = (huminite × 100 + exinite × 50 − vitrinite × 75 − inertinite × 100)/100), which is used to classify kerogen types based on this index.
Huminite is characterized by amorphous humic matter, appearing orange-brown to brown-orange under microscopic examination, dispersed without specific shapes. Exinite kerogen predominantly consists of amorphous humic matter, retaining remnants of original plant morphology (Figure 3G–I).

4.4.3. Kerogen Carbon Isotopes

In HC125, Lei3-2 shale samples exhibit kerogen isotopic values ranging from −27.7‰ to −30.8‰, with an average of −28.79‰. The isotopic range for kerogen in shale samples spans from −28.1‰ to −30.8‰, averaging −28.87‰.

4.4.4. Biomarkers

Normal Alkanes
The UCM peak on the saturated hydrocarbon chromatogram was flat, indicating weak degradation. Therefore, biomarker compounds are highly reliable in their indication of the environment. HC125 samples show dominance of short-chain normal alkanes (Figure 7), with maximum abundance between carbon numbers C17–C20. Using the Carbon Preference Index (CPI) formula: 22(nC23 + nC25 + nC27 + nC29)/(nC22 + 2(nC24 + nC26 + nC28) + nC30) and the Odd-Even Preference (OEP) formula: (nC21 + 6nC23 + nC25)/(4nC22 + 4nC24), average CPI and OEP values are 0.94 and 0.99, respectively, close to the equilibrium value of 1. The ratio of C21+22/C28+29 ranges from 1.33 to 4.45, averaging 2.29. Specifically, clay shale averages 2.55, calcareous shale averages 1.71, and mixed shale averages 3.32.
Sesquiterpanes
In HC125 samples, the proportion of Pr and Ph in sesquiterpanes is lower compared to adjacent normal alkanes, with Pr content being less than Ph. The Pr/Ph ratio in shale samples ranges from 0.52 to 1.12, averaging 0.82 (Figure 7). Pr/nC17 and Ph/nC18 are used to create cross-plots to reflect the source of organic matter (Figure 8).
Terpanes
In the ion mass spectrum (m/z191) (Figure 9A,C), similar features are observed in all samples: high content of C30 hopanes, followed by tricyclic terpanes. In HC125, the distribution of Ts/(Ts + Tm) values ranges from 0.45 to 0.58, with an average of 0.51. Additionally, the samples contain elevated levels of gammacerane.
Steroids
In the ion mass spectrum (m/z217) (Figure 9B,D), significant amounts of pregnane compounds are present in the HC125 shale samples, particularly C215α-(H)-pregnane. However, there are variations in the content of cholestane compounds, with cholestane content exceeding that of pregnane compounds. C27 and C29 regular steranes are associated with algae and plants, respectively [62]. All samples contain C27–C28–C29 regular steranes and exhibit an “L-shaped” carbon number distribution pattern.

5. Discussion

5.1. Organic Geochemical Characteristics

5.1.1. Hydrocarbon Generation Potential

The total organic carbon (TOC) content in HC125 samples ranges from 0.11 to 4.01%, with an average of 1.32%. In the study by Huo et al. (2022), the TOC content of the Anisian shale of the Leikoupo Formation from wells CT1 and JY1 ranges from 0.13 to 1.78, averaging 0.81 [42]. HC125, CT1, and JY1 samples exhibit similar TOC distribution characteristics, with TOC values ranking as follows: mixed shale > clay shale > calcareous shale > carbonate rock, and showing a decreasing trend vertically. The C29αα20S/(20S + 20R) and C29ββ/(ββ + αα) ratios are effective indicators for maturity assessment [63,64]. In HC125 samples, the C29αα20S/(20S + 20R) ranges from 0.40 to 0.41, averaging 0.405; CT1 ranges from 0.47 to 0.58; and JY1 ranges from 0.47 to 0.54 [42], all indicating high maturity to overmaturity of Anisian shale. Additionally, maturity can be identified using the Ts/(Ts + Tm) ratio, typically indicating high maturity when Ts/(Ts + Tm) ranges from 0.4 to 0.6 [65]. In HC125 samples, the Ts/(Ts + Tm) ratio ranges from 0.45 to 0.58, averaging 0.51; CT1 ranges from 0.45 to 0.58 (mean: 0.51); and JY1 ranges from 0.43 to 0.56 (mean: 0.5) (Huo et al., 2022) [42], indicating high maturity of Anisian shale in the Sichuan Basin. The relationship between Tmax and PI values in pyrolysis results also reflects maturity characteristics, indicating that the samples are mainly in the wet gas and dry gas/condensate oil stages [66,67]. In the study by Huo et al. (2022) [42], Tmax values for CT1 samples range from 452 to 576 °C (mean: 498 °C) and for JY1 samples from 442 to 528 °C (mean: 494.5 °C), both lower than HC125 samples (424–592 °C, mean: 528.25 °C). This suggests higher maturity of HC125 potentially due to its closer proximity to the sedimentation and hydrocarbon generation center. The type of organic matter is determined by the TI: Type I when TI > 80; Type II1 when TI is between 40 and 80; Type II2 when TI ranges from 0 to 40; and Type III when TI is less than 0 (referenced). Based on the results of vitrinite reflectance (Figure 3G–I), calculated TI indices (66.5–80.75), and the distribution of δ13C values of kerogens in the samples (−29‰ to −26‰), the predominant kerogen types in the Anisian sample are Type II1, with some Type I. Compared to Type III kerogens, Type II1 and Type I kerogens are richer in hydrogen and lipid components, indicating a significant hydrocarbon generation potential [68]. It is important to note that high thermal maturity alters the optical properties of macerals, making visual kerogen typing and the calculated Type Index (TI) less reliable. Therefore, the results of cheese root microscopy should be used cautiously as auxiliary evidence.
Despite the high maturity characteristics of the Anisian shale, parameters such as S1, S2, HI, and OI from pyrolysis lose their ability to assess original hydrocarbon generation potential. Indeed, Tmax results from rocks hosting black shale gas reservoirs in other basins around the world indicate highly evolved shale gas production (Table 1). High maturity hinders the identification of original hydrocarbon generation potential but indicates a more thorough and complete process of oil and gas generation. Under conditions of high maturity and with the expulsion of most hydrocarbon products, the original organic abundance in Anisian shale should theoretically be much higher than modern residual organic matter. Therefore, the hydrocarbon generation potential of Anisian shale is considerable. Furthermore, based on residual TOC comparisons, mixed shale exhibits greater hydrocarbon generation potential than clay shale, while clay shale surpasses calcareous shale and carbonate rocks in this regard.

5.1.2. Organic Matter Sources

The distribution characteristics of n-alkanes in saturated hydrocarbon chromatograms reveal a predominance of short-chain n-alkanes (peak abundance between carbon numbers C17–C20) in the samples. These short-chain n-alkanes are primarily derived from lower marine planktonic organisms. Detection of n-alkanes in marine organic matter shows dominance of (C21 + C22) hydrocarbons, while terrestrial plant organic matter predominantly contains (C28 + C29) n-alkanes [73]. In asphaltene samples from shale, the average C21+22/C28+29 ratio is 2.29, indicating dominance of C21 and C22, suggesting a marine origin. The C21+22/C28+29 ratio decreases vertically from mixed shale (3.32) to clay shale (2.55) to calcareous shale (1.71), reflecting a relative decrease in marine organic matter sources, similar to the TOC variation trend, possibly related to regression phases and indirectly proving the marine origin of the organic matter.
Isoprenoid alkanes such as pristane (Pr) and phytane (Ph), derived from chlorophyll phytol side chains in photosynthetic organisms, are sensitive indicators of redox conditions. The Pr/Ph ratio, with ranges indicating anoxic, suboxic, and oxic conditions [74], in HC125 samples (average 0.82) indicates a reducing environment for the Anisian shale deposition, corroborated by Pr/nC17 and Ph/nC18 diagrams (Figure 8). Through characteristics of n-alkanes and isoprenoid alkanes, the depositional environment is conclusively determined as a reducing marine water environment.
Hopanes like gammacerane are typically associated with cyanobacteria (blue-green algae) and other prokaryotes [75]. And the tricyclic terpanes primarily occur in marine source rocks as products of original bacteria and prokaryotes post-diagenesis, largely independent of terrestrial sources. This suggests that primary producers in the depositional water bodies were likely marine plankton—cyanobacteria. Gamacerane, a reduction product of tetramethylated triterpenoids in foraminifera, correlates closely with their living conditions, typically indicating vertical stratification in saline and anoxic conditions [76]. The GI index (GI = gammacerane/αβC30 hopane) ranges from 0.22 to 0.28 in HC125, averaging 0.25, reflecting the degree of water column stratification.
C27 and C29 regular steranes are associated respectively with algae and plants [62]. Furthermore, the significant presence of C27αααR > C29αααR > C28αααR in HC125 samples exhibits an asymmetric anti-“L-type” distribution pattern. This confirms the organic matter in the shale originates from marine algae. Similar characteristics are observed in CT1 and JY1 samples [42].
In conclusion, the deposition and organic matter enrichment processes of Anisian shale occurred in a reducing marine environment, exhibiting vertical water column stratification in salinity and redox conditions.

5.2. Depositional Environment

The above discussion indicates that the tectonic movement had a significant controlling effect on black shale sedimentation in the study area, but the influence of the sedimentary environment on the shale could not be ignored.

5.2.1. Paleo-Weathering and Recycling Processes

Chemical weathering is closely related to paleoclimate and is an effective method for tracking paleoclimate changes [77]. Various geochemical indicators have been used for paleoclimate reconstruction, such as the chemical index of alteration (CIA) to infer the intensity of weathering [78,79,80]. Before calculating CIA values, it is necessary to exclude sedimentary recycling and post-sedimentary diagenesis alterations [51], so that the measured CIA value can be used to estimate the paleoclimate and degree of weathering.
(1)
Sediment recycling and sorting
Polycyclic reworking processes tend to accumulate stable minerals like quartz [53]. The coarse-grained sediments are rich in minerals such as quartz and zircon, and the fine sediments are rich in clay [81,82,83,84,85,86]. Analysis of the shale deposited in the Lei3 Member revealed that the influence of sedimentary differentiation could be excluded and that clay minerals were relatively enriched.
Generally, for the study of older strata, the index of compositional variability (ICV) can be used to estimate the degree of change in the composition of fine-grained clastic rocks by sediment recycling [51]. Mature sediments from the stable continental crust display lower ICV values (<1.0), whereas immature sediments from arc-related volcanic and plutonic rocks exhibit higher values [51].
Most of the ICV values of the Lei3 shale samples were greater than 1 (mean: 1.15). This meant that the mineral maturity was low, so Lei3 shale samples were weakly affected by recirculation.
(2)
Post-depositional diagenetic alteration
The vertical decrease in CIA values (Figure 5) may have been due to tectonic movements and changes in the sedimentary environment, paleoclimate, and the composition of source rocks. The CIA value is sensitive to the composition of source rocks, especially K metasomatism [78]. Therefore, the effect of K metasomatism should be excluded when calculating CIA values [87].
In the A–CN–K diagram, the trend of weathering was simulated by plotting the values of the samples on a line parallel to the A–CN axis (Figure 10). The weathering trend from the source rocks was parallel to the A–CN axis, and the CIA values were relatively concentrated and not tilted toward the K apex. Therefore, the involvement of K was not the main factor leading to the decrease in CIA values [78]. This further supported the decrease in CIA being affected by tectonic movements and paleoclimate changes. The previous analysis showed that regional tectonic uplift in the study area was related to physical weathering.
In the deposition period of mixed shale, the paleoclimate was relatively warm and humid with the CIA values of 77.91–83.22 (mean: 80.59). Vertically, the CIA values were relatively stable and gradually increased upwards in the Lei3 member. The range of CIA values of clay shale during the deposition period was 75.25–80.05 (mean: 77.51) changed to 66.81–76.05 (mean: 70.89) during the deposition period of calcareous shale.
Based on analysis of the A–CN–K diagram, it was concluded that the paleoclimate changed from warm and humid to semi-humid during the deposition period of shale in the Lei3 member. Under humid conditions, values of Fe, Mn, Cr, V, Ni, Co, and other elements in sedimentary rocks are high. However, under arid conditions, values of Ca, Mg, K, Na, Sr, and Ba are high because these elements separate out and catalyze the deposition of various salts at the bottom of the water body due to the evaporation of water and the enhanced alkalinity. The climate index was calculated from the following ratio:
C = ∑ (Fe + Mn + Cr + V + Ni + Co)/(Ca + Mg + K + Na + Sr + Ba)
According to the classification criteria, values changing from 0 to 0.2 represents an arid climate; from 0.2 to 0.4 represent a semi-arid climate; from 0.4 to 0.6 represent a semi-arid to semi-humid climate; from 0.6 to 0.8 represent a semi-humid climate; and more than 0.8 represents a humid climate [77,89,90].
In this study, the range of climate index C values of the sample is 0.63–1.01 (mean: 0.81), indicating a humid climate. The C values decreased from the mixed shale at the bottom of Lei 3 member to the calcareous shale at the top, indicating that the climate changed from warm humid to a semi-humid climate (Figure 11).
According to this analysis, the paleoclimate during the deposition of the Lei3 member changed from warm and humid to semi-humid. The longitudinal change in climate to semi-arid was not conducive to the reproduction of aquatic microorganisms, so their productivity gradually decreased. As a result, the TOC content corresponded to a slight downward trend as time progressed.

5.2.2. Marine Redox Conditions

The marine redox condition is a key factor for the sedimentary environment. It even affected the recovery of ecosystems after the P-T mass extinction.
(1)
Organic carbon isotope δ13Corg
δ13Corg values mainly reflect photosynthesis, carbon assimilation, and the isotopic composition of carbon sources [91]. The δ13Corg values of sedimentary organic matter are closely related to the source and sedimentary environment of the original organic matter. However, as thermal maturation does not markedly change the carbon isotope composition, the change in δ13Corg in the sedimentary organic matter reflects the change in δ13Corg of the original organic matter—in most cases.
Freudenthal et al. (2000) and Lehmann et al. (2002) believe that the widespread distribution of a hypoxic environment has a significant effect on the original isotope balance of δ13Corg: when environmental anoxic conditions are dominant, δ13Corg tends to be lighter [4,92]. The organic matter input of the Lei3 shale was dominated by algae and plankton. For marine algae, the δ13Corg value is usually −20‰ to −22‰ [93]; the δ13Corg value of the Lei3 shale was generally lower than −28‰ (Figure 5). The lighter δ 13Corg value is indicative of a generally hypoxic environment [94]. In our study, the co-occurrence of the most negative δ13Corg with the highest values of productivity proxies (Pxs, Baxs) and the strongest anoxia indicators (e.g., high V/Cr, MoEF) suggests that both processes were likely operating in concert. The high productivity fueled organic flux to the seafloor, promoting anoxia, which in turn enhanced the preservation of this 12C-enriched organic matter. This synergistic effect makes it difficult to isolate the primary driver, but the multi-proxy evidence strongly supports a high-productivity anoxic basin model.
(2)
δ15Norg isotope
According to previous studies, changes in marine redox and productivity and regional climate conditions are closely related to the marine N-cycle balance [95]. There are four biogeochemical processes of N cycling in the ocean: biological nitrogen fixation and nitrification mainly occur near the ocean surface, while biological denitrification and anaerobic oxidation of ammonia occur in deep water [95,96,97,98]. During the Lei3 shale deposition period, the sea level rose rapidly, and the extensive transgression resulted in a stratified and anoxic marine environment. At the same time, denitrification and biological nitrogen fixation were strengthened under high productivity, which led to the enrichment of δ14N in biological organic matter and its negative deviation in sediments. During the middle and later periods, due to the long-term anoxic reduction conditions, the remaining nitrate in the seawater inorganic N pool was relatively enriched by 15N. Tectonic uplift led to a decline in sea level, promoted water-gas exchange, and changed the seawater redox conditions. On the one hand, these factors limited the process of denitrification; on the other hand, organisms absorbed the remaining 15N-rich nitrate ions in seawater, resulting in the positive deviation of δ15N isotopes (Figure 5).
(3)
Elemental geochemistry
To define the redox environment, the three-part classification of oxic, dysoxic, and anoxic was first proposed by Rhodes and Morse [99]. For reconstruction of the paleoenvironment, selected trace elements that have few sources and relative stability after deposition are used as the geochemical indexes of paleo-oxygen facies [48]. Vxs, Crxs, Nixs, Coxs, Uxs, and Thxs were chosen as the trace element substitution indexes for determining the paleo-redox environment.
All these elements dissolve in seawater as stable, high-valence ions [47]. In a weak reduction environment, these high-valence ions are reduced to lower valences and deposited. Under strongly reducing conditions, these high-valence ions are reduced to a lower-valence state and deposited as oxides or hydroxides [47]. Therefore, the redox conditions can be determined.
The scatter diagram shows that there was a positive correlation between V/Cr, Ni/Co, V/(V + Ni), U/Th and TOC (Figure 6) and that from mixed shale to clay shale to calcareous shale, the degree of hypoxia gradually decreased. That is, mixed shale was mainly deposited in an anoxic environment, clay shale was mostly deposited in an anoxic environment, with some in a dysoxic environment, and calcareous shale was mainly deposited in a dysoxic environment (Figure 12).
The oxygen content in seawater varies over time and depends to a large extent on the geographic environment and prevailing climate [100,101]. It has been determined that the tectonic movement gradually increased during the sedimentary period of the Lei3 shale. With the gradual decrease in sea level, the degree of hypoxia did not decrease significantly. A simple sea-level fall model would predict improved ventilation and more oxic conditions. However, we propose a more nuanced mechanism linked to the interplay between tectonic uplift and basin geometry. On the one hand, regional tectonic uplift intensified physical weathering, increasing the flux of terrigenous nutrients (e.g., phosphorus) into the basin. This could have sustained high surface productivity, consuming oxygen in the water column through the degradation of organic matter [100,102]. On the other hand, the same tectonic event (e.g., the formation of the Luzhou paleo-uplift) likely altered basin topography, increasing its restriction. This is supported by the increasing Mo/TOC and MoEF-UEF trends in the clay and calcareous shales. A more restricted basin, even if shallower, would have impeded the renewal of oxygenated open-ocean water, trapping organic matter and maintaining stratification and bottom-water anoxia [42]. Although the mechanisms differ, they act in concert and ultimately modulate the redox conditions of the water body through water column stratification.
Figure 12. Cross plots of redox proxies of Lei3 shale in HC125. (A) Vxs/Crxs vs. Nixs/Coxs, (B) Nixs/Coxs vs. Uxs/Thxs, (C) Nixs/Coxs vs. Vxs/(Vxs + Nixs). The empirical values of redox proxies are derived from [48,103,104,105].
Figure 12. Cross plots of redox proxies of Lei3 shale in HC125. (A) Vxs/Crxs vs. Nixs/Coxs, (B) Nixs/Coxs vs. Uxs/Thxs, (C) Nixs/Coxs vs. Vxs/(Vxs + Nixs). The empirical values of redox proxies are derived from [48,103,104,105].
Minerals 15 01230 g012

5.2.3. Paleoproductivity

(1)
Organic carbon isotope (δ13Corg)
Higher ocean productivity leads to higher organic flux and more oxygen consumption, which is beneficial to maintain or even prolong anoxic conditions in the water column [48,56]. δ 13Corg values are closely related to the source of organic matter. The relationship is as follows: plankton preferentially absorbs atmospheric 12CO2 as the carbon source for photosynthesis [106,107]. As a result, the generated organic matter contains more 12C, resulting in lighter δ13Corg values in mudstone. Therefore, more negative δ 13Corg values reflect higher productivity and the presence of more organic matter [108]. As the δ13Corg values were lowest in the lower part of the Lei3 member and gradually increased upward, this indicated a gradual decrease in productivity.
The δ13Corg values in the study area were relatively low, which indicated algae to be the source of organic matter. The increasing δ13Corg trend in the upward direction may have been caused by the weakening contribution of algae to the organic matter. At the initial stage of shale deposition (the mixed shale deposition period), the marine productivity was the highest and more 12C entered organisms, resulting in the negative deviation of δ13Corg [106,107]. Subsequently, marine productivity gradually decreased, with less 12C entering organisms, leading to an increase in the 13Corg isotope values upward.
Hein and Sand proposed that increased atmospheric CO2 concentrations could improve the primary productivity of marine phytoplankton in 1997 [109]. And many studies have confirmed that changes in atmospheric CO2 concentrations are negatively correlated with the evolution of oceanic organic carbon isotope δ13Corg [110]. Therefore, according to the longitudinal distribution trend shown in Figure 5, the δ13Corg values of the Lei3 member shale in the study area during the deposition period were positively biased from the lower mixed shale to the upper calcareous shale, which was related to the gradual decrease in CO2 concentration and productivity. In addition, the vertical decrease in CO2 concentration led to an increase in the CaCO3 production rate, meaning that the calcification effect is negatively correlated with ocean acidification [111,112]. Therefore, the vertical increase of carbonate content in the study area means that calcification is enhanced, and it has evolved into calcareous shale in the upper part of the study area, corresponding to relatively low TOC and weakened paleoproductivity (2) N isotope δ15Norg.
Vertically, the excursion of N isotopes and C isotopes was similar, with a slight increase in the upwards direction in the Lei3 member. The sea level during the deposition period of the Lei3 shale was relatively high, so denitrification and biological nitrogen fixation were strengthened against the background of higher productivity, resulting in negative excursion of δ15Norg in the sediment. Additionally, the water mass in the study area was restricted, and nitrate could not be fully supplemented. Nitrogen-fixing microorganisms absorbed atmospheric N2 and increased the concentration of 14N (biological organic matter), resulting in the relative decrease in N isotopes in the early stage of shale deposition.
However, long-term anoxic reductive conditions resulted in the relative enrichment of the remaining nitrate (NO3−) in the seawater inorganic nitrogen pool (DIN) by 15N. During the middle and late stages of the Lei3 shale deposition, the decrease in sea level promoted exchange between seawater and the atmosphere. On the one hand, the process of denitrification was restricted and decreased productivity. On the other hand, the biological absorption of residual 15N nitrate ions in seawater, which affected the N isotopic composition in the middle and late stages of shale deposition, eventually led to the positive excursion of δ15Norg.
(2)
Elemental geochemistry
P and Ba are the main elements used to assess productivity [48]. The content of Porg and Babio in sediments and sedimentary rocks is positively correlated with TOC, which can be a good indicator of productivity.
In the present study, biogenic P and biogenic Ba can be represented by Pxs and Baxs, respectively. In this study, the Pxs and Baxs values had a good correlation, implying that the accumulation of P was less disturbed by redox conditions and could be used as an effective paleoproductivity indicator. Moreover, in the upward direction, both of them showed a gradually decreasing trend, indicating that the ancient productivity gradually declined. The overall productivity of the HC125 well showed the following vertical trend: mixed shale > clay shale > calcareous shale.
Previous studies have shown that warm and wet climates are conducive to biological development, which increases primary productivity, while dry and cold climates decrease productivity [113,114]. There was a correlation between productivity and climate in the Sichuan Basin (Figure 5). The values of CIA, P/Ti, and Ba all showed slight decreasing trends, upwards, indicating that the climatic conditions and primary productivity were stable.
The elemental indexes were consistent with organic carbon isotope values and the CO2 curve, both of which showed a gradually decreasing trend in paleoproductivity in the vertical direction. The overall 13Corg values in the study area were relatively light, which was the source of algae. Marine productivity was the highest during the early stage of the shale deposition period of the Lei3 member (the mixed shale deposition period). Nevertheless, as the CIA value decreased vertically, which suggested that chemical weathering abated, the sea level dropped, and the overall salinity increased, which was unfavorable to algae breeding and led to a relative decline in ancient productivity, which was consistent with the 13Corg results. However, due to the gradual tectonic uplift and longitudinal physical weathering enhancement, the nutrients carried by terrigenous matter contributed some of the nutrients required by marine organisms; However, as these nutrient levels were lower than those brought by relatively strengthened chemical weathering, there was an overall downward trend in TOC in the vertical direction. So, there was no sudden decrease in organic matter content, which resulted in no significant positive excursions of organic carbon isotopes. Therefore, regional tectonic uplift had a certain impact on productivity, and climate change was the dominant factor affecting productivity.

5.2.4. Water Salinity

Typically, the depositional environment (marine or terrestrial) and water salinity can be inferred using the SrXS/BaXS ratio [115]. Variations in SrXS/BaXS ratios are generally positively correlated with paleosalinity, with high Sr content often associated with arid climates. Longitudinally, Lei S3 shale samples exhibit an overall increasing trend in SrXS/BaXS ratio, indicating a rise in salinity. This is interpreted as seawater concentration during marine regression, gradual aridification, and localization processes, leading to increased salt concentrations. Among the shale types, sedimentary water salinity increases in the following order: mixed shale, clay shale, and calcareous shale. Paleosalinity shows a clear negative correlation with TOC (Figure 6). This provides a reference for determining whether salt reduction can serve as an indicator of organic-rich accumulation in polymictic shales.

5.3. Water Column Restriction

The Mo/TOC ratio and the Mo-U covariant model are usually used to evaluate the restriction of a water mass [59,116].

5.3.1. Mo/TOC Pattern

The study area is in the Yangtze Platform, surrounded by uplifts or ancient land, and is connected to the open sea through a relatively narrow strait. Therefore, the study area is subject to certain restrictions. Mo is more enriched in anoxic environments than in dysoxic and oxygenated environments [59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117].
There was a strong positive correlation between Mo and TOC in the Lei3 member shale (Figure 6), indicating that the enrichment of Mo was closely related to the organic matter content and that it formed in an anoxic-dysoxic environment [48]. The shale of the Lei3 member was characterized by Mo enrichment and relatively low TOC, resulting in high Mo/TOC values. The Mo/TOC distribution range was 15–46.5 with a mean of 32.5; the Mo/TOC values of the Cariaco Basin and the Saanich Inlet were 25 and 45, respectively (Figure 13A).
The higher Mo content in the sedimentary period mixed shale may be due to the low degree of restriction, high sea level and smooth water exchange at that time. By the middle and late stages of shale deposition (clay shale to calcareous shale), a few samples fell into the Framvarent Fjord area, exhibiting a more confined character with significantly decreased elemental Mo. However, there was still a positive correlation between Mo and TOC, which may have been caused by regional tectonic uplift, the sea level decrease and the increased restriction of the water mass, resulting in insufficient Mo supplementation. These results revealed that the Lei3 shale in the study area was formed in neither an extremely confined environment nor a completely open environment.

5.3.2. MoEF-UEF Covariant Model

The Mo-U covariant model can be used to identify the restriction of a water mass and redox conditions [57,59,116]. There have been multiple authigenic Mo-U covariation models reported for the sedimentary period of the Lei3 number in the study area. Common to all these models are the positive covariant characteristics of Moauth (authigenic Mo) and Uauth (authigenic). As the TOC content increased, the degree of reduction in the water body increased, and the enrichment coefficients of Mo and U both increased with the same amplitude.
The MoEF/UEF values of most samples in the study area were between those of current seawater (Mo-EF/U-EF = 1*SW) and the “particulate shuttle” (Figure 13B). There was no downward trend as Mo/U increased until the particulate shuttle was reached (e.g., Mn-oxides particulate) [59]. This indicates that the sedimentary environment was not strongly restricted because the Mo supply rate in the water exceeded the Mo consumption rate in the sediment. When the water column is strongly reducing, the absorption of Mo increases faster than the absorption of U. Therefore, this result reflected the transition from anoxic to dysoxic conditions. The degree of restriction from clay shale to calcareous shale increases [57,59], especially as mixed shale samples showed higher MoEF and UEF values. Additionally, the MoEF and UEF values in upward sedimentary clay shale and calcareous shale gradually decreased, reflecting the decrease in oxygen minimum zone (OMZ), weakening reduction conditions and the decreased Mo and U in the water body [116,118,119].
The trends in UEF and MoEF were similar to those for Mo/TOC. This may have been because the sea level was generally high during the Lei3 shale deposition period, then gradually decreased over time. The regional tectonic movement led to the uplift of the study area, resulting in an increased degree of restriction.

5.4. Depositional Model for Organic Matter Accumulation

During the sedimentary period of Lei3 shale, the organic carbon isotope 13Corg in the lower part suddenly increased, and the corresponding elements changed significantly. For example, the paleoproductivity indexes (P, Ba, P/Ti) suddenly decreased, as did the terrigenous clastic indexes (Al and Ti), and the corresponding N isotopes showed an overall increasing trend. Redox indexes (V, Ni, U, Mo, Ni/Co, V/Cr, U/Th, V (V + Ni)) suddenly decreased. In fact, the correlation between TOC values and elemental geochemical indicators is used to identify the controlling factors of organic matter enrichment. TOC values and water salinity (Figure 14A) suggest that the products of halophilic organisms’ life activities and the input of terrestrial nutrients may not be the primary sources of organic matter. Indicators of reducing conditions in depositional waters, such as V/Cr, Ni/Co, and U/Th, show a positive correlation with TOC values (Figure 14B–D), highlighting the importance of reducing environments for the preservation of organic matter in shale. The positive correlation between PXS/Ti and TOC indicates the control of ancient water productivity on organic matter abundance (Figure 14E). Additionally, the negative correlation between terrestrial input and organic matter abundance suggests that terrestrial organic matter is not the main source of organic matter in Lei3 shale (Figure 14F,G). Enhanced chemical weathering and rising C-index both promote the enrichment of organic matter (Figure 14H,I).
In terms of the mineral content, the quartz content decreased, the clay content increased, and there was no significant change in the carbonate content. The sedimentary environment changed during this period, and clay shale began to be deposited. Then, upon entering the calcareous shale deposition period, the paleoproductivity index and redox environment index showed overall downward trends. However, in the middle of the calcareous shale deposition period, the redox index suddenly increased, and the paleoproductivity index also correspondingly increased. There may have been a short period of sea-level rise during this period, following which the sea level fell to the lowest value.
During the Lei3 shale deposition period, the lithofacies associations transitioned from mixed shale to clay shale and then changed to lithofacies dominated by calcareous shale. It was apparent that the lithofacies transformations represented changes in the sedimentary environment. Accordingly, based on the structure and sedimentary background of the Lei3 member shale as determined in this study, a sedimentary model of shale was established to explain the organic matter enrichment mechanism in different lithofacies associations.
In the early depositional stage of the Lei3 member, there was an extensive transgression [120]—even larger than the Olenekian transgression [10]—and mixed shale was deposited under this background. The 13Corg and 15N isotopes showed a negative bias during this stage, indicating that the sea level was high, the seawater was anoxic and organic matter was enriched. The CIA value was high, reflecting strong chemical weathering that provided high levels of nutrient minerals for plankton in the surface seawater. Additionally, with the warm and humid climate and the high levels of marine primary productivity, marine organisms absorbed atmospheric CO2, and their subsequent burial and storage in the seabed resulted in a decrease in CO2 partial pressure. This in turn promoted the gradual decline in temperature during the middle and later stages of Lei3 member formation, which was also consistent with the transition from a warm-humid climate to a semi-humid climate during the Lei3 deposition period.
Although the Sichuan Basin was in a shallow sea environment during this period, the water body was not as smooth as the open sea. This formed an anoxic-dysoxic or even a sulfidic bottom seawater environment. Such an environment resulted in the death of many benthic organisms, but it was highly conducive to the preservation of organic matter. Therefore, the TOC content of mixed shale was also the highest in the Lei3 shale. Vertically, the clay shale to calcareous shale deposition periods were increasingly affected by regional tectonic uplift, and the sea level gradually decreased. 13Corg and 15N isotopes both increased, and the climate changed from a warm-humid climate to a semi-humid climate.
The gradual decrease in seawater CO2 concentration resulted in a weakening of ocean acidification accompanied by an increase in the production rate of CaCO3 [111,112]. Therefore, in the study area, the longitudinal calcification was enhanced and the carbonate content increased, resulting in the deposition of mainly calcareous shale in the upper part of the shale. During this period, chemical weathering gradually weakened (but the CIA value was still high, generally greater than 70), and the input of nutrients decreased, which limited plankton reproduction. Compared with the early stage, the primary productivity gradually decreased.
This is also consistent with the paleo-sedimentary background at that time. After the P-T mass extinction event, the sedimentary environment did not improve in a direction that is obviously conducive to biological reproduction. Until entering the Anisian period, marine algae and other small organisms began to recover and change with the times. As the times changed, the paleoenvironment in the early Anisian period did not improve enough for large organisms to thrive. Subsequently, the temperature decreased from 36° to 30°. As the temperature became moderate [121], the oxygen content gradually increased and the CO2 concentration gradually decreased, suggesting that “calcification feedback” occurred during this period [122]. This negative feedback greatly promoted the gradual recovery of macrofauna and plants, and the development of macrofauna and plants began in the middle and late Anisian period.
Vertically, there was an increasing flux of debris (Al and Ti), and the TOC decreased (Figure 5), indicating that there was an organic matter dilution effect from debris inputs. Therefore, the input of terrigenous clastic materials was not the dominant factor of organic matter enrichment in the study area. It is certain that the sea level decreased during the Middle and Late Mesozoic (Anisian period) and the Sichuan Basin was in a relatively active tectonic environment. In tectonically active areas, the physical dissolution rate is usually higher than the chemical weathering rate and non-steady-state weathering often occurs [123]. Therefore, longitudinally, chemical weathering (CIA) weakened, but terrigenous debris increased.
While changes in paleoclimate, productivity, and redox conditions all played a role, the regional tectonic activity served as the overarching driver that orchestrated these changes. The argument for a tectonic driver is built upon a cascade of interconnected evidence. First, the observed increase in terrigenous input (Al, Ti) coincides with a decrease in chemical weathering intensity (CIA). This decoupling is a key indicator. A simple progradation or change in oceanic currents would likely increase terrigenous input, but it would not necessarily suppress chemical weathering. In contrast, regional tectonic uplift provides a coherent mechanism for both: it enhances physical erosion and sediment supply (increasing Al, Ti) while simultaneously creating topographic barriers that reduce hydrological connectivity and shift the climate towards a more sub-humid state (lowering CIA).
Furthermore, previous systematic paleogeographic and tectonic reconstructions of the Middle Triassic Sichuan Basin have delineated a distinctive sedimentary configuration in the study area. Over time, continental cracking of Pangea influenced the uplift of the overall structure of the Sichuan Basin and the gradual closure of the Paleo-Tethys Ocean basin [124]. The continental collision stage represented by the Indosinian movement began in the Middle and Late Triassic. The final collision and assembly of the South China Block and the North China Block arose from the combined effect of the stress from the Northwest and the reaction force of the Central Guizhou Uplift and the Jiangnan Uplift [125] under the background of the convergent stress field. The Jiangnan-Xuefeng orogenic belt on the southeastern margin of the Yangtze block gradually expanded and uplifted during the Middle Triassic. The Luzhou paleo-uplift [37] was formed during the process of compression and migration to the west. Due to the enhancement of physical erosion caused by uplifting, it has become an important provenance area (terrigenous fluxes) in the study area [37]. By the way, the tectonic model does not rely solely on Al and Ti trends. Instead, it integrates them with independent proxies for climate (CIA), basin restriction (Mo/TOC), and redox conditions (V/Cr, etc.), all of which show synchronous changes that are best explained by a common tectonic trigger.
The enhancement of physical weathering leads to an increase in terrigenous debris input but results in fewer nutrient inputs than from chemical weathering. Additionally, many terrigenous debris inputs dilute the proportion of organic matter in rock. Therefore, the input of terrigenous debris increased longitudinally, but the TOC decreased (Figure 5).
In summary, the global sea level and paleoclimate controlled the material basis of shale deposition, while the regional tectonic uplift promoted shale deposition. The study area has formed a large-scale starvation-type and relatively blocked retention environment, surrounded by a semi-restricted retention basin. Therefore, the Lei3 shale was formed under semi-closed conditions. The enrichment mechanism of organic matter in Lei3 shale is summarized as follows: the preservation mechanism was driven by phased regional tectonic movement and climate change. This is shown in Figure 15, below.

6. Conclusions

This integrated study of the Middle Triassic Lei3 shale in the Sichuan Basin yields the following principal conclusions:
  • The depositional environment evolved from a productive, anoxic system (mixed shale) to a less productive, dysoxic one (calcareous shale), concurrent with a climatic shift from warm-humid to sub-humid conditions. This transition is consistently recorded by geochemical proxies for redox (e.g., V/Cr, U/Th), productivity (Pxs, Baxs), and weathering (CIA).
  • Regional tectonic uplift, associated with the Indosinian Orogeny, is identified as the primary driver of this evolution. It induced a sea-level fall, enhanced physical weathering, increased basin restriction, and moderated the climate, thereby orchestrating the observed facies succession. This finding supports a tectonically modulated depositional model for organic matter accumulation.
  • The mixed shale facies, characterized by the highest residual TOC content and deposition under the most favorable environmental conditions, is identified as the most promising interval for shale gas exploration.
In summary, this work reconstructs the paleoenvironmental evolution of the Lei3 shale, establishes tectonic activity as its fundamental control, and defines the mixed shale as the primary exploration target, thereby contributing to both the understanding of Anisian ecosystem recovery and the evaluation of hydrocarbon resources in the Sichuan Basin.

Author Contributions

Conceptualization, Z.D., D.C., Y.H., X.Y., G.J. and F.H.; Methodology, Z.D., D.C., Y.H., X.Y., S.S. and F.H.; Software, Y.X. and G.J.; Validation, G.J. and B.M.; Formal analysis, D.C., Y.H., X.Y. and F.H.; Investigation, Z.D., D.C., S.S. and F.H.; Writing—original draft, Z.D.; Writing—review & editing, D.C., Y.H., X.Y., Y.X., G.J., S.S., B.M. and F.H.; Visualization, Z.D. and Y.H.; Supervision, Y.X. and B.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Sichuan Provincial Youth Science & Technology Innovative Research Group Fund [No. 2022JDTD0004] and the APC was funded by China University of Petroleum (Beijing).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors would like to acknowledge financial support provided by Sichuan Provincial Youth Science & Technology Innovative Research Group Fund (No. 2022JDTD0004).

Conflicts of Interest

Authors Zhenjing Du, Yaodong Xu, Guanbo Jiang, Shilong Shi, Bingjie Ma and Xiaomin Yang were employed by the Shengli Oilfield Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bottjer, D.J.; Clapham, M.E.; Fraiser, M.L.; Powers, C.M. Understanding mechanisms for the end-Permian mass extinction and the protracted Early Triassic aftermath and recovery. GSA Today 2008, 18, 4–10. [Google Scholar] [CrossRef]
  2. Alroy, J. Dynamics of origination and extinction in the marine fossil record. Proc. Natl. Acad. Sci. USA 2008, 105, 11536–11542. [Google Scholar] [CrossRef] [PubMed]
  3. Lehrmann, D.J.; Jiayong, W.; Enos, P. Controls on facies architecture of a large Triassic carbonate platform; the Great Bank of Guizhou, Nanpanjiang Basin, South China. J. Sediment. Res. 1998, 68, 311–326. [Google Scholar] [CrossRef]
  4. Lehmann, M.F.; Bernasconi, S.M.; Barbieri, A.; McKenzie, J.A. Preservation of organic matter and alteration of its carbon and nitrogen isotope composition during simulated and in situ early sedimentary diagenesis. Geochim. Cosmochim. Acta 2002, 66, 3573–3584. [Google Scholar] [CrossRef]
  5. Lehrmann, D.J.; Stepchinski, L.; Altiner, D.; Orchard, M.J.; Montgomery, P.; Enos, P.; Ellwood, B.B.; Bowring, S.A.; Ramezani, J.; Wang, H.; et al. An integrated biostratigraphy (conodonts and foraminifers) and chronostratigraphy (paleomagnetic reversals, magnetic susceptibility, elemental chemistry, carbon isotopes and geochronology) for the Permian–Upper Triassic strata of Guandao section, Nanpanjiang Basin, South China. J. Asian Earth Sci. 2015, 108, 117–135. [Google Scholar] [CrossRef]
  6. Ovtcharova, M.; Goudemand, N.; Hammer, Ø.; Guodun, K.; Cordey, F.; Galfetti, T.; Schaltegger, U.; Bucher, H. Developing a strategy for accurate definition of a geological boundary through radio-isotopic and biochronological dating: The early–middle Triassic boundary (South China). Earth-Sci. Rev. 2015, 146, 65–76. [Google Scholar] [CrossRef]
  7. Li, C. Origin and Geological Significances of Claystone (Mung BeanRock) in the Early-Middle Triassic Boundary, Western Margin of Sichuan Basin. Master’s Thesis, Chengdu University of Technology, Chengdu, China, 2021. (In Chinese). [Google Scholar]
  8. Sun, S.; Li, J.L.; Chen, H.H.; Peng, H.; Hsü, K.J.; Shelton, J.W. Mesozoic and Cenozoic sedimentary history of South China. AAPG Bull. 1989, 73, 1247–1269. [Google Scholar] [CrossRef]
  9. Zeng, H. Sedimentation and hydrocarbon potential of early triassic carbonate rocks of shiwan dashan basin in guangxi. Prtroleum Geol. Exp. 1984, 2, 95–100. (In Chinese) [Google Scholar]
  10. Sun, Y.D. Paleoclimate Impact and Biodiversity Response of Volcanism at the Paleo-Mesozoic Transition in South China. Ph.D. Thesis, China University of Geosciences, Wuhan, China, 2013. (In Chinese). [Google Scholar]
  11. Wang, Y.; Jiang, W.; Liu, H.; Liu, B.; Zheng, H.; Song, X.; Wang, Q.; Wang, W.; Li, Y. Sedimentary and Diagenetic Features and Their Impacts on Microbial Carbonate Reservoirs in the Fourth Member of the Middle Triassic Leikoupo Formation, Western Sichuan Basin, China. Energies 2020, 13, 2293. [Google Scholar] [CrossRef]
  12. Sun, H.F.; Luo, B.; Wen, L.; Wang, J.X.; Zhou, G.; Wen, H.G.; Huo, F.; Dai, X.; He, C.L. The first discovery of organic-rich shale in Leikoupo Formation and new areas of subsalt exploration, Sichuan Basin. Nat. Gas Geosci. 2021, 32, 233–247. [Google Scholar]
  13. Boucot, A.J.; Chen, X.; Scotese, C.R.; Fan, J.X. Phanerozoic Paleoclimate: An Atlas of Lithologic Indicators of Climate; Science Press: Beijing, China, 2009; Volume 1, pp. 90–102. (In Chinese) [Google Scholar]
  14. Chen, X.; Boucot, A.J.; Scotese, C.R.; Junxuan, F.; Yuan, W.; Xiujuan, Z. Pangaean aggregation and disaggregation with evidence from global climate belts. J. Palaeogeogr. 2012, 1, 5–13. [Google Scholar]
  15. Scotese, C.R. Paleomap Website. 2002. Available online: http://www.scotese.com (accessed on 5 May 2022).
  16. Tabor, N.J.; Poulsen, C.J. Palaeoclimate across the Late Pennsylvanian-Early Permian tropical palaeolatitudes: A review of climate indicators, their distribution, and relation to palaeophysiographic climate factors. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2008, 268, 293–310. [Google Scholar] [CrossRef]
  17. Dobson, J.P.; Heller, F. Triassic paleomagnetic results from the Yangtze Block, SE China. Geophys. Res. Lett. 1993, 20, 1391–1394. [Google Scholar] [CrossRef]
  18. Lin, Y.T.; He, J.Q. Study on palaeomagnetism and salt formation conditions of the stratum of Lower-Middle Trias Series (Tlj–T2j) of Sichuan Basin. J. Salt Lake Res. 2005, 13, 1–4, (In Chinese with English abstract). [Google Scholar]
  19. Opdyke, N.D.; Huang, K.; Xu, G.; Zhang, W.Y.; Kent, D.V. Paleomagnetic results from the Triassic of the Yangtze Platform. J. Geophys. Res.-Solid Earth Planets 1986, 91, 9553–9568. [Google Scholar] [CrossRef]
  20. Wu, H.; Zhu, R.; Bai, L.; Guo, B.; Lv, J.J. Revised apparent polar wander path of the Yangtze Block and its tectonic implications. Sci. China 1998, 41, 78–90. [Google Scholar]
  21. Zhu, R.X.; Yang, Z.Y.; Wu, H.N.; Ma, X.H.; Huang, B.C.; Meng, Z.F.; Fang, D.J. Paleomagnetic constraints on the tectonic history of the major blocks of China duing the Phanerozoic. Sci. China 1998, 41, 1–19. [Google Scholar]
  22. Guo, Z.W.; Deng, K.L.; Han, Y.H. The Formation and Development of Sichuan Basin; Geological Publishing House: Beijing, China, 1996; pp. 1–200. (In Chinese) [Google Scholar]
  23. He, D.F.; Li, D.S.; Zhang, G.W.; Zhao, L.Z.; Fan, C.; Lu, R.Q.; Wen, Z. Formation and evolution of multi-cycle superposed Sichuan Basin, China. Chin. J. Geol. 2011, 46, 589–606, (In Chinese with English abstract). [Google Scholar]
  24. Meng, Q.R.; Wang, E.; Hu, J.M. Mesozoic sedimentary evolution of the northwest Sichuan basin: Implication for continued clockwise rotation of the South China block. Geol. Soc. Am. Bull. 2005, 117, 396–410. [Google Scholar] [CrossRef]
  25. Enos, P.; Wei, J.; Lehrmann, D.J. Death in Guizhou-Late Triassic drowning of the Yangtze carbonate platform. Sediment. Geol. 1998, 118, 55–76. [Google Scholar] [CrossRef]
  26. Xu, X.D.; Liu, B.J. Sequence Boundary and Sea Level Changes in Western Margin of Upper Yangtze Platform during Permian and Triassic. J. China Univ. Geosci. 1996, 7, 105–111. [Google Scholar]
  27. Koenig, I.; Lougear, A.; Bruns, P.; Gruetzner, J.; Trautwein, A.X.; Dullo, W.C. Iron oxidation in sediment cores (Site 1062) during six months of storage in the Ocean Drilling Program archive. In Proceedings of the Ocean Drilling Program Scientific Results; 2001; Volume 172, pp. 25–35. Available online: https://www-odp.tamu.edu/publications/172_SR/chap_02/chap_02.htm (accessed on 13 November 2025).
  28. Li, S.G.; Hart, S.; Zheng, S.G.; Liu, D.L.; Zhang, G.W.; Guo, A.L. Sm-Nd isotopic evidences of collision between North China plate and South China plate. Sci. China (Ser. B) 1989, 3, 312–319, (In Chinese with English abstract). [Google Scholar]
  29. Li, S.G.; Sun, W.D.; Zhang, G.W.; Chen, J.Y.; Yang, Y.C. Chronology and geochemistry of metavolcanic rocks from Heigouxia Valley in the Mian-Lüe tectonic zone, South Qinling-Evidence for a Paleozoic oceanic basin and its close time. Sci. China (Ser. D) 1996, 39, 300–310. [Google Scholar]
  30. Meng, Q.; Zhang, G. Timing of collision of the North and South China blocks: Controversy and reconciliation. Geology 1999, 27, 123. [Google Scholar] [CrossRef]
  31. Zhang, G.W. Orogenesis and dynamics of the Qinling Orogen. Sci. China 1996, 3, 225–234. [Google Scholar]
  32. Chu, Y.; Faure, M.; Lin, W.; Wang, Q. Early Mesozoic tectonics of the South China block: Insights from the Xuefengshan intracontinental orogen. J. Asian Earth Sci. 2012, 61, 199–220. [Google Scholar] [CrossRef]
  33. Chu, Y.; Faure, M.; Lin, W.; Wang, Q.; Ji, W. Tectonics of the Middle Triassic intracontinental Xuefengshan Belt, South China: New insights from structural and chronological constraints on the basal décollement zone. Int. J. Earth Sci. 2012, 101, 2125–2150. [Google Scholar] [CrossRef]
  34. Wang, Y.; Fan, W.; Zhang, G.; Zhang, Y. Phanerozoic tectonics of the South China Block: Key observations and controversies. Gondwana Res. 2013, 23, 1273–1305. [Google Scholar] [CrossRef]
  35. Zhang, G.; Guo, A.; Wang, Y.; Li, S.; Dong, Y.; Liu, S.; He, D.; Cheng, S.; Lu, R.; Yao, A. Tectonics of South China continent and its implications. Sci. China 2013, 56, 1804–1828. [Google Scholar] [CrossRef]
  36. Faure, M.; Lin, W.; Chu, Y.; Lepvrier, C. Triassic tectonics of the southern margin of the South China Block. Comptes Rendus Geosci. 2016, 348, 5–14. [Google Scholar] [CrossRef]
  37. Huang, H.; He, D.; Li, Y.; Fan, H. Determination and formation mechanism of the Luzhou paleo-uplift in the southeastern Sichuan Basin. Earth Sci. Front. 2019, 26, 102–120. [Google Scholar]
  38. Feng, Z.Z.; Bao, Z.D.; Shenghe, W.; Yongtie, L.; Guoli, W. Lithofacies paleogeography of the early and middle triassic of south china. Chin. J. Geol. 1997, 2, 212–220. (In Chinese) [Google Scholar]
  39. Yin, H.F. Discussion on the lainian stage in china. Geol. Rev. 1982, 28, 235–239. (In Chinese) [Google Scholar]
  40. Benton, M.J.; Zhang, Q.; Hu, S.; Chen, Z.Q.; Wen, W.; Liu, J.; Huang, J.; Zhou, C.; Xie, T.; Tong, J.; et al. Reprint of “Exceptional vertebrate biotas from the Triassic of China, and the expansion of marine ecosystems after the Permo-Triassic mass extinction”. Earth-Sci. Rev. 2014, 137, 85–128. [Google Scholar] [CrossRef]
  41. SY/T5163-2010; Analysis Method for Clay Minerals and Ordinary Non-Clay Minerals in Sedimentary Rocks by the X-Ray Diffraction. National Energy Administration: Beijing, China, 2010.
  42. Huo, F.; Wen, H.G.; Li, L.; Luo, B.; Zhou, G.; Xu, W.L.; Sun, H.F.; Wang, X.Z.; Jiang, H.C.; Chen, S.Z.; et al. Influence of the depositional environment on the formation of organic-rich marine shale: A case study of the first discovery of Anisian shale in the Sichuan Basin. J. Pet. Sci. Eng. 2022, 214, 110577. [Google Scholar] [CrossRef]
  43. GB/T 19145-2003; Determination of Total Organic Carbon in Sedimentary Rock. General Administration of Quality Supervision. Inspection and Quarantine of the People’s Republic of China: Beijing, China, 2003.
  44. Kikumoto, R.; Tahata, M.; Nishizawa, M.; Sawaki, Y.; Maruyama, S.; Shu, D.; Han, J.; Komiya, T.; Takai, K.; Takai, K.; et al. Nitrogen isotope chemostratigraphy of the Ediacaran and Early Cambrian platform sequence at Three Gorges, South China. Gondwana Res. 2014, 25, 1057–1069. [Google Scholar] [CrossRef]
  45. Pi, D.H.; Liu, C.Q.; Shields-Zhou, G.A.; Jiang, S.Y. Trace and rare earth element geochemistry of black shale and kerogen in the early Cambrian Niutitang Formation in Guizhou province, south China: Constraints for redox environments and origin of metal enrichments. Precambrian Res. 2013, 225, 218–229. [Google Scholar] [CrossRef]
  46. Li, Y.F.; Zhang, T.W.; Ellis, G.S.; Shao, D.Y. Depositional environment and organic matter accumulation of Upper Ordovician-Lower Silurian marine shale in the Upper Yangtze Platform, South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2017, 466, 252–264. [Google Scholar] [CrossRef]
  47. Algeo, T.J.; Maynard, J.B. Trace-element behavior and redox facies in core shales of Upper Pennsylvanian Kansas-type cyclothems. Chem. Geol. 2004, 206, 289–318. [Google Scholar] [CrossRef]
  48. Tribovillard, N.; Algeo, T.J.; Lyons, T.; Riboulleau, A. Trace metals as paleoredox and paleoproductivity proxies: An update. Chem. Geol. 2006, 232, 12–32. [Google Scholar] [CrossRef]
  49. Nance, W.B.; Taylor, S.R. Rare earth element patterns and crustal evolution-I. Australian post-Archean sedimentary rocks. Geochim. Cosmochim. Acta 1976, 40, 1539–1551. [Google Scholar] [CrossRef]
  50. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell: Oxford, UK, 1985; 312p. [Google Scholar]
  51. Cox, R.; Lowe, D.R.; Cullers, R.L. The influence of sediment recycling and basement composition on evolution of mudrock chemistry in the southwestern United States. Geochim. Cosmochim. Acta 1995, 59, 2919–2940. [Google Scholar] [CrossRef]
  52. Schatz, A.K.; Scholten, T.; Kühn, P. Paleoclimate and weathering of the Tokaj (Hungary) loess–paleosol sequence. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2015, 426, 170–182. [Google Scholar] [CrossRef]
  53. McLennan, S.M.; Hemming, S.; McDaniel, D.K.; Hanson, G.N. Processes Controlling the Composition of Clastic Sediments; Geological Society of America: Boulder, CO, USA, 1993; pp. 21–40. [Google Scholar]
  54. Wang, C.; Zhang, B.Q.; Shu, Z.G.; Lu, Y.C.; Lu, Y.Q.; Bao, H.Y.; Li, Z.; Liu, C. Lithofacies types and reservoir characteristics of marine shales of the Wufeng Formation-Longmaxi Formation in Fuling area, the Sichuan Basin. Oil Gas Geol. 2018, 39, 485–497. (In Chinese) [Google Scholar]
  55. Lewan, M.D.; Maynard, J.B. Factors controlling enrichment of vanadium and nickel in the bitumen of organic sedimentary rocks. Geochim. Cosmochim. Acta 1982, 46, 2547–2560. [Google Scholar] [CrossRef]
  56. Algeo, T.J.; Ingall, E. Sedimentary Corg: P ratios, paleocean ventilation, and Phanerozoic atmospheric PO2. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2007, 256, 130–155. [Google Scholar] [CrossRef]
  57. Algeo, T.J.; Tribovillard, N. Environmental analysis of paleoceanographic systems based on molybdenum–uranium covariation. Chem. Geol. 2009, 268, 211–225. [Google Scholar] [CrossRef]
  58. Ross, D.J.K.; Bustin, R.M. Investigating the use of sedimentary geochemical proxies for paleoenvironment interpretation of thermally mature organic-rich strata: Examples from the Devonian-Mississippian shales, Western Canadian Sedimentary Basin. Chem. Geol. 2009, 260, 1–19. [Google Scholar] [CrossRef]
  59. Tribovillard, N.; Algeo, T.J.; Baudin, F.; Riboulleau, A. Analysis of marine environmental conditions based onmolybdenum-uranium covariation-Applications to Mesozoic paleoceanography. Chem. Geol. 2012, 324–325, 46–58. [Google Scholar] [CrossRef]
  60. Ruan, Y.B.; Zhou, G.; Huo, F.; Sun, H.F.; Guo, P.; Luo, T.; Jiang, H.C.; Wen, H.G. Source-reservoir characteristics and configuration of the third member of Middle Triassic Leikoupo Formation in central Sichuan Basin. Lithol. Reserv. 2023, 34, 139–151. (In Chinese) [Google Scholar]
  61. Zhang, B.J.; Sun, H.F.; Luo, Q.; Zeng, H.C.; Xu, L.; Li, L.; Huo, F.; Wen, H.G. Geochemical characteristics and hydrocarbon origin of Leikoupo Formation platform lagoon facies shale in central Sichuan Basin. Nat. Gas Ind. 2023, 43, 15–29. (In Chinese) [Google Scholar]
  62. Huang, W.; Meinschein, W.G. Sterols as ecological indicators. Geochim. Cosmochim. Acta 1979, 43, 739–745. [Google Scholar] [CrossRef]
  63. Lewan, M.D.; Bjorøy, M.; Dolcater, D.L. Effects of thermal maturation on steroid hydrocarbons as determined by hydrous pyrolysis of Phosphoria Retort Shale. Geochim. Cosmochim. Acta 1986, 50, 1977–1987. [Google Scholar] [CrossRef]
  64. Peters, K.E.; Walters, C.C.; Moldowan, J.M. The biomarker guide. In Biomarkers and Isotopes in the Environment and Human History; Cambridge University Press: Cambridge, UK, 2005; Volume 2, pp. 75–76. [Google Scholar]
  65. Xue, H.T.; Lu, S.F.; Zhong, N.N. Research on the lower limit of organic matter abundance in carbonate gas source rocks. Sci. China (Ser. D Earth Sci.) 2004, 34, 127–133. [Google Scholar]
  66. Farouk, S.; Lofty, N.M.; Qteishat, A.; Ahmad, F.; Shehata, A.M.; Al-Kahtany, K.; Hsu, C.S. Source and thermal maturity assessment of the Paleozoic-Mesozoic organic matter in the Risha gas field, Jordan. Fuel 2023, 335, 126998. [Google Scholar] [CrossRef]
  67. Qteishat, A.; El-Shafeiy, M.; Farouk, S.; Ahmad, F.; Al-Kahtany, K.; Gentzis, T.; Hamdy, D. Organic geochemical characterization and hydrocarbon generation modeling of Paleozoic-Paleogene shales, Wadi Sirhan basin, south-eastern Jordan. Mar. Pet. Geol. 2024, 170, 107152. [Google Scholar] [CrossRef]
  68. Jarvie, D.M.; Hill, R.J.; Ruble, T.E.; Pollastro, R.M. Unconventional shale-gas systems: The Mississippian Barnett Shale of north-central Texas as one model for thermogenic shale-gas assessment. AAPG Bull. 2007, 91, 475–499. [Google Scholar] [CrossRef]
  69. Tan, X.C.; Li, L.; Liu, H.; Cao, J.; Wu, X.Q.; Zhou, S.Y.; Shi, X.W. Mega-shoaling in carbonate platform of the middle Triassic Leikoupo formation, Sichuan basin, southwest China. Sci. Sin. 2014, 44, 457–471. [Google Scholar] [CrossRef]
  70. Du, X.; Song, X.D.; Zhang, M.Q.; Lu, Y.C.; Lu, Y.B.; Chen, P.; Liu, Z.H.; Yang, S. Shale gas potential of the lower Permian Gufeng formation in the western area of the lower Yangtze platform, China. Mar. Petrol. Geol. 2015, 67, 526–543. [Google Scholar] [CrossRef]
  71. Egbobawaye, E.I. Petroleum source-rock evaluation and hydrocarbon potential in Montney Formation unconventional reservoir, Northeastern British Columbia, Canada. Nat. Resour. 2017, 8, 716–756. [Google Scholar] [CrossRef]
  72. Ibad, S.M.; Padmanabhan, E. Methane sorption capacities and geochemical characterization of Paleozoic shale Formations from Western Peninsula Malaysia: Implication of shale gas potential. Int. J. Coal Geol. 2020, 224, 103480. [Google Scholar] [CrossRef]
  73. Lu, J.; Zhou, K.; Yang, M.; Shao, L. Jurassic continental coal accumulation linked to changes in palaeoclimate and tectonics in a fault epression superimposed basin, Qaidam Basin, NW China. Geol. J. 2020, 55, 7998–8016. [Google Scholar] [CrossRef]
  74. Didyk, B.M.; Simoneit, B.R.; Brassell, S.C.; Eglinton, G. Organic geochemical indicators of palaeoenvironmental conditions of sedimentation. Nature 1978, 272, 216–222. [Google Scholar] [CrossRef]
  75. International Committee for Coal and Organic Petrology (ICCP). The new vitrinite classification (ICCP System 1994). Fuel 1998, 77, 349–358. [Google Scholar] [CrossRef]
  76. Zhang, L.P.; Huang, D.F.; Liao, Z.Q. Gammacerane-geochemical indicator of water column stratification. Acta Sedimentol. Sin. 1999, 1, 136–141. (in Chinese). [Google Scholar]
  77. Wang, Z.; Fu, X.; Feng, X.; Song, C.; Wang, D.; Chen, W.; Zeng, S. Geochemical features of the black shales from the Wuyu Basin, southern Tibet: Implications for palaeoenvironment and palaeoclimate. Geol. J. 2017, 52, 282–297. [Google Scholar] [CrossRef]
  78. Fedo, C.M.; Nesbitt, H.W.; Young, G.M. Unraveling the effects of potassium metasomatism in sedimentary-rocks and paleosols, with implications for paleoweathering conditions and provenance. Geology 1995, 23, 921–924. [Google Scholar] [CrossRef]
  79. Harnois, L. The CIW index: A new chemical index of weathering. Sediment. Geol. 1988, 55, 319–322. [Google Scholar] [CrossRef]
  80. Nesbitt, H.W.; Young, G.M. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 1982, 299, 715–717. [Google Scholar] [CrossRef]
  81. Cullers, R.L.; Basu, A.; Suttner, L.J. Geochemical signature of provenance in sand-size material in soils and stream sediments near the Tobacco Root batholith, Montana, USA. Chem. Geol. 1988, 70, 335–348. [Google Scholar] [CrossRef]
  82. Garzanti, E.; Padoan, M.; Setti, M.; Najman, Y.; Peruta, L.; Villa, I.M. Weathering geochemistry and Sr-Nd fingerprints of equatorial upper Nile and Congo muds. Geochem. Geophys. Geosyst. 2013, 14, 292–316. [Google Scholar] [CrossRef]
  83. Morey, G.B.; Setterholm, D.R. Rare earth elements in weathering profiles and sediments of Minnesota: Implications for provenance studies. J. Sediment. Res. 1997, 67, 105–115. [Google Scholar] [CrossRef]
  84. Nesbitt, H.W.; Young, G.M. Petrogenesis of sediments in the absence of chemical weathering: Effects of abrasion and sorting on bulk composition and mineralogy. Sedimentology 1996, 43, 341–358. [Google Scholar] [CrossRef]
  85. Singh, P.; Rajamani, V. REE geochemistry of recent clastic sediments from the Kaveri floodplains, southern India: Implication to source area weathering and sedimentary processes. Geochim. Cosmochim. Acta 2001, 65, 3093–3108. [Google Scholar] [CrossRef]
  86. Yang, J.; Du, Y.; Cawood, P.A.; Xu, Y. Modal and geochemical compositions of the lower silurian clastic rocks in north qilian, nw china: Implications for provenance, chemical weathering, and tectonic setting. J. Sediment. Res. 2012, 82, 92–103. [Google Scholar] [CrossRef]
  87. Xu, X.T.; Sao, L.Y. Limiting factors in utilization of chemical index of alteration of mudstones to quantify the degree of weathering in provenance. J. Palaeogeogr. 2018, 20, 515–522. (In Chinese) [Google Scholar]
  88. Nesbitt, H.W.; Young, G.M. Prediction of some weathering trends of plutonic and volcanic rocks based on thermodynamic and kinetic considerations. Geochim. Cosmochim. Acta 1984, 48, 1523–1534. [Google Scholar] [CrossRef]
  89. Cao, J.; Wu, M.; Chen, Y.; Hu, K.; Bian, L.; Wang, L.; Zhang, Y. Trace and rare earth element geochemistry of Jurassic mudstones in thenorthern Qaidam Basin, northwest China. Chem. Erde/Geochem. 2012, 72, 245–252. [Google Scholar]
  90. Feng, X.L.; Fu, X.G.; Tan, F.W.; Chen, W.B. Sedimentary Environment Characteristics of Upper Carboniferous Cameng Formation in Kongkong Chaka Area of Northern Qiangtang Basin, Tibet. Geoscience 2014, 28, 953–961. (In Chinese) [Google Scholar]
  91. Li, J. Characteristics of Organic Carbon, Carbon Nitrogen Ratio and Organic Carbon Isotope in Lake Surface Sediments of Northern Tibetan Plateau and Their Environmental Significance. Master’s Thesis, Lanzhou University, Lanzhou, China, 2014. (In Chinese). [Google Scholar]
  92. Freudenthal, T.; Wagner, T.; Wenzhofer, F.; Zabel, M.; Wefer, G. Early diagenesis of organic matter from sediments of the eastern subtropical Atlantic: Evidence from stable nitrogen and carbon isotopes. Geochim. Cosmochim. Acta 2001, 65, 1795–1808. [Google Scholar] [CrossRef]
  93. Meyers, P.A. Preservation of elemental and isotopic source identification of sedimentary organic matter. Chem. Geol. 1994, 114, 289–302. [Google Scholar] [CrossRef]
  94. Teng, G.E.; Liu, W.H.; Xu, Y.C.; Chen, J.F. Identification of effective source rocks of Ordovician Marine sediments in Ordos Basin. Prog. Nat. Sci. 2004, 14, 1249–1256. (In Chinese) [Google Scholar]
  95. Lei, L.D. Marine Redox Variability and Its Coupling with Oceanic C-N Biogeochemical Cycling Across the Permian-Triassic Boundary in South China. Ph.D. Thesis, China University of Geosciences, Wuhan, China, 2017. (In Chinese). [Google Scholar]
  96. Sigman, D.M.; DiFiore, P.J.; Hain, M.P.; Deutsch, C.; Wang, Y.; Karl, D.M.; Knapp, A.N.; Lehmann, M.F.; Pantoja, S. The dual isotopes of deep nitrate as a constraint on the cycle and budget of oceanic fixed nitrogen. Deep. Sea Res. Part I Oceanogr. Res. Pap. 2009, 56, 1419–1439. [Google Scholar] [CrossRef]
  97. Ader, M.; Sansjofre, P.; Halverson, G.P.; Busigny, V.; Trindade, R.I.F.; Kunzmann, M.; Nogueira, A.C.R. Ocean redox structure across the Late Neoproterozoic Oxygenation Event: A nitrogen isotope perspective. Earth Planet. Sci. Lett. 2014, 396, 1–13. [Google Scholar] [CrossRef]
  98. Stüeken, E.E.; Kipp, M.A.; Koehler, M.C.; Buick, R. The evolution of Earth’s biogeochemical nitrogen cycle. Earth-Sci. Rev. 2016, 160, 220–239. [Google Scholar] [CrossRef]
  99. Rhoads, D.C.; Morse, J.W. Evolutionary and ecologic significance of oxygen-deficient marine basins. Lethaia 1971, 4, 413–428. [Google Scholar] [CrossRef]
  100. Wei, H.; Shen, J.; Schoepfer, S.D.; Krystyn, L.; Richoz, S.; Algeo, T.J. Environmental controls on marine ecosystem recovery following mass extinctions, with an example from the Early Triassic. Earth-Sci. Rev. 2015, 149, 108–135. [Google Scholar] [CrossRef]
  101. Algeo, T.J.; Liu, J. A re-assessment of elemental proxies for paleoredox analysis. Chem. Geol. 2020, 540, 119549. [Google Scholar] [CrossRef]
  102. Song, H.; Wignall, P.B.; Tong, J.; Bond, D.P.G.; Song, H.; Lai, X.; Zhang, K.; Wang, H.; Chen, Y. Geochemical evidence from bio-apatite for multiple oceanic anoxic events during Permian–Triassic transition and the link with end-Permian extinction and recovery. Earth Planet. Sci. Lett. 2012, 353–354, 12–21. [Google Scholar] [CrossRef]
  103. Jones, B.; Manning, D.A. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chem. Geol. 1994, 111, 111–129. [Google Scholar] [CrossRef]
  104. Wignall, P.B.; Twitchett, R.J. Oceanic anoxia and the end Permian mass extinction. Science 1996, 272, 1155–1158. [Google Scholar] [CrossRef]
  105. Rimmer, S.M. Geochemical paleoredox indicators in Devonian-Mississippian black shales, central Appalachian Basin (USA). Chem. Geol. 2004, 206, 373–391. [Google Scholar] [CrossRef]
  106. Mckenzie, J.A. Carbon-13 cycle in Lake Greifen: A model for restricted ocean basins. In Nature and Origin of Cretaceous Carbon-Rich Facies; Academic Press: Cambridge, MA, USA, 1982; pp. 197–207. [Google Scholar]
  107. Mckenzie, J.A. Carbon isotopes and productivity in the lacustrine and marine environment. In Chemical Processes in Lakes; Wiley-Blackwell: Hoboken, NJ, USA, 1985; pp. 99–118. [Google Scholar]
  108. Yin, X.Z. Paleolake Productivity in Central Songliao Basin During the Early Late Cretaceous. Ph.D. Thesis, China University of Geosciences, Beijing, China, 2008. (In Chinese). [Google Scholar]
  109. Hein, M.; SandJensen, K. CO2 increases oceanic primary production. Nature 1997, 388, 526–527. [Google Scholar] [CrossRef]
  110. Des Marais, D.J. Isotopic evolution of the biogeochemical carbon cycle during the Proterozoic Eon. Org. Geochem. 1997, 27, 185–193. [Google Scholar] [CrossRef]
  111. Kleypas, J.A.; Buddemeier, R.W.; Eakin, C.M.; Gattuso, J.P.; Guinotte, J.; Hoegh-Guldberg, O.; Iglesias-Prieto, R.; Jokiel, P.L.; Langdon, C.; Skirving, W.; et al. Comment on “Coral reef calcification and climate change: The effect of ocean warming”. Geophys. Res. Lett. 2005, 32, L08601. [Google Scholar] [CrossRef]
  112. Lombard, F.; da Rocha, R.E.; Bijma, J.; Gattuso, J.P. Effect of carbonate ion concentration and irradiance on calcification in planktonic foraminifera. Biogeosciences 2010, 7, 247–255. [Google Scholar] [CrossRef]
  113. Li, J.; Sha, J.; McLoughlin, S.; Wang, X. Mesozoic and Cenozoic palaeogeography, palaeoclimate and palaeoecology in the eastern Tethys. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 515, 1–5. [Google Scholar] [CrossRef]
  114. Meng, Q.; Liu, Z.; Bruch, A.A.; Liu, R.; Hu, F. Palaeoclimatic evolution during Eocene and its influence on oil shale mineralisation, Fushun basin, China. J. Asian Earth Sci. 2012, 45, 95–105. [Google Scholar] [CrossRef]
  115. Armstrong-Altrin, J.S.; Lee, Y.I.; Kasper-Zubillaga, J.J.; Trejo-Ramírez, E. Mineralogy and geochemistry of sands along the Manzanillo and El Carrizal beach areas, southern Mexico: Implications for palaeoweathering, provenance and tectonic setting. Geol. J. 2016, 52, 559–582. [Google Scholar] [CrossRef]
  116. Algeo, T.J.; Lyons, T.W. Mo–total organic carbon covariation in modern anoxic marine environments: Implications for analysis of paleoredox and paleohydrographic conditions. Paleoceanography 2006, 21, 1–23. [Google Scholar] [CrossRef]
  117. Helz, G.R.; Miller, C.V.; Charnock, J.M.; Mosselmans, J.F.W.; Pattrick, R.A.D.; Garner, C.D.; Vaughan, D.J. Mechanism of molybdenum removal from the sea and its concentration in black shales: EXAFS evidence. Geochim. Cosmochim. Acta 1996, 60, 3631–3642. [Google Scholar] [CrossRef]
  118. Sweere, T.; van den Boorn, S.; Dickson, A.J.; Reichart, G.J. Definition of new trace-metal proxies for the controls on organic matter enrichment in marine sediments based on Mn, Co, Mo and Cd concentrations. Chem. Geol. 2016, 441, 235–245. [Google Scholar] [CrossRef]
  119. Lu, Y.; Huang, C.; Jiang, S.; Zhang, J.; Lu, Y.; Liu, Y. Cyclic late Katian through Hirnantian glacioeustasy and its control of the development of the organic-rich Wufeng and Longmaxi shales, South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 526, 96–109. [Google Scholar] [CrossRef]
  120. Wu, S.X.; Li, H.T.; Long, S.X.; Liu, Z.L.; Wang, C.L. A study on characteristics and diagenesis of carbonate reservoirs in the Middle Triassic Leikoupo Formation in western Sichuan Depression. Pet. Geol. 2011, 32, 542–550. (In Chinese) [Google Scholar]
  121. Trotter, J.A.; Williams, I.S.; Nicora, A.; Mazza, M.; Rigo, M. Long-term cycles of Triassic climate change: A new δ18O record from conodont apatite. Earth Planet. Sci. Lett. 2015, 415, 165–174. [Google Scholar] [CrossRef]
  122. Sarmiento, J.L.; Gruber, N. Ocean Biogeochemical Dynamics; Princeton University Press: Princeton, NJ, USA; Oxford, UK, 2006; 528p. [Google Scholar]
  123. Nesbitt, H.W.; Fedo, C.M. Quartz and Feldspar Stability, Steady and Non-steady-State Weathering, and Petrogenesis of Siliciclastic Sands and Muds. J. Geol. 1997, 105, 173–191. [Google Scholar] [CrossRef]
  124. An, Z.X. Paleogeology at the end of Triassic in Sichuan Basin. Acta Geol. Sin. 1962, 42, 20–28. (In Chinese) [Google Scholar]
  125. Zhao, Z.; Zhu, Y.; Deng, H.; Xu, Y. Control of paleouplifts to the meso-paleozoic primary oil and gas pools in the south of China. Exp. Pet. Geol. 2003, 1, 10–18. (In Chinese) [Google Scholar]
Figure 1. Geology of South China in the Middle Triassic. (A) Tectonic map illustrating major blocks of South China of the Middle Triassic interpreted suture zones and extent of the Nanpanjiang Basin and Yangtze Platform. ‘Old land’ is brown, shallow seas are pale blue, and deep marine basins are dark blue. (B). Lithostratigraphic column of the Central Sichuan Basin, showing major gas combination. Modified from reference [40]. (C). Stratigraphic column and shale sample (for subsequent geochemical analysis) location of the Lei3 shale of the HC125 Well.
Figure 1. Geology of South China in the Middle Triassic. (A) Tectonic map illustrating major blocks of South China of the Middle Triassic interpreted suture zones and extent of the Nanpanjiang Basin and Yangtze Platform. ‘Old land’ is brown, shallow seas are pale blue, and deep marine basins are dark blue. (B). Lithostratigraphic column of the Central Sichuan Basin, showing major gas combination. Modified from reference [40]. (C). Stratigraphic column and shale sample (for subsequent geochemical analysis) location of the Lei3 shale of the HC125 Well.
Minerals 15 01230 g001
Figure 2. Ternary diagram showing the mineralogy compositions of the three lithofacies in the Lei3 member of HC125 well [54].
Figure 2. Ternary diagram showing the mineralogy compositions of the three lithofacies in the Lei3 member of HC125 well [54].
Minerals 15 01230 g002
Figure 3. Photographs showing the characteristics of common minerals in Lei3 member. (A) HC125, calcareous shale, 2275.2 m, crack filling calcium, plane polarized light; (B) HC125, clay shale, 2291.6 m, visible calcite particles and bitumen/clay, plane polarized light; (C) HC125, mixed shale, 2313.5 m, shale visible calcareous and bitumen, plane polarized light; (D) HC125, calcareous shale, 2275.2 m, visible calcite particles and clay minerals, SEM; (E) HC125, clay shale, 2291.6 m, shale visible in a large number of clay minerals, SEM; (F) HC125, 2313.5 m, calcareous, pyrite and quartz minerals are visible in the mixed shale, SEM; (G). HC125, 2291.6 m, Sapropelinite and Exinite; (H) HC125, 2275.2 m, Sapropelinite and Exinite; (I) HC125, 2313.5 m, Sapropelinite and Exinite.
Figure 3. Photographs showing the characteristics of common minerals in Lei3 member. (A) HC125, calcareous shale, 2275.2 m, crack filling calcium, plane polarized light; (B) HC125, clay shale, 2291.6 m, visible calcite particles and bitumen/clay, plane polarized light; (C) HC125, mixed shale, 2313.5 m, shale visible calcareous and bitumen, plane polarized light; (D) HC125, calcareous shale, 2275.2 m, visible calcite particles and clay minerals, SEM; (E) HC125, clay shale, 2291.6 m, shale visible in a large number of clay minerals, SEM; (F) HC125, 2313.5 m, calcareous, pyrite and quartz minerals are visible in the mixed shale, SEM; (G). HC125, 2291.6 m, Sapropelinite and Exinite; (H) HC125, 2275.2 m, Sapropelinite and Exinite; (I) HC125, 2313.5 m, Sapropelinite and Exinite.
Minerals 15 01230 g003
Figure 4. The vertical evolution of mineral content and TOC values of the Lei3 shale in HC125 Well.
Figure 4. The vertical evolution of mineral content and TOC values of the Lei3 shale in HC125 Well.
Minerals 15 01230 g004
Figure 5. The vertical evolution of paleoproductivity (Pxs/Ti, Pxs, Baxs, CO2 concentration) and Paleoclimate (CIA, δ13Corg, δ15Norg isotope) and terrigenous fluxes (Al, Ti) proxies in Lei3 shale of HC125 well.
Figure 5. The vertical evolution of paleoproductivity (Pxs/Ti, Pxs, Baxs, CO2 concentration) and Paleoclimate (CIA, δ13Corg, δ15Norg isotope) and terrigenous fluxes (Al, Ti) proxies in Lei3 shale of HC125 well.
Minerals 15 01230 g005
Figure 6. The vertical evolution of redox (Vxs, Nixs, Vxs/Crxs, Nixs/Coxs, Uxs/Thxs, Vxs/(Vxs + Nixs) and water column restriction (Uxs and Moxs) proxies in Lei3 shale of HC125 well.
Figure 6. The vertical evolution of redox (Vxs, Nixs, Vxs/Crxs, Nixs/Coxs, Uxs/Thxs, Vxs/(Vxs + Nixs) and water column restriction (Uxs and Moxs) proxies in Lei3 shale of HC125 well.
Minerals 15 01230 g006
Figure 7. Distribution of n-alkanes in Lei3-2 Shale. (A) HC125 well, mixed shale, 2311.55 m, Lei3; (B) HC125 well, clay shale, 2303.45 m, Lei3; (C) CT1, Anisian [60]; (D) S47, Anisian [61].
Figure 7. Distribution of n-alkanes in Lei3-2 Shale. (A) HC125 well, mixed shale, 2311.55 m, Lei3; (B) HC125 well, clay shale, 2303.45 m, Lei3; (C) CT1, Anisian [60]; (D) S47, Anisian [61].
Minerals 15 01230 g007
Figure 8. Ratio diagram of Pr/nC17 and Ph/nC18. The data for HC125 shale source rocks falls within Zone of algal OM reducing environment.
Figure 8. Ratio diagram of Pr/nC17 and Ph/nC18. The data for HC125 shale source rocks falls within Zone of algal OM reducing environment.
Minerals 15 01230 g008
Figure 9. Distribution of Sesquiterpanes and Steranes in Lei3 shale. (A) HC125, Mixed Shale, 2311.55 m, Lei3-2; (B) HC125, Clay Shale, 2303.45 m, Lei3-2; (C) JY1, 2642.38 m [42]; (D) CT1, 3560.75 m [42].
Figure 9. Distribution of Sesquiterpanes and Steranes in Lei3 shale. (A) HC125, Mixed Shale, 2311.55 m, Lei3-2; (B) HC125, Clay Shale, 2303.45 m, Lei3-2; (C) JY1, 2642.38 m [42]; (D) CT1, 3560.75 m [42].
Minerals 15 01230 g009
Figure 10. Chemical compositional variations in the Lei4 shale in the A–CN–K ternary diagram and associated chemical index of alteration (CIA) variations [88].
Figure 10. Chemical compositional variations in the Lei4 shale in the A–CN–K ternary diagram and associated chemical index of alteration (CIA) variations [88].
Minerals 15 01230 g010
Figure 11. Cross plots of redox proxies of Lei3 shale in HC125. (A) Cross plots of Sr/Cu and C-value. (B) cross plots of Sr/Ba and C-value.
Figure 11. Cross plots of redox proxies of Lei3 shale in HC125. (A) Cross plots of Sr/Cu and C-value. (B) cross plots of Sr/Ba and C-value.
Minerals 15 01230 g011
Figure 13. (A) Mo-TOC correlations showing the water column restriction level of Lei3 shale [116]; (B) Covariation of MoEF and UEF [57] showing the water column redox conditions of Lei3 shale.
Figure 13. (A) Mo-TOC correlations showing the water column restriction level of Lei3 shale [116]; (B) Covariation of MoEF and UEF [57] showing the water column redox conditions of Lei3 shale.
Minerals 15 01230 g013
Figure 14. Correlation of TOC with Geochemical Indicators of Lei3 shale. (A) Correlation of SrXS/BaXS and TOC; (B) Correlation of VXS/CrXS and TOC; (C) Correlation of NiXS/CoXS and TOC; (D) Correlation of UXS/ThXS and TOC; (E) Correlation of PXS/Ti and TOC; (F) Correlation of Ti and TOC; (G) Correlation of Al and TOC; (H) Correlation of CIA and TOC; (I) Correlation of C index and TOC.
Figure 14. Correlation of TOC with Geochemical Indicators of Lei3 shale. (A) Correlation of SrXS/BaXS and TOC; (B) Correlation of VXS/CrXS and TOC; (C) Correlation of NiXS/CoXS and TOC; (D) Correlation of UXS/ThXS and TOC; (E) Correlation of PXS/Ti and TOC; (F) Correlation of Ti and TOC; (G) Correlation of Al and TOC; (H) Correlation of CIA and TOC; (I) Correlation of C index and TOC.
Minerals 15 01230 g014
Figure 15. Sedimentary model of the Lei3 shale of HC125. The location of well and variations in shale components are schematic. (A) Early sedimentary period of Lei3 shale, accompanied by mixed shale deposition. (B) Middle sedimentary period of Lei3 shale, accompanied by clay shale deposition. (C) Late sedimentary period of Lei3 shale, accompanied by calcareous shale deposition.
Figure 15. Sedimentary model of the Lei3 shale of HC125. The location of well and variations in shale components are schematic. (A) Early sedimentary period of Lei3 shale, accompanied by mixed shale deposition. (B) Middle sedimentary period of Lei3 shale, accompanied by clay shale deposition. (C) Late sedimentary period of Lei3 shale, accompanied by calcareous shale deposition.
Minerals 15 01230 g015
Table 1. Comparison of Residual TOC and Tmax of the Lei3 Shale with Selected Paleozoic Shale Gas Rocks Worldwide.
Table 1. Comparison of Residual TOC and Tmax of the Lei3 Shale with Selected Paleozoic Shale Gas Rocks Worldwide.
FormationAgeTOC (wt%)S1 (mgHC/g Rock)S2 (mgHC/g Rock)Tmax (°C)HI (mg HC/g TOC)OI (mg CO2/g TOCReference
Longmaxi fm. (China)Lower Silurian0.34–40.02–0.050.26–0.48313–6018–388–69[69]
Niutitang fm. (China)Lower Cambrian0.39–10.20.01–0.050.09–0.62281–6003–864–240[69]
Muskwa, Besa & Fort Simpson Fm. (Canada)Devonian- Mississippian0.18–4.720.01–0.040.04–0.24369–6111–94–12[58]
Barnett Shale (USA)Mississippian2.62–11.470.26–3.60.59–54.53425–54414–475 [68]
Gufeng Fm. (China)Lower Permian0.04–22.10–0.310–7.95339–5540–603–429[70]
Montney Fm. (Canada)Lower Triassic0.03–8.20.02–0.810–1.83347–5261–3670–933[71]
Baling and Bendang Riang Fm.Silurian-Devonian0.73–24.60.01–0.080.02–0.37322–5020.28–25.180.89–40.89[72]
Timah Tasoh and Sanai Fm.Devonian0.1–9.10.01–0.070.02–0.14330–5020.55–1102.79–900[72]
Kubang Pasu (Perlis), Singa, Batu Gajah Fm.Carboniferous0.06–2790.01–0.060.02–0.14301–60815–115.942.27–598.43[72]
Kubang Pasu (Kedah) Fm.L. Permian1.01–19.650.01–0.050.02–0.09328–5020.21–3.334.27–15[42]
Leikoupo fm. (China)Middle Triassic0.11–4.010.01–0.110.02–0.6424–592 This study
Note: Due to the high maturity of the Lei3 samples, Rock-Eval parameters S1, S2, HI, and OI are not comparable. The comparison is focused on TOC and thermal maturity (Tmax).
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

Du, Z.; Chen, D.; Huang, Y.; Yang, X.; Xu, Y.; Jiang, G.; Shi, S.; Ma, B.; Huo, F. Petrology and Geochemistry Features of the Middle Triassic Anisian Shale in Sichuan Basin, South China: Implications for Climatic and Environmental Condition Change. Minerals 2025, 15, 1230. https://doi.org/10.3390/min15121230

AMA Style

Du Z, Chen D, Huang Y, Yang X, Xu Y, Jiang G, Shi S, Ma B, Huo F. Petrology and Geochemistry Features of the Middle Triassic Anisian Shale in Sichuan Basin, South China: Implications for Climatic and Environmental Condition Change. Minerals. 2025; 15(12):1230. https://doi.org/10.3390/min15121230

Chicago/Turabian Style

Du, Zhenjing, Dongxia Chen, Yuhan Huang, Xiaomin Yang, Yaodong Xu, Guanbo Jiang, Shilong Shi, Bingjie Ma, and Fei Huo. 2025. "Petrology and Geochemistry Features of the Middle Triassic Anisian Shale in Sichuan Basin, South China: Implications for Climatic and Environmental Condition Change" Minerals 15, no. 12: 1230. https://doi.org/10.3390/min15121230

APA Style

Du, Z., Chen, D., Huang, Y., Yang, X., Xu, Y., Jiang, G., Shi, S., Ma, B., & Huo, F. (2025). Petrology and Geochemistry Features of the Middle Triassic Anisian Shale in Sichuan Basin, South China: Implications for Climatic and Environmental Condition Change. Minerals, 15(12), 1230. https://doi.org/10.3390/min15121230

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