Next Article in Journal
Cretaceous Changes of Strike-Slip Tectonics on the North Pacific Margins: Implications for the Earth’s Rotation
Previous Article in Journal
Petrophysical Acoustic Characterization of Weathered Granite: A Case Study of Mesozoic Granites in the Coastal Area of Fujian Province, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalysis of Minerals in Pyrolysis Experiments

1
School of Energy Resources, China University of Geosciences (Beijing), Beijing 100089, China
2
College of Earth Sciences, Yangtze University, Wuhan 430100, China
3
Department of Geoscience, University of Calgary, 2500 University Drive NW, Calgary, AB T2N 1N4, Canada
4
Changqing Industry Group Co., Ltd., China Petroleum Changqing Oilfield Branch, Xi’an 710021, China
*
Author to whom correspondence should be addressed.
Submission received: 1 March 2023 / Revised: 4 April 2023 / Accepted: 5 April 2023 / Published: 6 April 2023

Abstract

:
Recent research in the field of oil and gas geochemistry has focused on the catalytic role of minerals in geological history. Thermal simulation experiments are considered a valuable means of studying the formation and transformation of hydrocarbons. In this paper, we review the catalytic mechanisms, processes, and various arguments for different types of minerals in thermal simulation experiments from the perspective of mineral additives. We focus on two categories: (1) minerals that provide direct catalysis, such as clay minerals, alkali metals, carbonate rocks, and some transition metal elements, and (2) minerals, such as serpentine, that promote aqueous hydrogen and act as the material basis, as well as the radioactive element uranium. We also discuss existing disputes and prospects for the development direction of mineral catalytic thermal simulation experiments.

1. Introduction

Since the discovery that oil and natural gas mainly derive from kerogen, organic geochemistry has investigated the composition, types, hydrocarbon (gas) generation law, and sedimentary environment of kerogen using pyrolysis, chromatography, and other methods [1,2,3]. The aim of this study is to discuss the types and transformation of organic matter [4], kerogen oil generation [5], gas from crude oil cracking [6], and the safe utilization of coal and oil shale [7]. To date, three types of pyrolysis experiments exist, namely closed, semi-open, and open pyrolysis methods. The gold tube closed-system pyrolysis experiment is the most widely used. Gold’s good ductility and stability enables the pyrolysis experiment to consider higher temperatures and pressure environments and thus more geological conditions. The open pyrolysis test uses different gas atmospheres (such as nitrogen) to purge the pyrolysis products into the analytical instrument, which has the advantages of a simple experiment and high analysis efficiency. However, the process of hydrocarbon formation and transport is not carried out in such an open-system experiment. So open systems are limited in what they can simulate. In the semi-open system, a device for discharging pyrolysis products is added on the basis of the closed system. Due to the limitation of equipment and technology, the internal pressure of the hydrocarbon expulsion process will change and affect the liquid hydrocarbon expulsion. In the design of pyrolysis experiments, the controlled factors of experimental purposes and samples should be fully considered to avoid systematic errors of experimental results caused by different pyrolysis systems [8].
The simulation method has gradually started to consider the pore throat characteristics of different types of compounds in various reservoirs and the features of hydrocarbon generation and expulsion under pressure [9,10]. Coal, oil, kerogen, carbonaceous mudstone, carbonate rock, and other organic matters can be the subjects of simulation [11,12,13]. Please refer to the review article by Peng et al. [14] for more information on the status of thermal simulation experiments. However, the geological history of crude oil and natural gas formation is much more complicated than previously thought, involving different surface interface behaviors and organic–inorganic interactions. Therefore, when designing experiments, geological factors, the selection of experimental equipment, and whether to add different minerals or water should be considered [15].
As far as current reports are concerned, different minerals provide varying catalytic effects in the process of thermal simulation experiments. The mechanism by which silicate clay minerals and carbonate minerals promote pyrolysis is still unclear, and the specific behavior of kerogen catalysis by some ions and minerals in geological processes is still controversial. Examples include coordinated metal ions in biomacromolecules from sedimentary environments, uranium captured during the formation of crude oil and natural gas, carbonate minerals such as calcite and dolomite, and potentially strongly catalytic metal ions. The catalytic mechanism of these metal ions during pyrolysis has not been thoroughly studied. This paper reviews the process of promoting kerogen to produce crude oil or natural gas from the perspectives of clay minerals, alkali metals, carbonate rocks, other metal minerals (mainly transition metals), uranium as thermal simulation additives, and serpentine olivine. This paper is mainly divided into two parts, one of which discusses how minerals directly play a catalytic role, while the other discusses how minerals and water coexist to generate hydrogen. The clear catalytic methods are summarized, while the controversial aspects are also reviewed.

2. Minerals Directly Promote Thermal Simulation Process

In modern crude oil and kerogen samples, a variety of minerals can be found. Prior to pyrolysis experiments, it is common practice to pretreat rocks with acidic fluids such as HCl and HF, which removes all minerals except kerogen. These minerals are influenced by the ancient environment during the earth’s redox environment changes, and their unique characteristics can be used to infer sedimentary location environments, tectonic histories, and whether they have undergone epigenesis.
Moreover, these minerals can catalyze the formation of hydrocarbons in various forms. Typically, silico-aluminate dominated clay minerals promote the formation of petroleum in the form of unique solid acids, while various metallic elements enriched in other sedimentary environments are oxidation–reduction catalysts. In addition, contact between these minerals and liquids can result in secondary changes that promote the pyrolysis process.

2.1. Effect of Clay Minerals on Pyrolysis

The formation and structure of clay minerals in reservoirs are influenced by ancient water flow, environment, and provenance debris during the sedimentary period. Therefore, sedimentologists consider clay minerals as good indicators to reflect changes in sedimentary environments. Clay minerals are often used to divide sequences, construct and restore paleoenvironments, and judge major geo-historical events [16]. These silicate layered minerals contain metal elements such as potassium and aluminum, and their particle size is usually less than 3.9 μm. Researchers in the petroleum industry are highly interested in clay minerals [17].
The common types of clay minerals found in sedimentary environments are illite, montmorillonite, kaolinite, chlorite, and their mixed layers. Although clay minerals are not welcomed by reservoir engineers due to their special physical and chemical changes [18], such as water absorption and swelling, dispersed migration, and the release of metal ions including Fe and Al, they have gained attention in recent years among researchers exploring and developing unconventional reservoirs such as tight sandstone and shale. The nano-scale pores provided by clay minerals are excellent for preserving organic matter, and researchers are investigating this aspect [19].
Clay minerals were first used in the petrochemical field in the mid-twentieth century, and montmorillonite clay was first used as the cracking catalyst for crude oil in the chemical industry [20]. Since then, geologists and geochemists have been interested in the promotion of various mineral types in crude oil production. For instance, Eltantawy et al. [21] and Shimoyama [22] investigated the catalytic effects of clay minerals by heating natural fatty acids and alcohols in clay. The results showed that paraffin wax with a similar composition to crude oil and black carbonized residues attached to minerals was obtained, indicating that the catalysis of clay played a significant role in crude oil formation.
The catalytic effect of clay minerals, especially montmorillonite, in kerogen pyrolysis is widely recognized. However, the organic matter–water–rock reaction involved has not been fully explained. The catalytic effect of clay minerals in sedimentary layers mainly has the following two points: (1) the removal of heteroatoms from gum and asphaltene, promoting the formation of oil and the production of non-hydrocarbon substances such as CO2 and H2S, and (2) the condensation polymerization of kerogen, which forms C1-C15 short-chain alkanes through condensation polymerization (Figure 1):
Currently, the accepted theory suggests that montmorillonite can promote the production of light hydrocarbons by facilitating the cracking of hydrocarbons and the isomerization of normal hydrocarbons via a mechanism involving carbocation intermediates [23]. The Brønsted acid sites in the interlayer space of montmorillonite are produced by interlayer dissociation water, which significantly promotes this process. In contrast, n-alkane isomerization and olefin-alkane conversion through hydrogenation lead to high saturated hydrocarbon/n-alkane and olefin/alkane ratios in pyrolysis products [24]. Brønsted acid sites can promote the pyrolysis of nitrogen-containing organic compounds through Hoffman elimination [25], while Lewis acid sites participate in the decarboxylation reaction of aliphatic compounds during pyrolysis and CO2 is generated [26].
Ma et al. [27] have proposed that the primary catalytic mechanism of montmorillonite for low-maturity lignite pyrolysis is the free radical mechanism at low temperatures. Specifically, in the decarboxylation of organic matter, clay minerals may receive electrons from it during the decarboxylation of organic matter, which leads to the formation of free radicals. Free radicals cause rearrangement, resulting in the cleavage of the C single bond and the formation of short chain free hydrocarbons (Figure 1). However, the presence of water during the pyrolysis experiment has complicated results. Montmorillonite has different isomerization effects on alkanes in the presence or absence of water. Water molecules are adsorbed on the surface of montmorillonite and in the interlayer space, providing an oxygen source for the generation of CO2. It also inhibits the isomerization of n-alkanes in the pyrolysis process, leading to the dominant role of n-alkanes in the pyrolysis products. A lower interlayer water and surface adsorbed water content can greatly promote the isomerization of hydrocarbons and reduce the yield of CO2 [28].
Due to the complexity of crude oil, we believe that the free radical mechanism and the carbocation intermediate mechanism will occur simultaneously in the transformation process of crude oil and kerogen. The clay minerals will also dehydrate and rearrange themselves when heated, resulting in a reduction in interlayer volume and the transformation to illite on a macro level. However, the addition of water will increase the solid acidity of minerals. During the pyrolysis process, the hydroxyl group in water is adsorbed on the surface of minerals and proton H is provided to the organic matter. Although the current report cannot distinguish the degree of contribution, future experiments could be designed to focus on this process.
When it comes to illite, while it does exhibit catalytic effects, it is still slightly inferior to montmorillonite [29,30]. Due to illite lacking interlayer water, catalysis can only occur on its surface through solid acid sites. Hence, Lewis acid sites mainly occur on its outer surface [31]. The addition of illite slightly increased the relative cycloalkane content of kerogen-generated hydrocarbons in the experiment [32]. In contrast, the effect of kaolinite on the anhydrous pyrolysis of bituminous coal is not consistent with its effect in the pyrolysis with water. The addition of kaolinite inhibited the generation of light hydrocarbons but increased the yield when water was added [33].
In a study conducted by Espitalie [34], the effects of different clay minerals on the pyrolysis of type II kerogen were investigated. Results indicated that calcium montmorillonite showed better catalytic effects, promoting the formation of light hydrocarbons, and had better catalytic effects than sodium montmorillonite. On the other hand, kaolinite, illite, and attapulgite showed poor catalytic effects, but the residual products of attapulgite had a higher H/C ratio compared to the other two. The study also suggested that pyrolysis experiments in open systems at high temperatures and without water can enhance the interactions between minerals and hydrocarbons.
Roets et al. [35] found that kaolinite and quartz have little influence on the pyrolysis products of coal. Öztaş [36] showed that the treatment of coal with HF and HCl increased the pyrolysis efficiency of the treated samples, suggesting that clay minerals in coal may reduce the pyrolysis yield. He et al. [37] used the relative concentrations of C7 compounds to evaluate the difference in the catalytic effect of different clay mineral contents, with toluene, 1,2-dimethylcyclopentane, and n-heptane showing the most significant differences. According to Cai et al. [38], the catalytic effect of clay minerals is ranked as montmorillonite > kaolinite > illite, with kaolinite dominating in the high-mature to over-mature stage due to its free radical mechanism and amphiphilic nature.
In summary, the catalytic effect of clay minerals on the pyrolysis of hydrocarbons varies depending on the type of clay mineral and the experimental conditions. Montmorillonite has been found to be the most effective due to its interlayer water and solid acid sites in the interlayer space, while illite and kaolinite have limitations in their catalytic effect.

2.2. Catalysis of Alkali Metals

The deposition of organic matter, especially type III kerogen, is influenced by the terrestrial environment and contains significant amounts of alkali metals and alkaline earth metals, such as Na, K, Mg, and Ca. Among these elements, Ca is considered the most active alkaline earth metal [39,40]. Alkali metals in low maturity coal can exist in the form of hydrogen ions in oxygen functional groups or can be freely present in reservoirs such as various crystalline or amorphous minerals [41]. When heated, these alkali metals and alkaline earth metals decompose and generate basic oxides, such as MgO, CaO, Na2O, and K2O [42]. The basic oxides in coal promote the cracking of oxygen-containing functional groups, leading to a more compact coal structure. This process strengthens the reaction of tar macromolecules to form semi-coke, prevents tar macromolecules from escaping, and enhances the effect of demethanization and dehydrogenation [12]. For aromatic hydrocarbons, the catalysis of alkali metals can reduce the aromatic hydrocarbons with multi-side groups in pyrolysis products and increase the yield of small molecular aromatic hydrocarbons [43]. Generally speaking, alkali metals can accelerate the overall pyrolysis reaction rate, reduce the liquid yield, greatly improve the yield of gaseous hydrocarbons, and also reduce the contents of oxygen and sulfur in liquid products.
However, compared to marine carbonate reservoirs with vast storage capacity, the existing reports on the pyrolysis of kerogen III (coal) appear to be inconsistent with the reaction of crude oil, clay minerals, carbonate rocks, and kerogen. Karabakan and Yürüm [44] reported that carbonate minerals promote the pyrolysis of kerogen, and the total activation energy of kerogen decreases without HF treatment. They believe that alkaline earth metal oxides produced by carbonate minerals after heating promote the pyrolysis process. The reason why silicate minerals inhibit pyrolysis is not clear. Ballice et al. [45] came to a similar conclusion, stating that the inhibition of silicate minerals is greater than the promotion of carbonate in shale, which is attributed to the strong adsorption of clay minerals on hydrocarbons.

2.3. Influence of Carbonate Rock on Pyrolysis

Most carbonate rocks have a marine sedimentary environment and were formed during the early sedimentary period. Common carbonate minerals found in these rocks include aragonite (CaCO3), calcite (CaCO3), and dolomite (CaMg(CO3)2). Giant carbonate oil and gas reservoirs located in the Middle East have played a significant role in contributing to the world’s energy reserves [46]. The secondary changes that carbonate rocks undergo during sedimentation and diagenesis greatly influence their reservoir properties. For example, the process of dolomitization causes a reduction in the volume of limestone, which has the potential to increase the reservoir area [47]. Furthermore, sedimentary processes and secondary changes contribute to the creation of various types of pores, providing carbonate reservoirs with ample storage space [48]. However, the uneven development of carbonate reservoirs and the high sulfur content in some of these reservoirs limit further exploration and development [49]. As such, carbonate reservoirs have been a focal point of research for petroleum geologists.
Petroleum geologists have always been interested in carbonate rocks, which they believe can serve as source rocks due to the discovery of both reservoirs and source rocks in these formations [50,51,52]. Scientists have also studied the oil generation process of carbonate rocks. Unfortunately, the content of organic carbon in modern carbonate reservoirs is often very low. Currently, there are several explanations for this phenomenon: (1) Diagenesis and dissolution during the formation of carbonate rocks can cause the loss of organic matter, leading to a loss of 15%–50% of organic carbon in clastic rocks and 80%–90% of organic carbon in carbonate rocks [53]. (2) Organic matter may not deposit with carbonate rocks during the deposition process [54]. (3) The organic matter may have reached the threshold of hydrocarbon generation and been expelled. After hydrocarbon expulsion, the total organic carbon (TOC) of type I, type II, and type III kerogen decreased by 80%, 50%, and 20% [55], respectively, and carbonate rocks with high organic carbon content are not conducive to organic matter preservation after hydrocarbon expulsion [56]. (4) Organic matter may transform during carbonate rock deposition processes (e.g., running water, oxidation, and recrystallization), but this process appears to have little effect on organic carbon [57]. Reservoir geochemists find it challenging to understand carbonate reservoirs with conventional petroleum geology methods, and the indicators used to evaluate source rocks (e.g., vitrinite reflectance, organic matter deterioration degree) seem to have no effect on carbonate rocks [58]. This has prevented the development of a deep understanding of carbonate oil generation.
Calcite in carbonate rocks was first observed by Shimoyama and Johns [59] and has very close contact with organic matter during the pyrolysis of calcite fatty acids. They suggest that calcite provides catalysis in the pyrolysis process. However, it can be seen from the surface/interface relationship that the interface between calcite and polar mac-romolecules in crude oil [60] is more easily adsorbed. Therefore, this phenomenon is not enough to conclude that calcite has catalytic effect. Tannenbaum and Aizenshtat et al. [61] suggested that the heteroatoms present in carbonate rocks provide the essential catalysis for hydrocarbon generation from organic matter and kerogen. Jones et al. [62] concluded that there is no significant difference in the oil production process between carbonate rocks and shale, whereas Huizinga et al. [23] pointed out that CaCO3 has little impact on the thermal evolution and petroleum cracking behavior of kerogen. Pan et al. [63] further observed that the addition of calcite during crude oil cracking experiments had little effect on the cracking of main components in crude oil, but the formation rate of methane and total hydrocarbon gas was relatively low compared to other experimental groups. They postulated that calcite inhibits the secondary cracking of moisture (C2–C5) into methane.
Chen et al. [64] combined Rock–Eval pyrolysis data and hydrocarbon generation kinetics to conduct pyrolysis experiments on middle-upper Ordovician argillaceous carbonate rocks in Tarim Basin, China. The kerogen conversion rate reached 90%, indicating a high original potential. The argillaceous component has a relatively strong ability to preserve organic matter [65], but on the whole, carbonate rocks contain few argillaceous components. Therefore, whether well-preserved argillaceous carbonate rock experiments can accurately represent the hydrocarbon generation process in geological processes remains to be discussed.
Dolomite, unlike calcite, is currently regarded as a cheap and excellent catalyst for biomass gasification in the industry. Islam et al. [66] suggested that its catalytic effect comes from decomposition products, such as MgO and CaO, that form when the temperature exceeds 600 °C. However, the temperature of pyrolysis typically does not reach this limit. Thus, the catalytic mechanism of dolomite pyrolysis remains unclear.
Wu et al. [67] studied the effect of dolomite and calcite on the pyrolysis of crude oil using closed flexible gold capsules. They found that the presence of dolomite slightly reduced the yield of C1–C5 but significantly increased the yield of C6–C13. Moreover, they conducted pyrolysis experiments of MgCO3 and MgCO3 + CaCO3, which produced similar results to those of petroleum and dolomite. Thus, it is evident that the influence of dolomite on petroleum pyrolysis is closely related to the content of Mg in dolomite. In addition, the lower content of dolomite can promote the TSR (Thermochemical Sulfate Reduction) of crude oil. Specifically, dolomite -MgCO3 promotes the formation of SO42−. The coexistence of SO42−, water, and H2S then initiates the sulfate thermochemical reduction (TSR) reaction, which destroys hydrocarbons and produces CO2 and H2S [68]. However, the high content of dolomite can absorb or convert some H2S into other forms [69]. Moreover, the reaction of MgCO3 with water can generate Mg(OH)2 (R2), which further reduces the amount of H2S generated.
Mg(OH)2 + H2S→MgS + 2H2O MgS + 2O2→MgSO4
Maimaiti et al. [70] conducted a pyrolysis study using the same method as Wu et al. [67] to investigate the effect of carbonate rocks on crude oil cracking. They observed that different mineral types, such as calcite and dolomite, had varying degrees of inhibition on methane production at high maturity stages (EasyRo > 3.5%), but had no effect on C2–C5 hydrocarbons, while simultaneously inhibiting C14+. Additionally, the yield of CO2 in the pyrolysis products with carbonate minerals increased significantly, and unique changes in carbon isotopes were observed. Unlike typical thermally driven isotope fractionation, the addition of dolomite promoted the shift of carbon isotopes of methane, whereas calcite inhibited it. In crude oil + dolomite, crude oil + calcite, and simple crude oil pyrolysis experiments, δ13C1 was −49.89, −54.08, and −51.77%, respectively (EasyRo = 0.96%). However, the carbon isotope fractionation process of ethane and propane remained unaffected by minerals. They speculate that this phenomenon may be due to dolomite increasing the diversity of methane sources during pyrolysis and the differential adsorption of various organic gases by different minerals. Nevertheless, there is still insufficient evidence to support this claim.
Furthermore, Wang et al. [71] observed that increasing the carbonate concentration in source rocks led to a corresponding increase in the yield of CO2 and heavier carbon isotopes in CO2. They believe that carbonate rocks significantly contribute to CO2 in pyrolysis products, which is derived from the dissolution of carbonate minerals by organic acids during pyrolysis, despite the lower pyrolysis temperature compared to the decomposition temperature of carbonate rocks.

2.4. Effect of Metal Minerals on Pyrolysis Experiments

The precursors of crude oil and kerogen contain many metal elements that exist in organisms in the form of metalloenzymes, playing an important role in the evolution of life and the maintenance of biological functions [72]. These metal elements can be captured at any stage of deposition, hydrocarbon generation, and migration, and are relatively enriched in macromolecules such as colloid and asphaltene, and even complexed or bonded into kerogen [73]. For instance, porphyrin, which is widely found in chlorophyll and heme, is also present in petroleum and kerogen, with the metal elements coordinated being different depending on the living body and its environmental conditions [74]. The proportions of vanadoporphyrin and nickel porphyrin in kerogen, crude oil, and rock represent marine and continental sedimentary environments, respectively. Additionally, these bio-lipids are also considered important catalysts for transforming into rearranged alkanes in crude oil, and the relative content of rearranged alkanes in crude oil produced in some geological periods lacking chlorophyll is lower [75]. Furthermore, these metals and their oxides are important catalysts for the transformation of the Earth’s environment from simple to complex [76]. Based on chemical principles, active intermediates are formed by the coordination between ligands and receptors due to the d orbitals around the metal elements not being fully occupied. This reduces the activation energy of the reaction and promotes the formation of organic matter [77]. Moreover, the larger internal pores generated after the organic–inorganic interaction between minerals and organic matter also promote this process [78]. However, Lewan et al. [79] presented different opinions, reporting no obvious difference in gas type, yield, dryness, and formation rate of low-cooked kerogen after 72 h of aqueous pyrolysis at 330 and 355 °C. They argued that various transition metals with catalytic properties were not in an active state during the geological maturation of kerogen and exhibited suitable catalytic activity.
Gao and colleagues [80] conducted anhydrous pyrolysis experiments on low-rank coal with the transition metal Ni. The addition of the transition metal Ni catalyzed the decomposition of organic matter and the synthesis of abiogenic methane during the maturity of source rocks. With the addition of Ni, the content of CH4 in pyrolysis products increased, and the yields of ethane and propane decreased. They concluded that Ni played a catalytic role in the whole process of kerogen, crude oil, and heavy hydrocarbon gas cracking, and Fischer–Tropsch Type (FTT) reaction occurred in the pyrolysis experiment, which catalyzed together with heavy hydrocarbon gas cracking, significantly improving methane yield and greatly improving CH4 yield at all temperatures (see Figure 2). In all the experimental groups, δD became smaller, but δ13C gradually became positive with the increase in temperature.
He et al. [81] investigated the impact of iron-containing minerals, specifically FeCO3, FeS2, and Fe2O3, on the pyrolysis of 1-methylnaphthalene under high pressure. They explored different catalytic mechanisms by utilizing isotope differences at different temperature ranges. The coexistence of water and minerals greatly increased the ratio of i-C4/n-C4 isomers, which is consistent with the pyrolysis trend observed in water-bearing minerals by Cai et al. [82]. The reactions at different mineral interfaces also varied. The presence of FeS2 and Fe2O3 minerals promoted the reaction between water and hydrocarbons, leading to a significant increase in CH4 yield. In contrast, the experimental group containing FeCO3 showed little difference in CH4 yield but exhibited positive carbon isotope. The authors hypothesized that the FTT reaction occurs at the interface between minerals and hydrocarbons at higher temperatures (400–450 °C), contributing to abiotic CH4 production.
The catalytic action of transition metals in the pyrolysis process is related to their chemical forms. For instance, in the low-temperature stage, Fe3+ combines with oxygen-containing functional groups in coal or π bonds in unsaturated hydrocarbons using its own d empty orbital, resulting in chemical adsorption. The cations in coal have a positive effect on the organic structure of coal, making the coal molecules more polar and increasing their pyrolysis activity. As the temperature increases, the form of iron changes into crystalline oxide, which can catalyze the transformation of heavy components into primary pyrolysis products, reducing tar yield and increasing the gaseous yield [82]. However, He et al. [83] argued that transition metals, including iron, have a catalytic effect only in the elemental state, which can explain the differences observed in the experiment by Lewan et al. [79]. Guan et al. [84] reported that the different occurrence states of the same element have a significant influence on the results of thermal simulation experiments. In thermal simulation experiments conducted at low temperatures for an extended period (150 °C, 30 days), the Mn solution promoted crude oil yield, whereas Mn elemental substance showed obvious inhibition. Interestingly, the final crude oil yield of low-concentration Mn solution was 18.19 kg/t TOC, much lower than that of high-concentration Mn solution, which yielded 5.4 kg/t TOC (Figure 3).

3. Thermal Simulation Process Promoted by Catalytic Hydrogenation of Minerals

Hydrogen plays a crucial role in the transformation of organic matter, as it not only provides the material basis for kerogen but also enhances the hydrogenation reaction during oil and gas generation and promotes the thermal evolution of source rocks. Specifically, hydrogen atoms can combine with carbon free radicals during the thermal simulation process, leading to the formation of smaller molecular hydrocarbons and ultimately increasing the final hydrocarbon yield [85,86]. Surprisingly, even a highly evolved solid pitch can generate hydrocarbons again under the influence of hydrogen [87].
Several sources of exogenous hydrogen are present in sedimentary basins, including water in sedimentary environments, hydrogen-containing fluids resulting from volcanic hydrothermal activity in the mantle, serpentine in ultrabasic volcanic rocks, and enrichment and radiation of radioactive minerals.

3.1. Water Is an Important Source of Hydrogen for Hydrocarbon in Natural Basins

Water has been reported to increase the production of gaseous hydrocarbons in pyrolysis experiments [88,89]. This is because water undergoes a disproportionation reaction during pyrolysis reaction, releasing hydrogen that promotes the cracking of kerogen oil or crude oil. This reduces cross-linking reactions and promotes thermal cracking, thereby altering the original pyrolysis law [90].
Water is transported to the deep part of the earth in the subduction zone of the subducting plate [91], where it reacts with iron under different temperature and pressure conditions to form a unique high-magnesium iron hydride [92,93]. Finally, dehydrogenation takes place and hydrogen-rich fluids in the mantle return to the surface through volcanic hydrothermal activity [94,95]. Meanwhile, hydrogen plays various roles in the Earth’s C cycle [96,97,98,99]. Mantle-derived gas also has a profound impact on the transformation of organic matter and the formation of hydrocarbons. The hydrocarbons that may be affected by mantle-derived gas can be determined by the difference in hydrogen isotopes [95,100,101,102,103].

3.2. The Role of Hydrogen Generation from Serpentine in Pyrolysis

During the early stages of Earth’s evolution, catalytic metals carried up from the mantle by volcanic activity were believed to play a significant role in the formation of primitive life and small organic molecules(Table 1). Transition metals such as iron, manganese, cobalt, zinc, and vanadium were transported into the ancient ocean via hydrothermal vents. Similar to their modern-day counterparts, the hydrothermal vents in the deep sea support a variety of complex biological and chemical processes, including sulfate reduction, iron oxidation and reduction, serpentine formation, and methane production, which contribute to the largest primary productivity in the deep ocean [104,105,106,107,108].
The process of igneous oxidation of olivine and pyroxene in the lower oceanic crust and upper mantle of the earth is referred to as serpentinization. In the presence of water, the Fe(II) present in these minerals is oxidized by Fe(III), leading to the precipitation of Fe3O4 and the release of diatomic H2, which can react with carbon to produce methane (as shown in Formula (2)). This process generates a substantial amount of heat and significantly affects the pH around the minerals, resulting in the deposition of carbonate rocks that are widespread in the hydrothermal vents of modern mid-ocean ridges.
(Fe,Mg)2SiO4 + H2O + CO2 → Mg3Si2O5(OH)4 + Fe3O4 + CH4
Hydrocarbons, primarily C1–C4 and formic acid, are the most commonly detected compounds in modern hydrothermal activities on the ocean floor. Although only a small portion of these compounds are believed to be produced by the thermal cracking of other hydrocarbons, recent studies have detected some macromolecular hydrocarbons and lipid compounds with the participation of microorganisms near the deep sea [109,110]. However, the source of macromolecular compounds near hydrothermal solutions in the marine environment is still controversial [111], and man-made organic pollution is one possible factor. It is also challenging to distinguish the contribution of organic and inorganic actions to hydrocarbons. Therefore, only a small portion of CH4 is currently considered to be generated by the cracking of other hydrocarbons [112].
When organic matter is affected by volcanoes, it can also be impacted [113,114,115]. Different stages of volcanic hydrothermal activity in sedimentary basins have different catalytic behaviors on hydrocarbon behavior. The upward movement of magma is selective, and without other favorable channels, sedimentary basins that lack rigidity are more susceptible to invasion [116]. Once organic matter accumulates, the heat brought by the magmatic hydrothermal solution accelerates its maturity, promotes the thermal metamorphism of mudstone and sandstone, squeezes clastic material, and produces structural cracks. If this effect occurs during the sedimentary period, it can change the original sedimentary environment [117]. Metal minerals and other inorganic gases with catalytic properties brought by hydrothermal solutions will provide all the necessary elements for primary productivity, resulting in explosive growth of aquatic organisms. The addition of volcanic ash leads to water stratification, which enhances the reduction and preservation conditions of organic matter in the basin, improves the primary biological productivity of the basin, and easily forms high-quality source rocks, as shown in Figure 4 [84].

3.3. The Role of Radioactive Hydrogen Production from Uranium in Pyrolysis

Uranium, a special transition metal element, is the cornerstone of the nuclear industry and has garnered significant attention for its potential in energy transformation [118]. Geologists are particularly interested in the unique sandstone-type pitchblende and its associated relationships and formation mechanisms. This phenomenon has been reported in most sedimentary basins worldwide [119,120,121,122,123,124].
In general, the deposition of uranium in sedimentary basins is closely related to the transportation, reduction, enrichment, and preservation of organic matter (Figure 5) [119,125,126]. Uranium is highly reactive and can react with most elements, especially in terms of its strong reduction ability. Under reduction conditions, U(VI) is converted to an immobile tetravalent state, U(IV) (see Formula (3)), which predominantly occurs in shale that is rich in organic matter [127]. The reduction and mineralization of uranium may occur simultaneously with the bacterial sulfate reduction in crude oil [128,129].
U O 2 2 + + C H 4 + 3 H 2 O 4 U O 2 + H C O 3 + 9 H +
When uranium is enriched to a certain extent, high radioactive anomalies can also be found in the formation. The decay of uranium can release α, β, and γ rays, and the particles in these rays can decompose water to provide  H 1 3 4  for hydrogen and oxygen, resulting in the generation of active hydrogen. The radioactive decay and the provision of hydrogen are also considered to promote the pyrolysis process of organic matter in natural strata. The reaction process is as follows:
H 2 O + ( α , γ ) H · + O H ·
2 O H · H O O H
2 H O O H 2 H 2 O + O 2
2 H · H 2
Uranium enrichment can result in high radioactive anomalies in formations, releasing α, β, and γ rays that can decompose water to generate foreign hydrogen and promote the pyrolysis process of organic matter in natural strata. This foreign hydrogen compensates for the hydrogen required for the thermal degradation of organic matter and enhances the methane yield and total hydrocarbon yield [86]. In sedimentary basins, compounds can reach metastable thermodynamic equilibrium during hydrocarbon generation, leading to the formation of olefin-alkane equilibrium when the H concentration is sufficient [130]. Uranium ore can catalyze the pyrolysis of immature kerogen II, as discussed by Cai et al. [131], with liquid and gas products increasing significantly during gold tube closed-system pyrolysis. Deuterium in hydrocarbon gas was consumed, indicating that water and uranium ore can participate in the thermal evolution of organic matter and provide a hydrogen source for hydrocarbon generation. However, under this condition, the carbocation mechanism is inhibited to some extent, as the i-C4/n-C4 ratio in the experimental group containing uranium ore and water is low.
According to Yang et al. [132], the catalysis of uranium can cause an exponential change in kerogen structure. With an increase in uranium content in kerogen, the gas/oil ratio and aromatic hydrocarbon content of pyrolysis products continuously increase compared to the hydrocarbon generation law of conventional marine shale. FT-ICR-MS results show that macromolecular compounds in uranium-rich samples are more alkylated and oxygenated compounds are more abundant, which can be attributed to free radicals generated by the in situ reaction of highly active OH produced by the radioactive cracking of water with organic matter during diagenesis. However, the more complex cross-linked structure of uranium-rich samples can hinder further cracking.

4. Conclusions and Outlook

Previous studies have shown that different minerals exhibit different forms of catalysis in laboratory-scale experiments. However, detailed consideration of the geological background and the corresponding design of thermal simulation experiments remain a primary concern, particularly with respect to shale and deep carbonate rocks that are currently the focus of research in China. Future research should focus on the development and application of combined measuring equipment, such as TG-IR, TG-MS, and TG-XRD, which can more accurately reflect the catalytic process and mechanism by monitoring changes in molecular markers, minerals, and parameters during the experimental process. Furthermore, with the continuous development of computational chemistry, there is a need to explore more accurate and precise modeling of thermal simulation experiments on computers. Numerical simulations can easily change experimental parameters and provide a more intuitive molecular reaction mechanism. They are also more environmentally friendly and faster than thermal simulation experiments.

Author Contributions

Conceptualization and funding acquisition, H.H.; methodology and writing—review and editing, M.Z.; data curation, P.X.; writing—original draft preparation, visualization, J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (Grant number 41873049).

Data Availability Statement

Data available on request from the authors.

Conflicts of Interest

We declare that we have no known competitive financial interests or personal relationships.

References

  1. Hunt, J.M. Distribution of carbon in crust of earth. AAPG Bull. 1972, 56, 2273–2277. [Google Scholar]
  2. Han, Q.; Li, M.; Liu, X.; Xiao, H.; Ren, J.; Guo, C. A maturation scale for molecular simulation of kerogen thermal degradation. Org. Geochem. 2023, 175, 104507. [Google Scholar] [CrossRef]
  3. Wang, Z.; Tang, Y.; Wang, Y.; Zheng, Y.; Chen, F.; Wu, S.; Fu, D. Kinetics of shale oil generation from kerogen in saline basin and its exploration significance: An example from the Eocene Qianjiang Formation, Jianghan Basin, China. J. Anal. Appl. Pyrolysis 2020, 150, 104885. [Google Scholar] [CrossRef]
  4. Lai, D.; Zhan, J.-H.; Tian, Y.; Gao, S.; Xu, G. Mechanism of kerogen pyrolysis in terms of chemical structure transformation. Fuel 2017, 199, 504–511. [Google Scholar] [CrossRef]
  5. Spigolon, A.L.; Lewan, M.D.; de Barros Penteado, H.L.; Coutinho, L.F.C.; Mendonça Filho, J.G. Evaluation of the petroleum composition and quality with increasing thermal maturity as simulated by hydrous pyrolysis: A case study using a Brazilian source rock with Type I kerogen. Org. Geochem. 2015, 83, 27–53. [Google Scholar] [CrossRef]
  6. Tian, H.; Wang, Z.; Xiao, Z.; Li, X.; Xiao, X. Oil cracking to gases: Kinetic modeling and geological significance. Chin. Sci. Bull. 2006, 51, 2763–2770. [Google Scholar] [CrossRef]
  7. Lu, Y.; Wang, Y.; Zhang, J.; Wang, Q.; Zhao, Y.; Zhang, Y. Investigation on the characteristics of pyrolysates during co-pyrolysis of Zhundong coal and Changji oil shale and its kinetics. Energy 2020, 200, 117529. [Google Scholar] [CrossRef]
  8. Behar, F.; Vandenbroucke, M.; Tang, Y.; Marquis, F.; Espitalie, J. Thermal cracking of kerogen in open and closed systems: Determination of kinetic parameters and stoichiometric coefficients for oil and gas generation. Org. Geochem. 1997, 26, 321–339. [Google Scholar] [CrossRef]
  9. Shao, D.; Ellis, G.S.; Li, Y.; Zhang, T. Experimental investigation of the role of rock fabric in gas generation and expulsion during thermal maturation: Anhydrous closed-system pyrolysis of a bitumen-rich Eagle Ford Shale. Org. Geochem. 2018, 119, 22–35. [Google Scholar] [CrossRef]
  10. Zhao, Z.; Feng, Q.; Liu, X.; Lu, H.; Peng, P.A.; Liu, J.; Hsu, C.S. Petroleum maturation processes simulated by high-pressure pyrolysis and kinetic modeling of low-maturity type I Kerogen. Energy Fuels 2022, 36, 1882–1893. [Google Scholar] [CrossRef]
  11. Nelson, P.F.; Smith, I.W.; Tyler, R.J.; Mackie, J.C. Pyrolysis of coal at high temperatures. Energy Fuels 1988, 2, 391–400. [Google Scholar] [CrossRef]
  12. Sathe, C.; Pang, Y.; Li, C.-Z. Effects of heating rate and ion-exchangeable cations on the pyrolysis yields from a Victorian brown coal. Energy Fuels 1999, 13, 748–755. [Google Scholar] [CrossRef]
  13. Chen, J.; Zhang, X.; Chen, Z.; Pang, X.; Yang, H.; Zhao, Z.; Pang, B.; Ma, K. Hydrocarbon expulsion evaluation based on pyrolysis Rock-Eval data: Implications for Ordovician carbonates exploration in the Tabei Uplift, Tarim. J. Pet. Sci. Eng. 2021, 196, 107614. [Google Scholar] [CrossRef]
  14. Peng, W.; Hu, G.; Liu, Q.; Jia, N.; Fang, C.; Gong, D.; Yu, C.; Lyu, Y.; Wang, P.; Feng, Z. Research status on thermal simulation experiment and several issues of concern. J. Nat. Gas Geosci. 2018, 3, 283–293. [Google Scholar] [CrossRef]
  15. Ma, W.; Hou, L.; Luo, X.; Liu, J.; Tao, S.; Guan, P.; Cai, Y. Generation and expulsion process of the Chang 7 oil shale in the Ordos Basin based on temperature-based semi-open pyrolysis: Implications for in-situ conversion process. J. Pet. Sci. Eng. 2020, 190, 107035. [Google Scholar] [CrossRef]
  16. Merriman, R.J. Clay minerals and sedimentary basin history. Eur. J. Mineral. 2005, 17, 7–20. [Google Scholar] [CrossRef] [Green Version]
  17. Sposito, G.; Skipper, N.T.; Sutton, R.; Park, S.-H.; Soper, A.K.; Greathouse, J.A. Surface geochemistry of the clay minerals. Proc. Natl. Acad. Sci. USA 1999, 96, 3358–3364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Wilson, M.; Wilson, L.; Patey, I.; Shaw, H. The influence of individual clay minerals on formation damage of reservoir sandstones: A critical review with some new insights. Clay Miner. 2014, 49, 147–164. [Google Scholar] [CrossRef]
  19. Chen, S.; Han, Y.; Fu, C.; Zhu, Y.; Zuo, Z. Micro and nano-size pores of clay minerals in shale reservoirs: Implication for the accumulation of shale gas. Sediment. Geol. 2016, 342, 180–190. [Google Scholar] [CrossRef]
  20. Thomas, C.L.; Hickey, J.; Stecker, G. Chemistry of clay cracking catalysts. Ind. Eng. Chem. 1950, 42, 866–871. [Google Scholar] [CrossRef]
  21. Eltantawy, I.; Arnold, P. Catalytic decomposition of organic molecules by clays. Nat. Phys. Sci. 1973, 244, 144. [Google Scholar] [CrossRef]
  22. Shimoyama, A.; Johns, W.D. Catalytic conversion of fatty acids to petroleum-like paraffins and their maturation. Nat. Phys. Sci. 1971, 232, 140–144. [Google Scholar] [CrossRef]
  23. Huizinga, B.J.; Tannenbaum, E.; Kaplan, I.R. The role of minerals in the thermal alteration of organic matter—IV. Generation of n-alkanes, acyclic isoprenoids, and alkenes in laboratory experiments. Geochim. Et Cosmochim. Acta 1987, 51, 1083–1097. [Google Scholar] [CrossRef] [PubMed]
  24. Yuan, P.; Liu, H.; Liu, D.; Tan, D.; Yan, W.; He, H. Role of the interlayer space of montmorillonite in hydrocarbon generation: An experimental study based on high temperature–pressure pyrolysis. Appl. Clay Sci. 2013, 75, 82–91. [Google Scholar] [CrossRef]
  25. Liu, H.; Yuan, P.; Liu, D.; Bu, H.; Song, H.; Qin, Z.; He, H. Pyrolysis behaviors of organic matter (OM) with the same alkyl main chain but different functional groups in the presence of clay minerals. Appl. Clay Sci. 2018, 153, 205–216. [Google Scholar] [CrossRef]
  26. Bu, H.; Yuan, P.; Liu, H.; Liu, D.; Liu, J.; He, H.; Zhou, J.; Song, H.; Li, Z. Effects of complexation between organic matter (OM) and clay mineral on OM pyrolysis. Geochim. Et Cosmochim. Acta 2017, 212, 1–15. [Google Scholar] [CrossRef]
  27. Ma, X.; Zheng, G.; Sajjad, W.; Xu, W.; Fan, Q.; Zheng, J.; Xia, Y. Influence of minerals and iron on natural gases generation during pyrolysis of type-III kerogen. Mar. Pet. Geol. 2018, 89, 216–224. [Google Scholar] [CrossRef]
  28. Song, H.; Liu, H.; Bu, H.; Liu, D.; Li, Y.; Du, P. Effects of montmorillonite charge reduction on the high-temperature/high-pressure pyrolysis of organic matter. Appl. Clay Sci. 2021, 213, 106243. [Google Scholar] [CrossRef]
  29. Espitalie, J.; Madec, M.; Tissot, B. Role of mineral matrix in kerogen pyrolysis: Influence on petroleum generation and migration. AAPG Bull. 1980, 64, 59–66. [Google Scholar]
  30. Horsfield, B.; Douglas, A. The influence of minerals on the pyrolysis of kerogens. Geochim. Et Cosmochim. Acta 1980, 44, 1119–1131. [Google Scholar] [CrossRef]
  31. Wu, L.M.; Zhou, C.H.; Keeling, J.; Tong, D.S.; Yu, W.H. Towards an understanding of the role of clay minerals in crude oil formation, migration and accumulation. Earth-Sci. Rev. 2012, 115, 373–386. [Google Scholar] [CrossRef]
  32. Thompson, K.F. Mechanisms controlling gas and light end composition in pyrolysates and petroleum: Applications in the interpretation of reservoir fluid analyses. Org. Geochem. 2006, 37, 798–817. [Google Scholar] [CrossRef]
  33. Franklin, H.D.; Peters, W.A.; Howard, J.B. Mineral matter effects on the rapid pyrolysis and hydropyrolysis of a bituminous coal. 1. Effects on yields of char, tar and light gaseous volatiles. Fuel 1982, 61, 155–160. [Google Scholar] [CrossRef]
  34. Espitalié, J.; Makadi, K.S.; Trichet, J. Role of the mineral matrix during kerogen pyrolysis. Org. Geochem. 1984, 6, 365–382. [Google Scholar] [CrossRef]
  35. Roets, L.; Bunt, J.R.; Neomagus, H.W.; Strydom, C.A.; Van Niekerk, D. The effect of added minerals on the pyrolysis products derived from a vitrinite-rich demineralised South African coal. J. Anal. Appl. Pyrolysis 2016, 121, 41–49. [Google Scholar] [CrossRef]
  36. Öztaş, N.; Yürüm, Y. Pyrolysis of Turkish Zonguldak bituminous coal. Part 1. Effect of mineral matter. Fuel 2000, 79, 1221–1227. [Google Scholar] [CrossRef]
  37. He, M.; Wang, Z.; Moldowan, M.J.; Peters, K. Insights into catalytic effects of clay minerals on hydrocarbon composition of generated liquid products during oil cracking from laboratory pyrolysis experiments. Org. Geochem. 2022, 163, 104331. [Google Scholar] [CrossRef]
  38. Cai, J.; Du, J.; Song, M.; Lei, T.; Wang, X.; Li, Y. Control of clay mineral properties on hydrocarbon generation of organo-clay complexes: Evidence from high-temperature pyrolysis experiments. Appl. Clay Sci. 2022, 216, 106368. [Google Scholar] [CrossRef]
  39. Linares-Solano, A.; Hippo, E.J.; Walker, P.L., Jr. Catalytic activity of calcium for lignite char gasification in various atmospheres. Fuel 1986, 65, 776–779. [Google Scholar] [CrossRef]
  40. Skodras, G.; Sakellaropoulos, G. Mineral matter effects in lignite gasification. Fuel Process. Technol. 2002, 77, 151–158. [Google Scholar] [CrossRef]
  41. Walker, P.L., Jr. Catalysis of lignite char gasification by exchangeable calcium and magnesium. Fuel 1984, 63, 1214–1220. [Google Scholar]
  42. Li, C.-Z.; Sathe, C.; Kershaw, J.R.; Pang, Y. Fates and roles of alkali and alkaline earth metals during the pyrolysis of a Victorian brown coal. Fuel 2000, 79, 427–438. [Google Scholar] [CrossRef]
  43. Qiao, L.; Deng, C.; Lu, B.; Wang, Y.; Wang, X.; Deng, H.; Zhang, X. Study on calcium catalyzes coal spontaneous combustion. Fuel 2022, 307, 121884. [Google Scholar] [CrossRef]
  44. Karabakan, A.; Yürüm, Y. Effect of the mineral matrix in the reactions of oil shales: 1. Pyrolysis reactions of Turkish Göynük and US Green River oil shales. Fuel 1998, 77, 1303–1309. [Google Scholar] [CrossRef]
  45. Ballice, L. Stepwise chemical demineralization of Göynük (Turkey) oil shale and pyrolysis of demineralization products. Ind. Eng. Chem. Res. 2006, 45, 906–912. [Google Scholar] [CrossRef]
  46. Trice, R. Challenges and insights in optimizing oil production from Middle East mega karst reservoirs. In Proceedings of the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 12 March 2005; Volume 3. [Google Scholar]
  47. Weber, J.; Cheshire, M.C.; Bleuel, M.; Mildner, D.; Chang, Y.-J.; Ievlev, A.; Littrell, K.C.; Ilavsky, J.; Stack, A.G.; Anovitz, L.M. Influence of microstructure on replacement and porosity generation during experimental dolomitization of limestones. Geochim. Et Cosmochim. Acta 2021, 303, 137–158. [Google Scholar] [CrossRef]
  48. Ibrahem, Y.; Morozov, V.; Sudakov, V.; Idrisov, I.; Kolchugin, A. Sedimentary diagenesis and pore characteristics for the reservoir evaluation of Domanik formations (Semiluksk and Mendymsk) in the central part of Volga-Ural petroleum province. Pet. Res. 2022, 7, 32–46. [Google Scholar] [CrossRef]
  49. Gürgey, K. An attempt to recognise oil populations and potential source rock types in Paleozoic sub-and Mesozoic-Cenozoic supra-salt strata in the southern margin of the Pre-Caspian Basin, Kazakhstan Republic. Org. Geochem. 2002, 33, 723–741. [Google Scholar] [CrossRef]
  50. Klemme, H.; Ulmishek, G.F. Effective petroleum source rocks of the world: Stratigraphic distribution and controlling depositional factors. AAPG Bull. 1991, 75, 1809–1851. [Google Scholar]
  51. Liu, W.; Xu, Y.; Chen, J. Comprehensive geochemical identification of highly evolved marine carbonate rocks as hydrocarbon-source rocks as exemplified by the Ordos Basin. Sci. China Ser. D 2006, 49, 384–396. [Google Scholar]
  52. Zhu, G.; Yan, H.; Chen, W.; Yan, L.; Zhang, K.; Li, T.; Chen, Z.; Wu, G.; Santosh, M. Discovery of Cryogenian interglacial source rocks in the northern Tarim, NW China: Implications for Neoproterozoic paleoclimatic reconstructions and hydrocarbon exploration. Gondwana Res. 2020, 80, 370–384. [Google Scholar] [CrossRef]
  53. Morad, S. Carbonate cementation in sandstones: Distribution patterns and geochemical evolution. In Carbonate Cementation in Sandstones: Distribution Patterns and Geochemical Evolution; Wiley-Blackwell: Oxford, UK, 1998; pp. 1–26. [Google Scholar]
  54. Hunt, J.M. The Origin of Petroleum in Carbonate Rocksk1. In Developments in Sedimentology; Elsevier: Amsterdam, The Netherlands, 1967; Volume 9, pp. 225–251. [Google Scholar]
  55. Daly, A.R.; Edman, J.D. EALoss of Organic Carbon from Source Rocks during Thermal Maturation; AAPG Bulletin: Tulsa, OK, USA, 1987. [Google Scholar]
  56. Xia, L.-W.; Cao, J.; Wang, M.; Mi, J.-L.; Wang, T.-T. A review of carbonates as hydrocarbon source rocks: Basic geochemistry and oil–gas generation. Pet. Sci. 2019, 16, 713–728. [Google Scholar] [CrossRef] [Green Version]
  57. Tissot, B.; du Petrole, E.N.S.; Welte, D. Petroleum Formation and Occurrence. In A New Approach to Oil and Gas Exploration; Springer-Verlag: Berlin, Germany, 1978. [Google Scholar]
  58. Huo, Z.; Pang, X.; Chen, J.; Zhang, J.; Song, M.; Guo, K.; Li, P.; Li, W.; Liang, Y. Carbonate source rock with low total organic carbon content and high maturity as effective source rock in China: A review. J. Asian Earth Sci. 2019, 176, 8–26. [Google Scholar] [CrossRef]
  59. Shimoyama, A.; Johns, W.D. Formation of alkanes from fatty acids in the presence of CaCO3. Geochim. Et Cosmochim. Acta 1972, 36, 87–91. [Google Scholar] [CrossRef]
  60. Subramanian, S.; Sørland, G.H.; Simon, S.; Xu, Z.; Sjöblom, J. Asphaltene fractionation based on adsorption onto calcium carbonate: Part 2. Self-association and aggregation properties. Colloids Surf. A Physicochem. Eng. Asp. 2017, 514, 79–90. [Google Scholar] [CrossRef]
  61. Tannenbaum, E.; Aizenshtat, Z. Formation of immature asphalt from organic-rich carbonate rocks—I. Geochemical correlation. Org. Geochem. 1985, 8, 181–192. [Google Scholar] [CrossRef]
  62. Jones, R. Comparison of Carbonate and Shale Source Rocks; AAPG Bulletin: Tulsa, OK, USA, 1984. [Google Scholar]
  63. Pan, C.; Jiang, L.; Liu, J.; Zhang, S.; Zhu, G. The effects of calcite and montmorillonite on oil cracking in confined pyrolysis experiments. Org. Geochem. 2010, 41, 611–626. [Google Scholar] [CrossRef]
  64. Chen, Z.; Chai, Z.; Cheng, B.; Liu, H.; Cao, Y.; Cao, Z.; Qu, J. Geochemistry of high-maturity crude oil and gas from deep reservoirs and their geological significance: A case study on Shuntuoguole low uplift, Tarim Basin, western China. AAPG Bull. 2021, 105, 65–107. [Google Scholar] [CrossRef]
  65. Lanyu, W.; Dongfeng, H.; Yongchao, L.; Ruobing, L.; Xiaofeng, L. Advantageous shale lithofacies of Wufeng Formation-Longmaxi Formation in Fuling gas field of Sichuan Basin, SW China. Pet. Explor. Dev. 2016, 43, 208–217. [Google Scholar]
  66. Islam, M.W. A review of dolomite catalyst for biomass gasification tar removal. Fuel 2020, 267, 117095. [Google Scholar] [CrossRef]
  67. Wu, L.; Fang, X.; Geng, A. Influence of carbonate minerals on the pyrolysis and phase behavior of oil. Mar. Pet. Geol. 2021, 134, 105374. [Google Scholar] [CrossRef]
  68. Machel, H.G.; Krouse, H.R.; Sassen, R. Products and distinguishing criteria of bacterial and thermochemical sulfate reduction. Appl. Geochem. 1995, 10, 373–389. [Google Scholar] [CrossRef]
  69. Shand, M.A. The Chemistry and Technology of Magnesia; John Wiley & Sons: Hoboken, NJ, USA, 2006. [Google Scholar]
  70. Maimaiti, A.; Wang, Q.; Hao, F.; Yang, X.; Zhang, H.; Tian, J. Oil pyrolysis with carbonate minerals: Implications for the thermal stability of deep crude oil. Mar. Pet. Geol. 2022, 146, 105929. [Google Scholar] [CrossRef]
  71. Wang, W.; Ji, L.; Song, D.; Zhang, D.; Lü, C.; Su, L. Origin of inorganic carbon dioxide associated with hydrocarbon generation: Evidence from hydrous pyrolysis experiments and natural and shale gases. J. Asian Earth Sci. X 2022, 7, 100079. [Google Scholar] [CrossRef]
  72. Bullock, R.M.; Chen, J.G.; Gagliardi, L.; Chirik, P.J.; Farha, O.K.; Hendon, C.H.; Jones, C.W.; Keith, J.A.; Klosin, J.; Minteer, S.D. Using nature’s blueprint to expand catalysis with Earth-abundant metals. Science 2020, 369, eabc3183. [Google Scholar] [CrossRef]
  73. Curiale, J. Distribution and Occurrence of Metals in Heavy Crude Oils and Solid Bitumens--Implications for Petroleum Exploration: Section II. In Characterization, Maturation, and Degradation; AAPG Bulletin: Tulsa, OK, USA, 1987. [Google Scholar]
  74. Treibs, A. Chlorophyll-und Häminderivate in bituminösen Gesteinen, Erdölen, Erdwachsen und Asphalten. Ein Beitrag zur Entstehung des Erdöls. Justus Liebigs Ann. Der Chem. 1934, 510, 42–62. [Google Scholar] [CrossRef]
  75. Mango, F.D. Transition metal catalysis in the generation of petroleum: A genetic anomaly in Ordovician oils. Geochim. Et Cosmochim. Acta 1992, 56, 3851–3854. [Google Scholar] [CrossRef]
  76. Martin, W.; Baross, J.; Kelley, D.; Russell, M.J. Hydrothermal vents and the origin of life. Nat. Rev. Microbiol. 2008, 6, 805–814. [Google Scholar] [CrossRef]
  77. Miller, D.M.; Buettner, G.R.; Aust, S.D. Transition metals as catalysts of “autoxidation” reactions. Free. Radic. Biol. Med. 1990, 8, 95–108. [Google Scholar] [CrossRef]
  78. Yan, J.; Jiang, X.; Han, X.; Liu, J. A TG–FTIR investigation to the catalytic effect of mineral matrix in oil shale on the pyrolysis and combustion of kerogen. Fuel 2013, 104, 307–317. [Google Scholar] [CrossRef]
  79. Lewan, M.; Kotarba, M.; Więcław, D.; Piestrzyński, A. Evaluating transition-metal catalysis in gas generation from the Permian Kupferschiefer by hydrous pyrolysis. Geochim. Et Cosmochim. Acta 2008, 72, 4069–4093. [Google Scholar] [CrossRef]
  80. Gao, J.; Zou, C.; Li, W.; Ni, Y.; Yuan, Y. Transition metal catalysis in natural gas generation: Evidence from nonhydrous pyrolysis experiment. Mar. Pet. Geol. 2020, 115, 104280. [Google Scholar] [CrossRef]
  81. He, K.; Zhang, S.; Wang, X.; Mi, J.; Zhang, W.; Guo, J.; Zhang, W. Pyrolysis of 1-methylnaphthalene involving water: Effects of Fe-bearing minerals on the generation, C and H isotope fractionation of methane from H2O-hydrocarbon reaction. Org. Geochem. 2021, 153, 104151. [Google Scholar] [CrossRef]
  82. He, L.; Li, S.; Lin, W. Catalytic cracking of pyrolytic vapors of low-rank coal over limonite ore. Energy Fuels 2016, 30, 6984–6990. [Google Scholar] [CrossRef]
  83. He, L.; Hui, H.; Li, S.; Lin, W. Production of light aromatic hydrocarbons by catalytic cracking of coal pyrolysis vapors over natural iron ores. Fuel 2018, 216, 227–232. [Google Scholar] [CrossRef]
  84. Guan, D.; Ma, Z.; Zhao, Y.; Wang, H.; Wang, Q.; Liu, Q. Oil Generation Theory and Simulation Experiment Technology of Continental Source Rocks in China; Petroleum Industry Press in China: Xi’an, China, 2022; pp. 199–245. (In Chinese) [Google Scholar]
  85. Jin, Z.; Zhang, J. Exploration strategy for deep basin gas reservoirs. Pet. Explor. Dev. 1999, 26, 4–5. [Google Scholar]
  86. Mao, G.; Liu, C.; Zhang, D.; Qiu, X.; Wang, J.; Liu, B.; Liu, J.; Qu, S.; Deng, Y.; Wang, F. Effects of uranium on hydrocarbon generation of hydrocarbon source rocks with type-III kerogen. Sci. China Earth Sci. 2014, 57, 1168–1179. [Google Scholar] [CrossRef]
  87. Liao, Y.; Fang, Y.; Wu, L.; Geng, A.; Hsu, C.S. The characteristics of the biomarkers and δ13C of n-alkanes released from thermally altered solid bitumens at various maturities by catalytic hydropyrolysis. Org. Geochem. 2012, 46, 56–65. [Google Scholar] [CrossRef]
  88. Waples, D.W. Time and temperature in petroleum formation: Application of Lopatin’s method to petroleum exploration. AAPG Bull. 1980, 64, 916–926. [Google Scholar]
  89. Lewan, M.D.; Roy, S. Role of water in hydrocarbon generation from Type-I kerogen in Mahogany oil shale of the Green River Formation. Org. Geochem. 2011, 42, 31–41. [Google Scholar] [CrossRef]
  90. Shuai, Y.; Zhang, S.; Luo, P.; Liu, J.; Hu, G. Experimental evidence for formation water promoting crude oil cracking to gas. Chin. Sci. Bull. 2012, 57, 4587–4593. [Google Scholar] [CrossRef] [Green Version]
  91. Wyllie, P. Plate tectonics and magma genesis. Geol. Rundsch. 1981, 70, 128–153. [Google Scholar] [CrossRef]
  92. Mao, H.-K.; Hu, Q.; Yang, L.; Liu, J.; Kim, D.Y.; Meng, Y.; Zhang, L.; Prakapenka, V.B.; Yang, W.; Mao, W.L. When water meets iron at Earth’s core–mantle boundary. Natl. Sci. Rev. 2017, 4, 870–878. [Google Scholar] [CrossRef]
  93. Ohtani, E. The role of water in Earth’s mantle. Natl. Sci. Rev. 2020, 7, 224–232. [Google Scholar] [CrossRef]
  94. Jin, Z.; Zhang, L.; Yang, L.; Hu, W. A preliminary study of mantle-derived fluids and their effects on oil/gas generation in sedimentary basins. J. Pet. Sci. Eng. 2004, 41, 45–55. [Google Scholar] [CrossRef]
  95. Bali, E.; Audétat, A.; Keppler, H. Water and hydrogen are immiscible in Earth’s mantle. Nature 2013, 495, 220–222. [Google Scholar] [CrossRef] [PubMed]
  96. Guélard, J.; Beaumont, V.; Rouchon, V.; Guyot, F.; Pillot, D.; Jézéquel, D.; Ader, M.; Newell, K.; Deville, E. Natural H2 in K ansas: Deep or shallow origin? Geochem. Geophys. Geosystems 2017, 18, 1841–1865. [Google Scholar] [CrossRef]
  97. Holloway, J.R.; O’DAY, P.A. Production of CO2 and H2 by diking-eruptive events at mid-ocean ridges: Implications for abiotic organic synthesis and global geochemical cycling. Int. Geol. Rev. 2000, 42, 673–683. [Google Scholar] [CrossRef]
  98. Zgonnik, V. The occurrence and geoscience of natural hydrogen: A comprehensive review. Earth-Sci. Rev. 2020, 203, 103140. [Google Scholar] [CrossRef]
  99. Catling, D.C.; Zahnle, K.J.; McKay, C.P. Biogenic methane, hydrogen escape, and the irreversible oxidation of early Earth. Science 2001, 293, 839–843. [Google Scholar] [CrossRef]
  100. Andersen, T.; Neumann, E.-R. Fluid inclusions in mantle xenoliths. Lithos 2001, 55, 301–320. [Google Scholar] [CrossRef]
  101. Liu, X.; Fu, X.; Liu, D.; Wei, W.; Lu, X.; Liu, C.; Wang, W.; Gao, H. Distribution of mantle-derived CO2 gas reservoir and its relationship with basement faults in Songliao Basin, China. J. Nat. Gas Sci. Eng. 2018, 56, 593–607. [Google Scholar] [CrossRef]
  102. Walter, B.F.; Giebel, R.J.; Steele-MacInnis, M.; Marks, M.A.; Kolb, J.; Markl, G. Fluids associated with carbonatitic magmatism: A critical review and implications for carbonatite magma ascent. Earth-Sci. Rev. 2021, 215, 103509. [Google Scholar] [CrossRef]
  103. Hiett, C.D.; Newell, D.L.; Jessup, M.J.; Grambling, T.A.; Scott, B.E.; Upin, H.E. Deep CO2 and N2 emissions from Peruvian hot springs: Stable isotopic constraints on volatile cycling in a flat-slab subduction zone. Chem. Geol. 2022, 595, 120787. [Google Scholar] [CrossRef]
  104. Jørgensen, B.B.; Isaksen, M.F.; Jannasch, H.W. Bacterial sulfate reduction above 100 C in deep-sea hydrothermal vent sediments. Science 1992, 258, 1756–1757. [Google Scholar] [CrossRef] [PubMed]
  105. Fisher, A.T. Marine hydrogeology: Recent accomplishments and future opportunities. Hydrogeol. J. 2005, 13, 69–97. [Google Scholar] [CrossRef]
  106. Russell, M. First life: Billions of years ago, deep under the ocean, the pores and pockets in minerals that surrounded warm, alkaline springs catalyzed the beginning of life. Am. Sci. 2006, 94, 32–39. [Google Scholar] [CrossRef]
  107. Lartaud, F.; Little, C.T.; De Rafelis, M.; Bayon, G.; Dyment, J.; Ildefonse, B.; Gressier, V.; Fouquet, Y.; Gaill, F.; Le Bris, N. Fossil evidence for serpentinization fluids fueling chemosynthetic assemblages. Proc. Natl. Acad. Sci. USA 2011, 108, 7698–7703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Konn, C.; Charlou, J.-L.; Holm, N.G.; Mousis, O. The production of methane, hydrogen, and organic compounds in ultramafic-hosted hydrothermal vents of the Mid-Atlantic Ridge. Astrobiology 2015, 15, 381–399. [Google Scholar] [CrossRef]
  109. Baross, J.A.; Hoffman, S.E. Submarine hydrothermal vents and associated gradient environments as sites for the origin and evolution of life. Orig. Life Evol. Biosph. 1985, 15, 327–345. [Google Scholar] [CrossRef]
  110. Ménez, B.; Pisapia, C.; Andreani, M.; Jamme, F.; Vanbellingen, Q.P.; Brunelle, A.; Richard, L.; Dumas, P.; Réfrégiers, M. Abiotic synthesis of amino acids in the recesses of the oceanic lithosphere. Nature 2018, 564, 59–63. [Google Scholar] [CrossRef]
  111. Reeves, E.P.; Fiebig, J. Abiotic synthesis of methane and organic compounds in Earth’s lithosphere. Elem. Int. Mag. Mineral. Geochem. Petrol. 2020, 16, 25–31. [Google Scholar] [CrossRef]
  112. Beaudry, P.; Stefánsson, A.; Fiebig, J.; Rhim, J.H.; Ono, S. High temperature generation and equilibration of methane in terrestrial geothermal systems: Evidence from clumped isotopologues. Geochim. Et Cosmochim. Acta 2021, 309, 209–234. [Google Scholar] [CrossRef]
  113. Coveney, R.M., Jr.; Goebel, E.D.; Zeller, E.J.; Dreschhoff, G.A.; Angino, E.E. Serpentinization and the origin of hydrogen gas in Kansas. AAPG Bull. 1987, 71, 39–48. [Google Scholar]
  114. Lollar, B.S.; Frape, S.; Weise, S.; Fritz, P.; Macko, S.; Welhan, J. Abiogenic methanogenesis in crystalline rocks. Geochim. Et Cosmochim. Acta 1993, 57, 5087–5097. [Google Scholar] [CrossRef]
  115. Klein, F.; Grozeva, N.G.; Seewald, J.S. Abiotic methane synthesis and serpentinization in olivine-hosted fluid inclusions. Proc. Natl. Acad. Sci. USA 2019, 116, 17666–17672. [Google Scholar] [CrossRef] [Green Version]
  116. Galland, O.; Holohan, E.; Van Wyk de Vries, B.; Burchardt, S. Laboratory modelling of volcano plumbing systems: A review. In Physical Geology of Shallow Magmatic Systems: Dykes, Sills and Laccoliths; Springer-Verlag: Berlin, Germany, 2018; pp. 147–214. [Google Scholar]
  117. Zahedi, M.K.; MacDonald, S. Volcanics: A Commonly Underestimated part of Petroleum Exploration. ASEG Ext. Abstr. 2018, 2018, 1–6. [Google Scholar] [CrossRef] [Green Version]
  118. Dincer, I. Green methods for hydrogen production. Int. J. Hydrog. Energy 2012, 37, 1954–1971. [Google Scholar] [CrossRef]
  119. Landais, P. Organic geochemistry of sedimentary uranium ore deposits. Ore Geol. Rev. 1996, 11, 33–51. [Google Scholar] [CrossRef]
  120. Fisher, W.L.; Galloway, W.; Proctor, C.V., Jr.; Nagle, J. Depositional Systems in the Jackson Group of Texas their Relationship to Oil Gas, and Uranium; The University of Texas at Austin, Bureau of Economic Geology: Austin, TX, USA, 1970. [Google Scholar]
  121. Lecomte, A.; Michels, R.; Cathelineau, M.; Morlot, C.; Brouand, M.; Flotté, N. Uranium deposits of Franceville basin (Gabon): Role of organic matter and oil cracking on uranium mineralization. Ore Geol. Rev. 2020, 123, 103579. [Google Scholar] [CrossRef]
  122. Gauthier-Lafaye, F.; Weber, F. Natural nuclear fission reactors: Time constraints for occurrence, and their relation to uranium and manganese deposits and to the evolution of the atmosphere. Precambrian Res. 2003, 120, 81–100. [Google Scholar] [CrossRef]
  123. Robertson, J.; Hendry, M.J.; Kotzer, T.; Hughes, K.A. Geochemistry of uranium mill tailings in the Athabasca Basin, Saskatchewan, Canada: A review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 1237–1293. [Google Scholar] [CrossRef]
  124. Kang, H.; Chen, Y.-l.; Li, D.-p.; Zhao, J.-x.; Cui, F.-r.; Xu, Y.-l. Deep-penetrating geochemistry for concealed sandstone-type uranium deposits: A case study of Hadatu uranium deposit in the Erenhot Basin, North China. J. Geochem. Explor. 2020, 211, 106464. [Google Scholar] [CrossRef]
  125. Schulz, H.-M.; Yang, S.; Schovsbo, N.H.; Rybacki, E.; Ghanizadeh, A.; Bernard, S.; Mahlstedt, N.; Krüger, M.; Amann-Hildebrandt, A.; Krooss, B.M. The Furongian to Lower Ordovician Alum Shale Formation in conventional and unconventional petroleum systems in the Baltic Basin–A review. Earth-Sci. Rev. 2021, 218, 103674. [Google Scholar] [CrossRef]
  126. Nakashima, S.; Disnar, J.; Perruchot, A.; Trichet, J. Experimental study of mechanisms of fixation and reduction of uranium by sedimentary organic matter under diagenetic or hydrothermal conditions. Geochim. Et Cosmochim. Acta 1984, 48, 2321–2329. [Google Scholar] [CrossRef]
  127. Partin, C.A.; Bekker, A.; Planavsky, N.J.; Scott, C.; Gill, B.C.; Li, C.; Podkovyrov, V.; Maslov, A.; Konhauser, K.O.; Lalonde, S.V. Large-scale fluctuations in Precambrian atmospheric and oceanic oxygen levels from the record of U in shales. Earth Planet. Sci. Lett. 2013, 369, 284–293. [Google Scholar] [CrossRef]
  128. Lovley, D.R.; Phillips, E.J. Reduction of uranium by Desulfovibrio desulfuricans. Appl. Environ. Microbiol. 1992, 58, 850–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Cai, C.; Dong, H.; Li, H.; Xiao, X.; Ou, G.; Zhang, C. Mineralogical and geochemical evidence for coupled bacterial uranium mineralization and hydrocarbon oxidation in the Shashagetai deposit, NW China. Chem. Geol. 2007, 236, 167–179. [Google Scholar] [CrossRef]
  130. Seewald, J.S. Organic–inorganic interactions in petroleum-producing sedimentary basins. Nature 2003, 426, 327–333. [Google Scholar] [CrossRef]
  131. Cai, Y.-W.; Zhang, S.-C.; He, K.; Mi, J.-K.; Zhang, W.-L.; Wang, X.-M.; Wang, H.-J.; Wu, C.-D. Effects of U-ore on the chemical and isotopic composition of products of hydrous pyrolysis of organic matter. Pet. Sci. 2017, 14, 315–329. [Google Scholar] [CrossRef] [Green Version]
  132. Yang, S.; Schulz, H.-M.; Horsfield, B.; Schovsbo, N.H.; Noah, M.; Panova, E.; Rothe, H.; Hahne, K. On the changing petroleum generation properties of Alum Shale over geological time caused by uranium irradiation. Geochim. Et Cosmochim. Acta 2018, 229, 20–35. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The main catalytic mechanism of clay minerals.
Figure 1. The main catalytic mechanism of clay minerals.
Minerals 13 00515 g001
Figure 2. Variation diagram of pyrolysis yield of increasing Ni (Adapted from [80]).
Figure 2. Variation diagram of pyrolysis yield of increasing Ni (Adapted from [80]).
Minerals 13 00515 g002
Figure 3. Comparison diagram of different forms of Mn element addition on oil production rate in thermal simulation experiment (adapted from [84]).
Figure 3. Comparison diagram of different forms of Mn element addition on oil production rate in thermal simulation experiment (adapted from [84]).
Minerals 13 00515 g003
Figure 4. Influence model of volcanic activity on sedimentary basins (Ref. [85]).
Figure 4. Influence model of volcanic activity on sedimentary basins (Ref. [85]).
Minerals 13 00515 g004
Figure 5. Correlation between uranium and TOC (quoted from Schulz et al. [125]).
Figure 5. Correlation between uranium and TOC (quoted from Schulz et al. [125]).
Minerals 13 00515 g005
Table 1. Primary metabolic reactions of various microorganisms near hydrothermal solution (cited from Martin et al. [76]).
Table 1. Primary metabolic reactions of various microorganisms near hydrothermal solution (cited from Martin et al. [76]).
MetabolismReactionExamples in Vent Environments
AnaerobicMethanogenesis 4 H 2 + C O 2 C H 4 + 2 H 2 O
  C H 3 C O 2 + H 2 O C H 4 + H C O 3
  4 H C O O + H + 3 H C O 3 + C H 4
Methanococcus spp. common in magma-hosted vents; Methanosarcinales at Lost City
S° reduction   S 0 + H 2 H 2 S Lithotrophic and heterotrophic; hyperthermophilic archaea
Anaerobic CH4 oxidation   C H 4 + S O 4 2 H S + H C O 3 + 2 H 2 O Methanosarcina spp. and epsilonproteobacteria at mudvolcanoes and methane seeps
Sulfate reduction   S O 4 2 + H + + 4 H 2 H S + 4 H 2 O Deltaproteobacteria
Fe reduction   8 F e 3 + + C H 3 C O 2 + 4 H 2 O 2 H C O 3 + 8 F e 2 + + 9 H + Epsilonproteobacteria, thermophilic bacteria and hyperthermophilic Crenarchaeota
Fermentation   C 6 H 12 O 6 2 C 6 H 6 O + 2 C O 2 Many genera of bacteria and archaea
AerobicSulfide oxidation   H S + 2 O 2 S O 4 2 + H + Many genera of bacteria; common vent animal symbionts
CH4 oxidation   C H 4 + 2 O 2 H C O 3 + H + + H 2 O Common in hydrothermal systems; vent animal symbionts
H2 oxidation   H 2 + 0.5 O 2 H 2 O Common in hydrothermal systems; vent animal symbionts
Fe oxidation   F e 2 + + 0.5 O 2 + H + F e 3 + + 0.5 H 2 O Common in low-temperature vent fluids; rock-hosted microbial mats
Mn oxidation   M n 2 + + 0.5 O 2 + H 2 O M n O 2 + 2 H + Common in low-temperature vent fluids; rock-hosted microbial mats; hydrothermal plumes
Respiration   C 6 H 12 O 6 + 6 O 2 6 C O 2 + 6 H 2 O Many genera of bacteria
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

Zhong, M.; Huang, H.; Xu, P.; Hu, J. Catalysis of Minerals in Pyrolysis Experiments. Minerals 2023, 13, 515. https://0-doi-org.brum.beds.ac.uk/10.3390/min13040515

AMA Style

Zhong M, Huang H, Xu P, Hu J. Catalysis of Minerals in Pyrolysis Experiments. Minerals. 2023; 13(4):515. https://0-doi-org.brum.beds.ac.uk/10.3390/min13040515

Chicago/Turabian Style

Zhong, Ming, Haiping Huang, Pengcheng Xu, and Jie Hu. 2023. "Catalysis of Minerals in Pyrolysis Experiments" Minerals 13, no. 4: 515. https://0-doi-org.brum.beds.ac.uk/10.3390/min13040515

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