Next Article in Journal
Fast Initial Model Design for Electrical Resistivity Inversion by Using Broad Learning Framework
Previous Article in Journal
Petrographic Insights into the Evolution of Nano-Scale Organic Matter Pores with Organic Matter Conversion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organic Matter Enrichment Mechanisms in the Lower Cambrian Shale: A Case Study from Xiangandi #1 Well

1
Institute of Geomechanics, Chinese Academy of Geological Sciences, Beijing 100081, China
2
Key Laboratory of Petroleum Geomechanics, Chinese Academy of Geological Sciences, Beijing 100081, China
3
Key Laboratory of Paleomagnetism and Tectonic Reconstruction of Ministry of Natural Resources, Chinese Academy of Geological Sciences, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(2), 183; https://doi.org/10.3390/min14020183
Submission received: 6 December 2023 / Revised: 24 December 2023 / Accepted: 27 December 2023 / Published: 8 February 2024

Abstract

:
In order to investigate the effect of primary productivity, organic matter dilution, and preservation on the accumulation of organic matter, geochemical data, and proxies of primary productivity, clastic influx, and redox conditions were obtained for organic-rich shales in the Cambrian Niutitang Formation. The primary productivity (total organic carbon [TOC], Mo, P, Ba, and Babio) and redox (Ni/Co, V/Cr, U/Al, and Th/U) proxies suggest the organic-rich shales were deposited in anoxic-euxinic conditions during periods of high primary productivity. Pyrite in the Niutitang Formation comprises spherical framboids, which also indicate that anoxic bottom waters were present during organic matter deposition. High primary productivity enhanced the organic C flux into the thermocline layer and bottom waters, which lead to the development of anoxic bottom waters owing to O2 consumption by microorganisms and organic matter degradation. The anoxic bottom waters were beneficial for the preservation of organic matter. In addition, Ti/Al ratios correlate well with TOC contents throughout the Niutitang Formation, indicating that clastic input increased the burial rate and prevented organic matter degradation during deposition. Therefore, the accumulation of organic matter in the Niutitang Formation was controlled mainly by primary productivity rather than bottom-water redox conditions.

1. Introduction

Variations of organic matter accumulation in sediments are relatively large, as a result of a complex interaction between productivity, preservation, and dilution of organic matter [1,2,3]. The primary productivity associated with different nutrient and organic matter types can provide important insights into the organic matter in shales and mudstones [4,5,6,7]. Dilution of organic matter is a complex function of sedimentation rates [8,9] and biogenic alteration of inorganic matter [2,10]. The redox conditions of seawater also have a significant effect on the preservation of organic matter and can prevent its oxidation and destruction in sediments [11,12,13,14,15]. Organic matter enrichment in shales has been explained by two distinct models, i.e., the primary productivity and preservation models [5,6,7,8,9,10,11,12].
The Lower Cambrian Niutitang Formation in the Middle and Upper Yangtze regions consists mainly of bedded siliceous shale, laminated shale, muddy siltstone, and sand-stone [16]. The third-order sequence of the Niutitang Formation is a transgressive systems tract (TST), consisting mainly of siliceous and calcareous shales, and a highstand systems tract (HST), consisting mainly of argillaceous and silty mixed shales. The absence of a lowstand systems tract (LST) can be attributed to rapid transgression at the Ediacaran–Cambrian boundary [17,18]. Organic matter contents vary in the different sedimentary facies and decrease from the deep-water shelf to the shallow-water shelf and tidal flat facies [19]. Owing to negligible fractionation during sediment transport and deposition, geochemical data and proxies obtained from the clastic sediments provide important insights into their palaeoenvironments during deposition [20,21,22,23]. Shields and Stille discussed the effects of post-depositional diagenesis on rare earth elements (REEs) in the basal Cambrian phosphorites of the Meishucun section [24]. Li et al. and Zhou et al. proposed that seawater made a greater contribution to REE contents than hydrothermal inputs in the Niutitang Formation phosphatic, siliceous, and silty mixed shales [25,26]. Several studies have investigated the geochronology and geochemistry of the Niutitang Formation and the processes responsible for organic matter enrichment [27,28,29,30]. However, few studies have investigated the effects of primary productivity and organic matter dilution on organic matter accumulation. In this study, the geochemical compositions of the Niutitang Formation shales were used to determine variations in primary productivity, organic matter dilution and preservation in order to develop a better understanding of organic matter accumulation in organic-rich shales.

2. Samples and Method

The Xiangandi #1 well (XAD 1 well) is located near Anhua, in central Hunan Province (Figure 1). The Niutitang Formation overlies the chert successions of the Liuchapo Formation and contains mainly black organic-rich shales intercalated with gray limestones. The 53 samples from the Niutitang Formation and 2 samples from the Liuchapo Formation were selected from XAD 1 well for major and trace elements analysis (Figure 1c).
Argon ion polished thin sections, taken perpendicular to the bedding orientation, were observed using a Quanta FEG 450 scanning electron microscope. The micrographs were employed to study the morphology of the minerals. Total organic carbon (TOC) content was determined using the carbon/sulfur analyzer. Major element concentrations of SiO2, Al2O3, CaO, TFe2O3, K2O, MgO, MnO, Na2O, P2O5, and TiO2 were determined using an AxiosmAX X-ray fluorescence spectrometer following experimental details of Li et al. [25]. The relative precision of major-element concentrations in the Lower Cambrian Niutitang Formation shales was <±5%. The trace elements concentrations of Mo, Ba, Cu, Ni, Co, V, Cr, and U were determined using an ELEMENT XR plasma mass spectrometer (National Research Center for Geoanalysis, Chinese Academy of Geological Sciences, Beijing, China) following experimental details of Zhou et al. [26], with a relative precision of <±5%. Element enrichment factors (EFelement) were defined as follows: EFelement = (element/Al)sample/(element/Al)average shale, where (element/Al)average shale refers to the data from an average shale [31]. The major elements concentrations, trace elements concentrations, and TOC content of the Niutitang Formation shales are presented in Table 1 and Table 2.

3. Results

3.1. TOC Content

TOC contents range from 2.01% to 19.8% (average = 7.62%) for the Niutitang Formation shales. TOC contents range from 0.13% to 4.05% (average = 2.09%) for the Liuchapo Formation sediments.

3.2. Major Element

SiO2 contents vary from 54.49% to 94.12% (average = 71.33%) for the Niutitang formation shales, and from 76.08% to 98.82% (average = 87.45%) for the Liuchapo Formation sediments (Table 1). Al2O3 contents are 0.8%–13.69% (average = 7.65%) for the Niutitang Formation shales, and 0.37%–9.32% (average Al2O3 = 4.85%) for the Liuchapo Formation sediments. CaO contents vary from 0.16% to 6.75% (average = 1.6%) for the Niutitang Formation shales, and from 0.1% to 0.38% (average = 0.24%) for the Liuchapo Formation sediments. Enrichment factors of major elements indicate that SiO2, TFe2O3, CaO, and P2O5 are enriched in the Niutitang Formation shales, whereas MgO, Na2O, K2O, MnO, and TiO2 are not markedly enriched (Figure 2).

3.3. Trace Element

Concentrations of Mo, Ba, Cu, Ni, Co, V, Cr, U, and Th in the Niutitang Formation shale samples range 4.65437 ppm (mean Mo = 78.82 ppm), 390–40,781 ppm (mean Ba = 5064 ppm), 15.3–1386 ppm (mean Cu = 128.2 ppm), 15.1–949 ppm (mean Ni = 128.9 ppm), 0.92–22.4 ppm (mean Co = 12.72 ppm), 58.4–7585 ppm (mean V = 852.7 ppm), 59.7–871 ppm (mean Cr = 148.4 ppm), 1.97–214 ppm (mean U = 32 ppm), and 0.4–14.9 ppm (mean Th = 7.7 ppm), respectively (Table 2). Mo, Ba, Cu, Ni, Co, V, Cr, U, and Th concentrations of the Liuchapo Formation sediments vary 0.7–0.98 ppm (mean Mo = 0.84 ppm), 265–2198 ppm (mean Ba = 1232 ppm), 2.83–27 ppm (mean Cu = 14.92 ppm), 4.89–45.9 ppm (mean Ni = 25.4 ppm), 0.48–5.31 ppm (mean Co = 2.9 ppm), 12.3–134 ppm (mean V = 73.2 ppm), 108–157 ppm (mean Cr = 132.5 ppm), 0.55–5.14 ppm (mean U = 2.85 ppm), and 0.45–6.9 ppm (mean Th = 3.68 ppm), respectively.

4. Discussion

4.1. Primary Production

Zircon U–Pb geochronology has yielded an age of 536.5 ± 5.5 Ma for the Liuchapo Formation sedimentary rocks [32]. Zircon U–Pb and Re–Os geochronology has yielded ages of 521–514 Ma for the Niutitang Formation sedimentary rocks, corresponding to the Tommotian of the Early Cambrian [27,28]. The Liuchapo and Niutitang formations pre-date the radiation of vascular plants in terrestrial ecosystems, and terrigenous organic C made a negligible contribution to their TOC contents. Although sponge spicule, acantho-morphic acritarch, and scyphozoa fossils are present in the shales, biomarker analysis of the Niutitang Formation shales indicates that planktonic algae were the main primary producers [30]. Relative to average shale [31], Niutitang Formation and Liuchapo Formation sediments have less average Al2O3 contents (average Al2O3 = 6.99%), indicating minor clay minerals. Relative to average shale [31], sedimentary rocks of the Niutitang and Liuchapo formations have lower Al2O3 contents (average = 6.99%), indicative of the presence of only minor clay minerals. The TOC contents of mudstones or shales reflect a relatively small fraction of the primary production due to photosynthesis in the photic zone of the surface ocean [31]. Most organic material generated by primary production sinks from the ocean surface into the thermocline and deep ocean, reaches the sediment–water interface, undergoes decomposition, and is lost [33,34]. Despite this decomposition and diagenesis, TOC contents can still be used to estimate the palaeo-primary productivity [5,22].
TOC contents reach a minimum value in the upper Liuchapo Formation (ULF) and then exhibit a slightly increasing trend (Figure 3). The increasing trend of TOC contents terminates at the marlstones of the base of the Niutitang Formation (BNF), in which TOC contents decrease to 2%. The TOC contents continue to increase in the lower Niutitang Formation (LNF) and subsequently decrease in the middle Niutitang Formation (MNF). The TOC contents remain uniform at 5% until a marked decrease in the marlstones of the upper Niutitang Formation (UNF).
Molybdenum is present as MoO42− in seawater and is rarely incorporated into natural minerals. Organic matter can scavenge Mo from the MoO42− in seawater [35,36], or Mo can be captured as Fe–Mo–S cluster compounds that are present due to H2S [37,38]. Owing to the positive relationships between TOC and Mo contents, Mo provides important insights into primary productivity [39].
Molybdenum exhibits a distinctive increase at the boundary between the Liuchapo and Niutitang formations and then decreases to ~10 ppm in the marlstones in the BNF (Figure 2). Molybdenum then increases (with slight fluctuations) and peaks in the LNF. Molybdenum decreases back to ~45 ppm in the MNF. Subsequently, Mo contents remain uniform, until a distinctive decrease in the marlstones of the UNF. In general, Mo contents correlate with TOC contents throughout the Niutitang Formation (Figure 4a).
Phosphorus is present in dissolved and particulate form in seawater and is a prominent nutrient for microorganisms [40]. Remineralised P after burial is preferentially retained and precipitated in sediments under oxygenated bottom-water conditions but not under O2-depleted conditions, which is attributable to differences in P fixation associated with redox cycling [41]. Although P accumulation mechanisms differ from those of TOC, P is also a proxy for primary productivity [42].
A prominent increase in P contents occurs across the boundary between the ULF and BNF, where P contents increase from 0.012 to 0.445% (Figure 2). Subsequently, P contents remain relatively high throughout the LNF, with slight fluctuations. Contents of P decrease in the overlying MNF, continue to decrease in the grey marlstones in the UNF, and then increase slightly.
Barium is derived mainly from three sources: terrigenous detrital input, hydrothermal fluids, and benthic organisms [43]. The substantial amount of Ba transported from the ocean surface zone to the sediment–water interface along with decaying organic matter is termed biogenic Ba [44]. In order to correct for the presence of detrital Ba, Dymond et al. proposed a method for determining the content of biogenic Ba (Babio) [45]:
Babio = Basamlpe − Alsample × (Ba/Al) dete
where (Ba/Al)detr is the detrital Ba fraction estimated from the upper continental crust.
Both Ba and Babio exhibit similar trends in the Niutitang Formation, and the contents of Ba and Babio are almost the same (Figure 2). A prominent increase in Ba and Babio contents occurs across the boundary between the ULF and BNF. Subsequently, Ba and Babio contents remain relatively high throughout the LNF and then decrease. Barium and Babio contents continue to decrease in the interval of grey marlstones in the UNF. However, the different Ba and Babio contents in the interval of grey marlstones in the UNF may be indicative of a prominent contribution from detrital Ba.
In summary, multi-proxy data for the primary productivity in the LNF are higher than those of the overlying grey–black shales in the MNF and UNF. The primary productivity increased throughout the LNF and subsequently decreased in the MNF and UNF. In addition, the primary productivity reached a minimum during deposition of the grey marlstone intervals at the base and top of the Niutitang Formation.

4.2. Clastic Influx

Most of Ti occurs in clay minerals in sediments, and Ti normalized to Al (Ti/Al ratios) represents the detrital non-aluminosilicate input [23,46]. The Ti/Al ratios decrease to minimum values in the marlstones in the BNF. Subsequently, the Ti/Al ratios have relatively high values (with slight fluctuations) in the LNF (Figure 5), decrease slightly in the MNF, and increase in the UNF. The Ti/Al ratios return to relatively low values in the marlstone intervals of the UNF before increasing again.
The Ti/Al ratios correlate with the TOC contents, but with some notable exceptions in the marlstone intervals of the BNF and UNF, where high Ti/Al ratios correspond to low TOC contents. This might be explained by the changes in Si/Al ratios and CaCO3 contents. The Si/Al ratios reach maximum values with prominent fluctuations in the LNF and subsequently decrease in the MNF and UNF (Figure 5). Where the Si/Al ratios begin to increase in the BNF, the Ti/Al ratios exhibit a small increase and then decrease to low ratios. The low Ti/Al ratios in the marlstone intervals of the BNF and UNF are attributable to dilution by carbonate and correspond to increases in CaCO3 contents.
Silica is derived mainly from terrigenous detrital and biogenic silica inputs. Previous studies have proposed that excess Si contents in the BNF and LNF are attributable to the generation of biogenic silica [25]. Despite the effects of biogenic silica, the higher Si/Al ratios and minimum Ti/Al ratios in the LNF are thought to represent relatively high sea level [5]. With a rise in relative sea level, terrestrial clastic input could then weaken or cut off, resulting in the production of carbonate.
Total organic C contents increase with the sedimentation rate until a threshold value, after which the sedimentation rate dilutes the organic matter accumulation [9]. Clastic inputs increase the burial rate and prevent organic matter degradation [47] or provide more particle sites for organic matter adsorption [48]. Accordingly, Ti/Al ratios correlate well with TOC contents throughout the Niutitang Formation, apart from where dilution occurred during deposition of the marlstone intervals.

4.3. Redox Conditions

Cobalt and Cr are soluble cations in oxygenated seawater, whereas Co and Cr are transported into sediments as authigenic sulphides under anoxic conditions [49]. Because authigenic sulphides of Co and Cr are scarce, detrital/terrigenous inputs contribute most Co and Cr in sediments [28]. Vanadium and Ni are precipitated from seawater under O2 deficient conditions, which are associated with organic matter decay [4]. Therefore, Ni/Co and V/Cr ratios are proxies of redox conditions (Figure 4b,c). Ratios of Ni/Co and V/Cr both increase (with fluctuations) across the ULP–BNF boundary and through the LNF and then decrease in the MNF and UNF (Figure 6). Relative to the MNF and UNF, the higher Ni/Co and V/Cr ratios in the LNF are indicative of an O2-depleted depositional environment.
Uranium occurs as dissolved U6+ in oxygenated seawater, and dissolved U6+ is reduced to insoluble U4+ under O2-depleted conditions, resulting in U enrichment in pelagic and hemipelagic sediments [50]. Algeo and Maynard proposed that organic matter accelerated the scavenging rate of U in sediments [49]. Ratios of U/Al are 6–137 × 10−4 in the LNF and 0.7–6.9 × 10−4 in the MNF and UNF. Ratios of U/Al increase across the Liuchapo–Niutitang formation boundary and remain relatively high (with fluctuations) throughout the LNF. Subsequently, the U/Al ratios exhibit a decrease in the MBF and UNF (Figure 6). The higher U/Al ratios of the LNF indicate the bottom waters were anoxic.
Thorium occurs as insoluble Th4+ in seawater and is unaffected by the redox conditions [51]. Wignall and Twitchett proposed that Th/U ratios generally increased with increasing oxygenation from 02 for anoxic conditions to 2–8 for oxic conditions [52]. Ratios of Th/U decrease across the Liuchapo–Niutitang formation boundary, remain relatively low with fluctuations in the LNF, and subsequently increase in the MNF and UNF (Figure 6). The lower Th/U ratios in the LNF relative to the MNF and UNF are indicative of anoxic bottom waters.
In order to discriminate euxinic from anoxic depositional conditions, Mo was used to recognise redox changes during deposition of the Niutitang Formation. Under euxinic conditions, Mo is precipitated preferentially as Fe sulphides owing to the presence of free H2S [49]. If only U and V are scavenged and enriched without Mo, free H2S did not exist in the seawater; however, if U, V, and Mo exhibit concurrent enrichments, then euxinic conditions existed with free H2S [22]. Uranium, V, and Mo are all enriched in the LNF, indicating that euxinic conditions existed during deposition of the LNF.

4.4. Organic Matter Accumulation Mechanism of Organic-Rich Shales

The redox and productivity proxies suggest that the black, organic-rich shales of the LNF were deposited under anoxic (or even euxinic) conditions during periods of high primary productivity. The grey–black, organic-poor shales of the MNF and UNF were deposited under suboxic conditions during periods of relatively low primary productivity.
Euhedral pyrite is either precipitated directly or formed indirectly by framboidal precursor regrowth, indicative of limited FeS2 saturation and intense late diagenesis [53]. Framboidal pyrite in marine sediments has two main mechanisms of formation: (1) under oxic–dysoxic bottom-water conditions, in which diagenetic, framboidal pyrite grows in anoxic pore waters in sediments [54]; and (2) under euxinic bottom-water conditions, in which syngenetic framboidal pyrite grows in anoxic bottom waters and sinks into the sediments [55]. Framboidal pyrite dominates over euhedral pyrite in the Niutitang Formation sediments (Figure 7), indicating the framboidal pyrite could be an indicator of the redox environment. The mean diameter and size distribution of framboidal pyrite in the Niutitang Formation sediments suggest the pyrite formed under euxinic conditions [55]. Based on organic biomarker analysis, the main primary producers were planktonic algae [35]. Planktonic algae might have produced mineralised tests that sank into the sediments.
Owing to tectonic subsidence in the Huanan rift and a relative rise in sea level [22], the accommodation space increased during deposition of the LNF. Because of the rapid transgression at the Ediacaran–Cambrian boundary, the Niutitang Formation is divided into transgressive and highstand systems tracts [21]. During deposition of the LNF, the relative sea level rose to the maximum flooding surface and the seasonal mixing could not affect the bottom waters [25,26]. Biomarker analyses show that nutrients were derived from planktonic algae [30], and upwelling provided nutrients for primary producers based on the occurrence of phosphorites and P enrichment. Productivity proxies (Mo and Babio) show that primary productivity was higher in the ocean surface zone during deposition of the LNF shales. The high primary productivity enhanced the organic C flux into the chemocline layer and bottom waters, leading to the development of anoxic bottom waters owing to O2 consumption by microorganisms and organic matter degradation. In addition, clastic inputs increased the burial rate and prevented organic matter degradation. The anoxic bottom waters were beneficial for the preservation of organic matter in the LNF shales (Figure 8).
During deposition of the grey–black MNF and UNF shales, relative sea level fell and primary productivity decreased. Seasonal mixing may have reached the bottom waters and, in addition, low primary productivity would not have formed anoxic bottom waters. The low primary productivity and suboxic bottom waters would have resulted in the relatively organic-poor MNF and UNF shales.

5. Conclusions

  • Redox and primary productivity proxies suggest that the black, organic-rich shales in the Niutitang Formation were deposited in anoxic–euxinic conditions during periods of high primary productivity;
  • Framboidal pyrite indicates that anoxic bottom waters existed during organic matter deposition. The high primary productivity enhanced the organic C flux into the thermocline layer and bottom waters, which formed the O2-depleted bottom waters;
  • Ti/Al ratios correlate with TOC contents throughout the Niutitang Formation, indicating the clastic input enhanced the burial rate and prevented organic matter degradation during deposition of the Niutitang Formation.

Author Contributions

Conceptualization, L.Z. (Lei Zhou); methodology, L.Z. (Linyan Zhang); software, L.Z. (Linyan Zhang); validation, L.W.; formal analysis, L.Z. (Linyan Zhang); investigation, L.Z. (Lei Zhou); resources, X.F. and L.W.; data curation, L.Z. (Linyan Zhang); writing—original draft, L.Z. (Lei Zhou); writing—review and editing, L.Z. (Lei Zhou); visualization, L.W.; supervision, X.F.; project administration, X.F.; funding acquisition, X.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (42372179), the China Postdoctoral Science Foundation (2018M631541), and the China Geological Survey (DD20221660).

Data Availability Statement

Data are contained within the article.

Acknowledgments

We thank the Institute of Geomechanics, Chinese Academy of Geological Sciences for the provision of geological data and borehole samples. The careful review and constructive suggestions of the manuscript by anonymous reviewers are greatly appreciated.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bohacs, K.M.; Grawbowski, G.J.; Carroll, A.R.; Mankeiwitz, P.J.; Miskell-Gerhardt, K.J.; Schwalbach, J.R.; Wegner, M.B.; Simo, J.A. Production, and dilution—The many paths to source-rock development. In Deposition of Organic-Carbonrich Sediments: Models, Mechanisms and Consequences; SEPM Special Publication: Tulsa, OK, USA, 2005; pp. 61–101. [Google Scholar]
  2. Passey, Q.R.; Bohacs, K.M.; Esch, W.L.; Klimentidis, R.; Sinha, S. From oil-prone source rock to gas-producing shale reservoir-geologic and petophysical characterization of unconventional shale gas reservoirs. In Proceedings of the SPE 131350, SPE International Oil & Gas Conference and Exhibition, Beijing, China, 8–10 June 2010. [Google Scholar]
  3. Wang, Z.W.; Wang, J.; Fu, X.G.; Zhan, W.Z.; Yu, F.; Feng, X.L.; Song, C.Y.; Chen, W.B.; Zeng, S.Q. Organic material accumulation of Carnian mudstones in the North Qiangtang Depression, eastern Tethys: Controlled by the paleoclimate, paleo- environment, and provenance. Mar. Petrol. Geol. 2017, 88, 440–457. [Google Scholar] [CrossRef]
  4. Calvert, S.E.; Pedersen, T.F. Geochemistry of recent oxic and anoxic marine sediments: Implications for the geological record. Mar. Geol. 1993, 113, 67–88. [Google Scholar] [CrossRef]
  5. Sageman, B.B.; Murphy, A.E.; Werne, J.P.; Ver Straeten, C.A.; Hollander, D.J.; Lyons, T.W. A tale of shales: The relative roles of production, decomposition, and dilution in the accumulation of organic-rich strata, Middle–Upper Devonian, Appalachian basin. Chem. Geol. 2003, 195, 229–273. [Google Scholar] [CrossRef]
  6. Wei, H.Y.; Chen, D.Z.; Wang, J.G.; Yu, H.; Tucker, M.E. Organic accumulation in the lower Chihsia Formation (Middle Permian) of South China: Constraints from pyrite morphology and multiple geochemical proxies. Paleogeoger. Palaeocl. 2012, 353–355, 73–86. [Google Scholar] [CrossRef]
  7. Algeo, T.J.; Henderson, C.M.; Tong, J.N.; Feng, Q.L.; Yin, H.F.; Tyson, R.V. Plankton and productivity during the Permian–Triassic boundary crisis: An analysis of organic carbon fluxes. Glob. Planet. Chang. 2013, 105, 52–67. [Google Scholar] [CrossRef]
  8. Creaney, S.; Passey, Q.R. Recurring patterns of total organic carbon and source rock quality within a sequence stratigraphic framework. AAPG Bull. 1993, 77, 386–401. [Google Scholar]
  9. Tyson, R.V. Sedimentation rate, dilution, preservation and total organic carbon: Some results of a modelling study. Org. Geochem. 2001, 32, 333–339. [Google Scholar] [CrossRef]
  10. Henrichs, S.M.; Reeburgh, W.S. Anaerobic mineralization of marine sediment organic matter: Rates and the role of anaerobic processes in the oceanic carbon economy. Geomicrobiol. J. 1987, 5, 191–238. [Google Scholar] [CrossRef]
  11. Demaison, G.J.; Moore, G.T. Anoxic environments and oil source bed genesis. Org. Geochem. 1980, 2, 9–31. [Google Scholar] [CrossRef]
  12. Calvert, S.E.; Bustin, R.M.; Ingall, E.D. Influence of water column anoxia and sediment supply on the burial and preservation of organic carbon in marine shales. Geochim. Cosmochim. Acta 1996, 60, 1577–1593. [Google Scholar] [CrossRef]
  13. Yan, D.T.; Wang, H.; Fu, Q.L.; Chen, Z.H.; He, J.; Gao, Z. Geochemical characteristics in the Longmaxi Formation (Early Silurian) of South China: Implications for organic matter accumulation. Mar. Petrol. Geol. 2015, 65, 290–301. [Google Scholar] [CrossRef]
  14. Peng, J. What besides redox conditions? Impact of sea-level fluctuations on redoxsensitive trace-element enrichment patterns in marine sediments. Sci. China Earth Sci. 2022, 65, 1985–2004. [Google Scholar] [CrossRef]
  15. Peng, J.W.; Larson, T. A novel integrated approach for chemofacies characterization of organic-rich mudrocks. AAPG Bull. 2022, 106, 437–460. [Google Scholar] [CrossRef]
  16. Liang, C.; Jiang, Z.X.; Guo, L.; Yang, Y.T. Characteristics of black shale, sedimentary evolution and shale gas exploration prospect of shelf face taking Wengan and Yongye profile of the Niutitang Formation as an example. J. Daqing Petrol. Inst. 2011, 35, 13–21. [Google Scholar]
  17. Mei, M.X.; Zhang, H.; Meng, X.Q.; Chen, Y.H. Sequence stratigraphic division and framework of the Lower Cambrian in the Upper Yangte region. Geol. China 2006, 33, 1292–1304. [Google Scholar]
  18. Wu, Y.; Fan, T.L.; Jiang, S.; Yang, X.Q. Lithofacies and sedimentary sequence of the lower Cambrian Niutitang shale in the upper Yangtze platform, South China. J. Nat. Gas Sci. Eng. 2017, 43, 124–136. [Google Scholar] [CrossRef]
  19. Wu, Y.; Fan, T.L.; Ding, H.Y. Lithofacies and Sedimentary Model of the Lower Cambrian Marine Shale in the Upper Yangtze Platform. Geoscience 2017, 31, 1222–1232. [Google Scholar]
  20. Arthur, M.A.; Sageman, B.B. Marine black shales: Depositional mechanisms and environments of ancient deposits. Annu. Rev. Earth Planet Sci. 1994, 22, 499–551. [Google Scholar] [CrossRef]
  21. Jones, B.; Manning, D.A.C. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones. Chem. Geol. 1994, 111, 111–129. [Google Scholar] [CrossRef]
  22. 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]
  23. 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, 200, 1–19. [Google Scholar] [CrossRef]
  24. Shields, G.; Stille, P. Diagenetic constraints on the use of cerium anomalies as palaeoseawater redox proxies: An isotopic and REE study of Cambrian phosphorites. Chem. Geol. 2001, 175, 29–48. [Google Scholar] [CrossRef]
  25. Li, Y.F.; Fan, T.L.; Zhang, J.C.; Zhang, J.P.; Wei, X.J.; Hu, X.L.; Zeng, W.T.; Fu, W. Geochemical changes in the Early Cambrian interval of the Yangtze Platform, South China: Implications for hydrothermal influences and paleocean redox conditions. J. Asian Earth Sci. 2015, 109, 100–123. [Google Scholar] [CrossRef]
  26. Zhou, L.; Kang, Z.H.; Wang, Z.X.; Peng, Y.Y.; Xiao, H.F. Sedimentary geochemical investigation for paleoenvironment of the Lower Cambrian Niutitang Formation shales in the Yangtze Platform. J. Petrol. Sci. Eng. 2017, 159, 376–386. [Google Scholar] [CrossRef]
  27. Mao, J.W.; Lehmann, B.; Du, A.D.; Zhang, G.D.; Ma, D.S.; Wang, Y.T.; Zeng, M.G.; Kerrich, R. Re-Os dating of polymetallic Ni-Mo-PGE-Au mineralization in Lower Cambrian black shales of South China and its geologic significance. Econ. Geol. 2002, 97, 1051–1061. [Google Scholar] [CrossRef]
  28. Jiang, S.Y.; Yang, J.H.; Ling, H.F.; Chen, Y.Q. Extreme enrichment of polymetallic Ni–Mo–PGE–Au in lower Cambrian black shales of South China: An Os isotope and PGE geochemical investigation. Paleogeogr. Paleocl. 2007, 254, 217–228. [Google Scholar] [CrossRef]
  29. Xu, L.G.; Lehmann, B.; Mao, J.W. Seawater contribution to polymetallic Ni–Mo–PGE–Au mineralization in Early Cambrian black shales of South China: Evidence from Mo isotope, PGE, trace element, and REE geochemistry. Ore Geol. Rev. 2013, 52, 66–84. [Google Scholar] [CrossRef]
  30. Liu, Z.H.; Zhuang, X.G.; Teng, G.E.; Xie, X.M.; Yin, L.M.; Bian, L.Z.; Feng, Q.L.; Algeo, T.J. The Lower Cambrian Niutitang Formation at Yangtiao (Guizhou, SW China): Organic matter enrichment, source rock potential, and hydrothermal influences. J. Petrol. Geol. 2015, 38, 411–432. [Google Scholar] [CrossRef]
  31. Wedepohl, K.H. Environmental influences on the chemical composition of shales and clays. In Physics and Chemistry of the Earth; Ahrens, L.H., Press, F., Runcorn, S.K., Urey, H.C., Eds.; Pergamon: Oxford, UK, 1971; Volume 8, pp. 307–331. [Google Scholar]
  32. Chen, D.; Wang, J.; Qing, H.; Yan, D.; Li, R. Hydrothermal venting activities in the Early Cambrian, south China: Petrological, geochronological and stable isotopic constraints. Chem. Geol. 2009, 258, 168–181. [Google Scholar] [CrossRef]
  33. Schoepfer, S.D.; Shen, J.; Wei, H.Y.; Tyson, R.V.; Ingall, E.; Algeo, T.J. Total organic carbon, organic phosphorus, and biogenic barium fluxes as proxies for paleomarine productivity. Earth-Sci. Rev. 2014, 149, 23–52. [Google Scholar] [CrossRef]
  34. Hedges, J.I.; Keil, R.G. Sedimentary organic matter preservation: An assessment and speculative synthesis. Mar. Chem. 1995, 49, 81–115. [Google Scholar] [CrossRef]
  35. Helz, G.R.; Miller, C.V.; Charnock, J.M.; Mosselmans, J.L.W.; Pattrick, R.A.D.; Garner, C.D.; Vaughan, D.J. Mechanisms of molybdenum removal from the sea and its concentration in black shales: EXAFS evidences. Geochim. Cosmochim. Acta 1996, 60, 3631–3642. [Google Scholar] [CrossRef]
  36. Adelson, J.M.; Helz, G.R.; Miller, C.V. Reconstructing the rise of recent coastal anoxia; molybdenum in Chesapeake Bay sediments. Geochim. Cosmochim. Acta 2001, 65, 237–252. [Google Scholar] [CrossRef]
  37. Bertine, K.K. The deposition of molybdenum in anoxic waters. Mar. Chem. 1972, 1, 43–53. [Google Scholar] [CrossRef]
  38. Ripley, E.M.; Shaffer, N.R.; Gilstrap, M.S. Distribution and geochemical characteristics of metal enrichment in the New Albany Shale (Devonian–Mississippian), Indiana. Econ. Geol. 1990, 85, 1790–1807. [Google Scholar] [CrossRef]
  39. Werne, J.P.; Sageman, B.B.; Lyons, T.W.; Hollander, D.J. An integrated assessment of a “type euxinic” deposit: Evidence for multiple controls on black shale deposition in the Middle Devonian Oatka Creek Formation. Am. J. Sci. 2002, 302, 110–143. [Google Scholar] [CrossRef]
  40. Paytan, A.; McLaughlin, K. The oceanic phosphorus cycle. Chem. Rev. 2007, 107, 563–576. [Google Scholar] [CrossRef] [PubMed]
  41. Ingall, E.D.; Kolowith, L.; Lyons, T.; Hurtgen, M. Sediment carbon, nitrogen and phosphorus cycling in an anoxic fjord, Effingham Inlet, British Columbia. Am. J. Sci. 2005, 305, 240–258. [Google Scholar] [CrossRef]
  42. Ingall, E.R.; Bustin, M.; Cappellen, P.V. Influence of water column anoxia on the burial and preservation of carbon and phosphorus in marine shales. Geochim. Cosmochim. Acta 1993, 57, 303–316. [Google Scholar] [CrossRef]
  43. Bertram, M.A.; Cowen, J.P. Morphological and compositional evidence for biotic precipitation of marine barite. J. Mar. Res. 1997, 55, 577–593. [Google Scholar] [CrossRef]
  44. Ganeshram, R.S.; François, R.; Commeau, J.; Brown-Leger, S.L. An experimental investigation of barite formation in seawater. Geochim. Cosmochim. Acta 2003, 67, 2599–2605. [Google Scholar] [CrossRef]
  45. Dymond, J.; Suess, E.; Lyle, M. Barium in deep-sea sediment: A geochemical proxy for paleoproductivity. Paleoceanography 1992, 7, 163–181. [Google Scholar] [CrossRef]
  46. Bertrand, P.; Shimmield, G.; Martinez, P.; Grousset, F.; Jorissen, F.; Paterne, M.; Pujol, C.; Bouloubassi, I.; Buat Menard, P.; Peypouquet, J.P.; et al. The glacial ocean productivity hypothesis: The importance of regional temporal and spatial studies. Mar. Geol. 1996, 130, 1–9. [Google Scholar] [CrossRef]
  47. Ibach, L.E.J. Relationship between sedimentation rate and total organic carbon content in ancient marine sediments. AAPG Bull. 1982, 66, 170–188. [Google Scholar]
  48. Zeng, S.Q.; Wang, J.; Fu, X.G.; Chen, W.B.; Feng, X.L.; Wang, D.; Song, C.Y.; Wang, Z.W. Geochemical characteristics, redox conditions, and organic matter accumulation of marine oil shale from the Changliang Mountain area, northern Tibet, China. Mar. Petrol. Geol. 2015, 64, 203–221. [Google Scholar] [CrossRef]
  49. 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]
  50. McManus, J.; Berelson, W.M.; Klinkhammer, G.P.; Hammond, D.E.; Holm, C. Authigenic uranium: Relationship to oxygen penetration depth and organic carbon rain. Geochim. Cosmochim. Acta 2005, 69, 95–108. [Google Scholar] [CrossRef]
  51. Kimura, H.; Watanabe, Y. Oceanic anoxia at the Precambrian-Cambrian boundary. Geology 2001, 29, 995–998. [Google Scholar] [CrossRef]
  52. Wignall, P.B.; Twitchett, R.J. Oceanic anoxia and the end Permian mass extinction. Science 1996, 272, 1155–1158. [Google Scholar] [CrossRef]
  53. Khan, D.; Qiu, L.W.; Liang, C.; Mirza, K.; Rehman, S.U.; Han, Y.; Hannan, A.; Kashif, M.; Kra, K.L. Genesis and distribution of pyrite in the lacustrine shale: Evidence from the Es3x shale of the Eocene Shahejie Formation, Zhanhua Sag, East China. ACS Omega 2022, 7, 1244–1258. [Google Scholar] [CrossRef]
  54. Zhao, J.; Liang, J.L.; Long, X.P.; Li, J.; Xiang, Q.R.; Zhang, J.C.; Hao, J.L. Genesis and evolution of framboidal pyrite and its implications for the oreforming process of Carlin-style gold deposits, southwestern China. Ore Geol. Rev. 2018, 102, 426–436. [Google Scholar] [CrossRef]
  55. Chang, J.Y.; Li, Y.Y.; Lu, H.L. The morphological characteristics of authigenic pyrite formed in marine sediments. J. Mar. Sci. Eng. 2022, 10, 1533. [Google Scholar] [CrossRef]
Figure 1. Simplified geological map in the southeast Yangtze Platform margin and lithological profile of XAD 1 well section. (a) Location map of the studied area in China, (b) Simplified geological map of Anhua County, Hunan Province, (c) lithological profile and sample sits of XAD 1 well.
Figure 1. Simplified geological map in the southeast Yangtze Platform margin and lithological profile of XAD 1 well section. (a) Location map of the studied area in China, (b) Simplified geological map of Anhua County, Hunan Province, (c) lithological profile and sample sits of XAD 1 well.
Minerals 14 00183 g001
Figure 2. Enrichment factors (EF) of major elements in the Niutitang Formation shales relative to average shale.
Figure 2. Enrichment factors (EF) of major elements in the Niutitang Formation shales relative to average shale.
Minerals 14 00183 g002
Figure 3. The primary production proxies (TOC, Mo, P, Ba, and Babio) in the Niutitang Formation.
Figure 3. The primary production proxies (TOC, Mo, P, Ba, and Babio) in the Niutitang Formation.
Minerals 14 00183 g003
Figure 4. Crossplots of Mo, Ni/Co, and V/Cr with TOC for the Niutitang and Liuchapo Formation. (a) Mo and TOC, (b) Ni/Co and TOC, (c) V/Cr and TOC.
Figure 4. Crossplots of Mo, Ni/Co, and V/Cr with TOC for the Niutitang and Liuchapo Formation. (a) Mo and TOC, (b) Ni/Co and TOC, (c) V/Cr and TOC.
Minerals 14 00183 g004
Figure 5. The dilution proxies (CaCO3, Si/Al, and Ti/Al) in the Niutitang Formation.
Figure 5. The dilution proxies (CaCO3, Si/Al, and Ti/Al) in the Niutitang Formation.
Minerals 14 00183 g005
Figure 6. The redox condition proxies (Ni/Co, V/Cr, U/Al, and Th/U) in the Niutitang Formation.
Figure 6. The redox condition proxies (Ni/Co, V/Cr, U/Al, and Th/U) in the Niutitang Formation.
Minerals 14 00183 g006
Figure 7. Pyrite in the organic-rich Niutitang Formation.
Figure 7. Pyrite in the organic-rich Niutitang Formation.
Minerals 14 00183 g007
Figure 8. Conceptual model showing the organic matter enrichment mechanisms of the Niutitang Formation shales.
Figure 8. Conceptual model showing the organic matter enrichment mechanisms of the Niutitang Formation shales.
Minerals 14 00183 g008
Table 1. The major element concentrations (%) of the Niutitang Formation shales.
Table 1. The major element concentrations (%) of the Niutitang Formation shales.
Samples
ID
Depth
/m
SiO2Al2O3TFe2O3MgOCaONa2OK2OMnOTiO2P2O5
Sam.01675.361.958.089.722.263.700.462.690.020.390.052
Sam.02677.168.5311.353.432.011.300.683.960.010.590.066
Sam.03681.274.417.913.442.321.400.572.560.010.380.051
Sam.0468676.148.553.031.480.560.392.95<0.010.420.068
Sam.0568966.9711.284.661.552.410.603.860.020.540.094
Sam.06692.871.899.583.971.410.810.393.33<0.010.470.062
Sam.07695.967.607.722.932.474.680.072.630.010.400.07
Sam.08698.973.768.623.731.210.160.052.89<0.010.440.056
Sam.09701.263.1213.695.542.520.360.074.330.010.710.10
Sam.10709.85.410.470.733.0248.340.130.070.030.030.036
Sam.11712.373.0910.713.691.410.160.623.06<0.010.470.074
Sam.12715.66.320.791.152.6147.570.160.090.040.040.044
Sam.13718.55.620.720.812.4948.52<0.050.100.030.040.035
Sam.14722.878.316.622.301.092.090.141.97<0.010.310.06
Sam.15726.666.45.786.094.724.580.060.950.030.220.082
Sam.16729.670.6211.53.721.590.720.063.52<0.010.370.063
Sam.17732.954.499.745.989.185.350.061.470.060.390.071
Sam.1873772.049.793.161.240.890.112.890.020.350.093
Sam.19740.572.229.44410.270.182.73<0.010.360.074
Sam.20743.473.379.872.961.170.430.112.97<0.010.380.069
Sam.21745.471.329.294.421.061.620.362.90.020.380.073
Sam.22750.669.311.154.131.371.450.133.320.020.460.082
Sam.23754.159.1410.725.71.286.750.383.020.050.390.081
Sam.24757.273.9110.432.791.140.410.163.04<0.010.390.057
Sam.2576074.7910.42.841.260.80.083.120.020.350.079
Sam.26763.671.4411.413.111.290.360.23.250.010.330.086
Sam.27767.667.1110.545.261.321.340.183.140.020.380.105
Sam.2877265.6611.355.322.382.620.113.290.040.370.101
Sam.29775.674.798.223.011.210.590.062.440.010.340.102
Sam.30778.678.737.042.010.890.830.072.040.020.290.094
Sam.31782.478.25.852.61.250.45<0.051.62<0.010.290.061
Sam.32786.464.4210.714.531.990.880.052.850.020.540.148
Sam.33789.372.44.672.131.444.26<0.051.060.030.220.158
Sam.34793.473.45.823.711.330.8<0.051.550.010.290.108
Sam.35797.678.15.082.660.770.72<0.051.41<0.010.20.147
Sam.3680059.98.014.891.260.29<0.052.17<0.010.350.153
Sam.3780458.865.844.532.640.730.081.70.010.340.466
Sam.38807.864.986.594.860.630.710.232.11<0.010.350.297
Sam.39810.562.297.034.221.631.440.342.130.010.420.221
Sam.40813.377.562.941.151.162.180.180.740.010.140.076
Sam.41819.3569.745.061.751.870.980.11.260.010.270.15
Sam.42822.683.662.381.040.350.340.10.65<0.010.110.166
Sam.43826.571.416.982.250.941.020.212.070.010.380.316
Sam.44829.779.772.691.180.831.840.060.720.010.140.138
Sam.45836.862.847.114.791.051.010.072.12<0.010.370.563
Sam.46840.676.591.590.910.782.15<0.050.290.010.070.344
Sam.47843.465.947.293.610.740.740.072.04<0.010.390.30
Sam.48847.535.344.993.3616.9316.36<0.050.760.080.210.10
Sam.4985294.120.80.440.080.85<0.050.05<0.010.010.25
Sam.50854.384.890.890.90.54.38<0.050.16<0.010.060.082
Sam.5185974.564.32.251.231.13<0.051.42<0.010.190.174
Sam.52862.525.240.90.182.6530.03<0.050.170.040.020.445
Sam.53865.579.014.772.822.673.10.161.460.030.270.035
Sam.54870.876.089.323.822.060.38<0.052.96<0.010.540.022
Sam.55882.398.820.370.30.080.1<0.050.13<0.010.020.012
Table 2. The trace element concentrations (ppm) and TOC contents (%) of the investigated shales.
Table 2. The trace element concentrations (ppm) and TOC contents (%) of the investigated shales.
Samples
ID
Depth
/m
TOCMoBaCuNiCoVCrUTh
Sam.01675.36.5735.810331178017.111074.810.49.27
Sam.02677.14.8549.5125764.179.716.318284.99.113.5
Sam.03681.24.3329.4103149.34814.386.867.811.59.89
Sam.046864.4230.5116652.849.213.992.676.514.210.6
Sam.056894.3936.4139473.45320.410077.311.712.4
Sam.06692.86.187.813079881.316.517872.419.711.2
Sam.07695.94.5245.597973.97510.915782.610.58.69
Sam.08698.95.6880.291911673.81812177.219.410.8
Sam.09701.26.5857.9108196.454.222.412995.119.514.9
Sam.10709.80.480.441752.69.081.35.718.050.580.56
Sam.11712.34.2848.8126610679.115.813374.11411.2
Sam.12715.60.842.271758.979.021.9512.510.41.30.82
Sam.13718.50.690.652133.558.51.6611.17.841.150.65
Sam.14722.83.5432.384311374.49.9718686.99.497.34
Sam.15726.65.415.139072.426.28.2461.986.29.795.43
Sam.16729.64.0633.9140772.890.616.2221719.3812
Sam.17732.94.6419.866266.740.817.278.578.47.6410.7
Sam.187375.4340133371.653.514.395.675.410.210.8
Sam.19740.56.0128.4144955.268.313.410467.87.219.91
Sam.20743.45.1933154770.843.413.682.866.77.7610.1
Sam.21745.45.7930.6141364.343.114.282.367.87.829.53
Sam.22750.65.3350.9163381.458.317.511975.611.911.3
Sam.23754.16.2249.8143669.482.114.915470.98.889.91
Sam.24757.24.4433.8173756.495.313.539171.59.019.92
Sam.257604.1826.4169050.578.713.822673.66.8510.5
Sam.26763.64.5962.7198765.912315.357187.911.910.3
Sam.27767.67.4863.5183069.999.313.851289.113.611.1
Sam.287725.1436.6168584.712217.534498.411.39.53
Sam.29775.66.7869.116335292.212.91857531.58.88
Sam.30778.65.929.6212135.958.49.6415062.215.96.56
Sam.31782.46.149.715086082.410.314970.522.66.9
Sam.32786.410.1110252410888.519.125388.454.912.5
Sam.33789.38.6388.4114955.342.98.513577.132.35.36
Sam.34793.49.7445.6227888.565.517.216159.735.97.05
Sam.35797.68.4107251149627810.1247961718.34.96
Sam.3680018.24270347415561717.64500480597.5
Sam.3780419.8943711,80174.249317.6127278.22145.26
Sam.38807.814.6324914,22217414712.41635127794.81
Sam.39810.512.6118517,78956.218111.8114884.679.64.72
Sam.40813.39.1334.1904130.366.33.7113010287.52.22
Sam.41819.359.9291.915,63113861556.59253817238.62.92
Sam.42822.61032.6773115.367.24.1642510149.71.75
Sam.43826.58.2715117,05511117312.1184119356.44.68
Sam.44829.79.2598.9903034.11514.760513982.62.04
Sam.45836.813.0917216,87646751614.2758551740.95.01
Sam.46840.612.9222.8893817271.83.62348318219.31.11
Sam.47843.412.4215440,78112522713.3334918943.63.96
Sam.48847.52.110.827,47535.429.59.736521413.522.24
Sam.4985213.126.43994126829.20.92121526529.60.4
Sam.50854.34.761826632169513.2312344451181.33
Sam.5185910.88144358493.494.914.6688087142.53.43
Sam.52862.54.5433.956,89397.449.44.745391339.150.26
Sam.53865.52.014.65433116.115.13.458.41751.975.5
Sam.54870.84.050.9821982745.95.311341085.146.9
Sam.55882.30.130.72652.834.890.4812.31570.550.45
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

Zhou, L.; Feng, X.; Zhang, L.; Wu, L. Organic Matter Enrichment Mechanisms in the Lower Cambrian Shale: A Case Study from Xiangandi #1 Well. Minerals 2024, 14, 183. https://doi.org/10.3390/min14020183

AMA Style

Zhou L, Feng X, Zhang L, Wu L. Organic Matter Enrichment Mechanisms in the Lower Cambrian Shale: A Case Study from Xiangandi #1 Well. Minerals. 2024; 14(2):183. https://doi.org/10.3390/min14020183

Chicago/Turabian Style

Zhou, Lei, Xingqiang Feng, Linyan Zhang, and Lin Wu. 2024. "Organic Matter Enrichment Mechanisms in the Lower Cambrian Shale: A Case Study from Xiangandi #1 Well" Minerals 14, no. 2: 183. https://doi.org/10.3390/min14020183

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