Next Article in Journal
High-Pressure Experimental and DFT Structural Studies of Aurichalcite Mineral
Previous Article in Journal
Isotopic Signatures of Microbial Mg-Carbonates Deposited in an Ephemeral Hyperalkaline Lake (Central Spain): Paleoenvironmental Implications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Solid Hydrocarbon on the Composition of Fluid Phase at the Subduction (Experimental Simulation)

V.S. Sobolev Institute of Geology and Mineralogy, Siberian Branch of Russian Academy of Sciences, 630090 Novosibirsk, Russia
*
Authors to whom correspondence should be addressed.
Minerals 2023, 13(5), 618; https://doi.org/10.3390/min13050618
Submission received: 9 March 2023 / Revised: 15 April 2023 / Accepted: 26 April 2023 / Published: 28 April 2023

Abstract

:
Experiments conducted in the olivine–serpentine–anthracene–metal (FeNi) system have shown that the recrystallization of olivines occurs under substantially reduced conditions with active participation of hydrocarbons, especially paraffins; moreover, their amount increases with increasing pressure and temperature. During the decomposition of serpentine, a large amount of water is released; therefore, the fluid at relatively low P-T parameters (2 GPa, 1100 °C) has mainly water–hydrocarbon composition. With an increase in pressure up to 3–4.5 GPa and temperature up to 1300–1400 °C, the composition of the fluid changes greatly towards an increase in the relative amount of hydrocarbons, while the main share is occupied by light (C1–C4) aliphatic hydrocarbons. Therefore, a biogenic material with a carbon–hydrogen composition can make a certain contribution to the carbon budget in subduction processes when falling into the subduction zones and may affect the oxygen fugitivity in the subducted slab.

1. Introduction

According to modern concepts, during the subduction of the oceanic crust, carbon mainly enters the mantle as carbonate minerals (initially CaCO3) or in a dissolved state in fluid or melt [1,2,3,4,5,6,7]. An oxidized fluid regime is assumed in this process (predominantly CO2-H2O composition) [8], although more complex oxygenated compounds can theoretically be present, such as acetic acid and acetate [9].
However, a slab can contain graphite in a carbon phase, which depends on rock mineral composition and, accordingly, on oxygen fugitivity [10,11,12,13]. Graphite may appear due to graphitization of carbonaceous material of biogenic origin from sedimentary rocks. Experimental data show that amorphized carbon from sedimentary rocks begins to crystallize at pressures of 0.5–3.0 GPa and temperatures of 350–420 °C and is completely transformed into well-structured graphite in the range of 450–600 °C [14]. Carbon-bearing material of biogenic origin contributes significantly to the carbon budget during subduction [15,16]. Solid hydrocarbons of abiogenic origin at a pressure of up to 8 GPa also decompose at relatively low temperatures (no more than 700 °C) into solid amorphized carbon and a fluid phase [17]. This was established experimentally in the examples of naphthalene (C10H8), anthracene (C14H10), pyrene (C16H10), and coronene (C24H12). In this case, the fluid is highly reduced (H2-CH4 prevail) [18,19]. Light hydrocarbons were identified in natural sites in fluid inclusions in minerals of rocks experienced metamorphism, metasomatosis, and hydrothermal changes [20,21,22,23]. The said hydrocarbons were identified using gas chromatography and Raman spectroscopy.
However, the actual composition of the fluid may be more complex. For example, hydrothermal fluids near the white smoker of Lost City on the MAR (Mid-Atlantic Ridge) were found to contain n-alkane acids (C8-C12) [24]. Ménez et al. [25] reported that samples lifted from a depth of 170 m below the sea floor during drilling near the smoker of Lost City contained a nitrogenated amino acid (aromatic alpha-amino acid C11H12N2O2) and other organic compounds. Pyrolysis-free gas chromatography–mass spectrometry (GC-MS) makes it possible to analyse the complex composition of the fluid in natural objects. Thus, such compounds as alkanes, alkenes, arenes, phenols, aldehydes, carboxylic acids, esters, ketones, nitriles, PAH (polycyclic aromatic hydrocarbon), and their halogenated, methylated, and sulfonated derivatives were found in volcanic gases from Vulcano (Aeolian Islands, Italy), as well as various heterocyclic compounds including thiophenes and furans [26]. Hydrocarbon compounds, including high-molecular ones, and their oxygenated, nitrogenated, sulfonated, and halogenated derivatives were diagnosed by GC-MS in fluid inclusions in natural diamonds [27,28,29,30], in melt and fluid inclusions in phenocrysts from basalts and rhyolites from Men’shii Brat Volcano (Medvezh’ya caldera, Iturup Island) [31], and fluid inclusions in minerals from polymetallic sulfide deposits at Ashadze-1, Semenov-2, Krasnov, and Rainbow, and from the Lost City carbonate structures of hydrothermal fields in the Mid-Atlantic Ridge [32]. Stability of high molecular weight carbon compounds is also proved experimentally at high P-T parameters (at 5.3 Gpa and 1300 °C in the Fe-C-S system) [33].
During subduction, a temperature increase causes dehydration of hydrous minerals in the oceanic crust [34]. It is serpentine that is believed to be the main source of water in the mantle, since it contains about 13 wt. % of water in the form of a hydroxyl group. According to experimental data, serpentinization of rocks of the peridotite association in the lithospheric mantle can occur at as low as 400–500 °C and low pressure as a result of hydrothermal transformation [35]. Above 700 °C, serpentine becomes unstable (at high pressure) and decomposes into olivine and orthopyroxene while releasing water [36,37]. The released water is concentrated in interstices between newly formed minerals and in fluid inclusions in these [38]. It was found that 0.13–2.43 wt. % of the fluid phase is preserved in newly formed solid phases in the form of fluid inclusions. The proportion of water proper was in the range of 0.1–2.06 wt. % according to chromatographic analysis [39]. High quality of fluid phase is assumed when olivine-bearing rocks are immersed in the mantle. Thus, H2O could reach the depths of the reduced mantle during its subduction with oxygen fugacity corresponding to the stability of metallic iron [40,41,42].
In connection with the above, it is of great scientific interest to specify the composition of fluid under reducing conditions in the presence of carbon and hydrogen in the system. We continued research on this topic considering the aforesaid information, especially the first results of studying the fluid phase released during anthracene decomposition in the presence of olivine and FeNi metal [43]. This publication provides experimental data on the study of the fluid phase composition in the system of metal FeNi–olivine–anthracene with the addition of serpentine under high pressure and temperature conditions simulating the stage of the oceanic crust immersing into the mantle.

2. Materials and Methods

2.1. Experiments

The experiments were carried out on a high-pressure, high-temperature apparatus of the “split sphere” type (BARS) according to the method developed following the state task at the IGM SB RAS (Novosibirsk). High-pressure cells (HPCs) were manufactured on a basis of refractory oxide ZrO2. The HPC heating circuit included a tubular graphite heater and molybdenum contacts. Pressure was calibrated by phase transitions in reference substances Bi and PbSe before heating the reaction volume [44]. Temperature was measured by a PtRh30/PtRh6 thermal couple installed directly near the reaction volume in the HPCs. The pressure and temperature measurement error in the experiments were ±0.2 GPa and ±25 °C, respectively. There were three experiments: 1100 °C and 2.0 GPa (No. 4-36-21), 1300 °C and 3.0 GPa (4-37-21), and 1400 °C and 4.5 GPa (4-38-21). The duration of all experiments was the same: 5 h.
The source substances were: Fe-Ni alloy, i.e., invar (36 wt. % Ni), natural olivine (fraction of 0.5–1 mm) of lherzolite xenoliths of Mongolian basalts, natural serpentine powder from Eastern Sayan ophiolites (Russia), and anthracene powder (ultra-high purity). Chemical composition of olivine (wt. %) was SiO2-40.47; FeO-9.00; MgO-49.62; CaO-0.04; TiO2-0.01; Cr2O3-0.04; MnO-0.14; NiO-0.41. Chemical composition of serpentine was SiO2-41.53; MgO-42.15; Fe2O3-2.74; Al2O3-0.95; TiO2-0.02; MnO-0.14; CaO-0.05; Na2O-0.30; K2O-0.02; P2O5-0.00; LOI of 12.42; LOI of 12.42; total of 100.32. The HPC reaction volume was filled with a mixture of olivine (90 wt. %) with anthracene (5 wt. %) and serpentine (5 wt. %); Fe-Ni alloy shaped as a disk was placed on the top. The samples of the source components were placed in capsules made of pressed MgO powder.
The assembly scheme of the HPC reaction volume is shown in Figure 1. When quenching the sample by disconnecting the heater power supply, the cooling rate of the samples was 2–3 s until reaching the metal solidification temperature; then the temperature reached 100–200 °C within a few seconds due to effective water cooling of the internal block of the hard-alloy punches of the device. The data on assembling a high-pressure cell and the procedures for conducting experiments are described in more detail in [43,45].
The compositions of gas and crystal phases in the fluid inclusions in the newly formed olivines were studied using a scanning electron microscope and by the methods of Raman spectroscopy and pyrolysis-free gas chromatography-mass spectrometry (GC-MS).

2.2. Scanning Electron Microscopy

After the experiments, the samples were studied using the methods of optical and scanning electron microscopy. The chemical composition of the newly formed mineral phases was studied using an electronic microscope MIRA 3 LMU (“TESCAN” Orsay Holding) equipped with the INCA Energy 450 + Xmax80 microanalysis system (Oxford Instruments Nanoanalysis Ltd., Oxford, UK) and the JEOL JXA-8100 electron probe microanalyser equipped with five wave spectrometers at the Research Equipment-Sharing Center of the IGM SB RAS. Measurements were taken at an accelerating voltage of 15–20 kV, a probe current of 20–50 nA, and a counting time of 10–60 s. A probe with a minimum beam diameter of about 3 microns was used to determine the composition of homogeneous phases.

2.3. Raman Spectroscopy Analysis

The compositions of gas and crystal phases in fluid inclusions were analysed using Raman spectroscopy on a Horiba Lab Ram HR 800 spectrometer in the Institute of Geology and Mineralogy SB RAS, Novosibirsk. The Raman signal was excited using a solid state Nd YAG laser with a wavelength 532 nm and power 75 mW. The spectra were registered using a semiconductor Endor detector with Peltier cooling. A confocal spectrometer based on the OLYMPUS BX-41microscope with a 100× lens with a large numerical aperture was used to locate the point in the analysed sample. The analysis was performed in a backscattering geometry. The time of signal accumulation and the size of confocal aperture varied depending on the size of the analysed phase. The spectra were obtained in the range of 100–4200 cm−1. The time of signal accumulation ranged from 25 s/spectral window for large objects and to 400 s/spectral window for small objects. The measurement error was within 1 cm−1.

2.4. GC–MS Analysis

Volatiles from the samples were analysed using the pyrolysis-free coupled gas chromatography–mass spectrometry (GC–MS) method on a Focus GC/DSQ II Series Single Quadrupole MS analyser (Termo Scientifc, Austin, TX, USA) at the V.S. Sobolev Institute of Geology and Mineralogy, Siberian Branch of the Russian Academy of Sciences, Novosibirsk [32,46,47,48,49]. The gas mixture was released from the fluid and melt inclusions of the samples by means of shock mechanical crushing in a custom-designed crusher. The crusher was heated to 160 °C and flushed with He to remove adsorbed volatiles. The released mixture was entrained in a He stream, without cryogenic focusing. Each analytical run was preceded and followed by blank analyses, which later were used in data processing. The gas mixture was injected into the analytical column of the GC–MS instrument through a 6-port 2-position Valco (USA) valve thermostated at 290 °C, at a constant He flow rate of 1.7 mL min−1, using vacuum compensation. The GC–MS transferline temperature was held at 300 °C. The gas mixture was separated in a Restek Rt-Q-BOND capillary column (100% divinylbenzene used as a stationary phase; length of 30 m; inner diameter of 0.32 mm; film thickness of 10 µm). The temperature program of the GC separation comprised an isothermal stage (70 °C for 2 min), followed by two heating ramps (25 °C min−1 to 150 °C and 5 °C min−1 to 290 °C), followed by the final isothermal stage at 290 °C for 100 min. Total ion current (TIC) electron ionization spectra were collected on a quadrupole mass-selective detector in full scan mode at an electron energy and emission current of 70 eV and 100 µA, respectively. Other experimental parameters were as follows: ion source T = 200 °C; multiplier voltage 1500 V; positive ion detection; the mass range of 5 to 500 amu; scan rate 1 s−1; and scan rate of 506.6 amu s−1. The start time of the analysis was synchronized with the samples’ shock crushing.
The procedure for preparing the sample for analysis excluded contact with any solvents and other possible contamination. The input of the mixture extracted from the sample during the shock crushing was carried out online in the He flow without concentration including cryofocus. This method does not pyrolyze the sample but heats it only in order to convert any water within the sample into a gas phase. In this case, it is the gas mixture that is analysed in situ rather than pyrolyzate, which contains more oxidized compounds (H2O, CO, CO2, etc.) due to the reactions between the gas mixture compounds, the gas mixture and accumulator surface, and the gas phase compounds and the sample. Blank online analyses were carried out before and after the ‘‘working” analysis. The previous analysis made it possible to control the release of gases adsorbed by the sample surface, including atmospheric components, and to record the system blank at the end of the process. The degree and completeness of hydrocarbon and polycyclic aromatic hydrocarbon elution from the analytical column during temperature programming in a chromatograph thermostat were determined using the results of subsequent analysis. If necessary, the analytical column was thermo-conditioned to achieve the required blank. The collected spectra were interpreted using both the AMDIS 2.73 (Automated Mass Spectral Deconvolution and Identification System) software and manually, with background correction against spectra from the NIST2020 and Wiley Registry 12th Edition Mass Spectral libraries (NIST MS Search 2.4). Peak areas in TIC chromatograms were determined using the ICIS algorithm Xcalibur (1.4SR1 Qual Browser). This method is suitable for detecting trace volatile concentrations exceeding tens of femtograms. The relative concentrations (rel. %) of volatile components in the studied mixture were obtained by normalizing the areas of individual chromatographic peaks to the total area of all peaks. The reliability of this normalization method was verified using external standards. Namely, certified Scotty Inc. NL 34522-PI and 34525-PI gas standards of methane–hexane alkanes were injected into the gas stream in splitless mode by means of a volumetric gas-tight syringe or a special valve with replaceable loops for volumes ranging from 2 to 500 µL. The calibration quality was assessed using the coefficients of determination R2 of the relationships between the peak area and the injected amount. The respective R2 values were as follows: 0.9975 (16 m/z, n = 22) for methane, 0.9963 (26 + 30 m/z, n = 16) for ethane, 0.9986 (29 + 43 m/z, n = 15) for propane, 0.9994 (29 + 43 m/z, n = 17) for butane, 0.9935 (43 + 72 m/z, n = 6) for pentane, and 0.9909 (57 + 86 m/z, n = 5) for hexane. The concentration ranges of alkanes during calibration were similar to concentrations encountered in the experiments. The relative analytical uncertainty for C1-C6 alkane determination was below 5 % (2σ).

3. Results

The source olivine was recrystallized in all experiments with the participation of the fluid phase. Fluid inclusions were found in the crystals of newly formed olivine (Figure 2 and Figure 3). According to the results of scanning electron microscopy and X-ray spectral microanalysis, the composition of the newly formed olivines (in wt. %) from the experiment 4-36-21 (T = 1100 °C and p = 2.0 GPa) was as follows: SiO2–41.0–42.21; FeO–2.06–7.22; MnO–0.0; MgO–51.05–55.54; CaO–0.0; NiO–0.27–0.53; Total—99.80–100.07, Fo = 91–99; olivines from the experiment 4-37-21 (T = 1300 °C and p = 3.0 GPA) were made of SiO2–40.66–40.93; FeO–4.81–8.10; MnO–0.0–0.15; MgO–50.39–52.25; CaO–0.0; NiO–0.26–0.41; Total—98.40–99.85, Fo = 92–97; and olivines from the experiment 4-38-21(T = 1400 °C and P = 4.5 GPa) contained SiO2–40.65–42.0; FeO–2-38-9.97; MnO–0.0–0.15; MgO–48.94–55.38; CaO–0.0; NiO–0.20–0.23; Total—99.94–100.0, Fo = 97.
Raman spectroscopy showed methane (Raman lines 2911 and 2912 cm−1), nitrogen (Raman line 2329 cm−1), hydrogen (Raman lines 4127, 4144, 4155, and 4162 cm−1) and hydrocarbons heavier than methane (Raman lines 2873, 2934, and 2970 cm−1) (Figure 4).
In addition, brucite (Raman lines 278, 442, 3645, and 3652 cm−1), calcite (Raman lines 152, 279, 712, 1086, and 1436 cm−1), and a non-identified solid phase (Raman lines 640 and 1114 cm−1) were found in fluid inclusions in olivine (Figure 5, Figure 6 and Figure 7). It is significant that carbonates were found in fluid inclusions (wt. %: SiO2–0.79; FeO–0.73; MnO–0.0; MgO–1.38; CaO–51.84; NiO–0.0; Total—54.75) (Figure 5 and Figure 6), although they were not included in the initial charge. Most likely they were formed directly in the fluid inclusions during the quenching and cooling of the experiment, just as in brucite. It is interesting to note that water in the form of OH-groups was detected in all newly formed olivines according to Raman spectroscopy data (Figure 7).
In addition to hydrocarbons and carbon dioxide, graphite is also present as a carbon phase. Graphite in the experimental products is present both as amorphized graphite mainly in fluid inclusions and as well-structured graphite both in fluid inclusions and in the form of solid inclusions in olivine (Raman lines 1579, 1580, and 1608 cm−1) (Figure 8a,b). What is more, amorphized graphite in fluid inclusions could be formed most likely from the fluid phase while the system cooled after the experiment.
According to GC-MS analysis, 213 to 248 volatile compounds were found in olivine inclusions obtained in the olivine–serpentine–anthracene–metal FeNi system at various temperatures and pressures (Table 1; Supplementary Tables S1–S3; Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13). The studies also showed that the major volatile components in the composition of fluid inclusions in olivines are hydrocarbons and their derivatives: aliphatic (paraffins, olefins), cyclic (naphthenes, arenes, PAHs), oxygenated (alcohols and esters, aldehydes, ketones, carboxylic acids), and heterocyclic compounds (dioxanes, furans) (Table 1; Supplementary Tables S1–S3; Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13).
The total content of hydrocarbons and their derivatives differs significantly in olivines from different experiments. The largest amount of hydrocarbons was found for olivines from experiment 4-38-21 and is 88.4 rel. %, and the lowest one is 37.6 rel. % for olivines from experiment 4-36-21 (Table 1; Figure 12d and Figure 13). It should be noted that olivines from experiment 4-38-21 showed the highest content of aliphatic hydrocarbons (82.9 rel. %) and the lowest content of oxygenated hydrocarbons (4.8 rel. %) compared with olivines from experiments 4-36-21 and 4-37-21, which contained 26.4 and 60.4 rel. % and 10.3 and 11.1 rel. %, respectively. (Table 1; Figure 12a).
The content of light saturated hydrocarbons (methane CH4 to n-butane C4H10) in olivines from experiment 4-38-21 (82.6 rel. %) is significantly higher (as compared with olivines from experiments 4-36-21 and 4-37-21, which were 25.6 and 60.1 rel. %, respectively (Figure 12b). The content of medium saturated hydrocarbons (n-pentane C5H12 to n-dodecane C12H26) in olivines from all experiments does not differ significantly and amounts to 0.6, 0.5, and 1.0 rel. %, respectively (Figure 12b). The content of heavy saturated hydrocarbons (n-pentane C5H12 to n-dodecane C12H26) for olivines from experiments 4-37-21 and 4-38-21 is close, amounting to 0.3 and 0.2 rel. %, respectively. Olivines from experiment 4-36-21 have a slightly larger share of heavy hydrocarbons, amounting to 1.2 rel. %.
The total content of cyclic hydrocarbons (C5H8-C16H26) in inclusions in olivines from all three experiments is quite close and amounts to 0.72, 0.75, and 0.6 rel. % for experiments 4-36-21 (at 1100 °C and at 2.0 GPa), 4-37-21 (at 1300 °C and at 3.0 GPa), 4-38-21 (at 1400 °C and at 4.5 GPa), respectively. Olivines from experiment 4-37-21 showed the highest content of oxygenated hydrocarbons (alcohols and esters CH4O-C14H18O4, aldehydes C2H4O-C15H30O, ketones C3H6O-C15H30O, and carboxylic acids C2H4O2-C14H28O2) amounting to 11.1 rel. %. The lowest content of the above was noted for olivines from experiment 4-38-21 being 4.8 rel. %.
According to the calculated H/(O+H) ratios (0.82, 0.97, and 0.98) (Table 1), the recrystallization of olivines occurred under more reduced conditions as the temperature in the experiments increased from 1100 to 1400 °C and the pressure from 2.0 to 4.5 GPa. The greatest reduction conditions [N/(O+H) = 0.98)] are characteristic of olivines obtained in experiment 4-38-21 at 1400 °C and 4.5 GPa (Table 1).
It should be noted that the water content in the inclusions in olivines from experiment 4-36-21 (T = 1100 °C and p = 2.0 GPa) is quite high and amounts to 47.0 % rel. At the same time, the increase in experiments 4-37-21 and 4-38-21 of temperature and pressure to 1300 °C and 1400 °C and 3.0 and 4.5 GPa, respectively, led to a dramatic decrease in the water content to 4.5 and 4.4 rel. %, respectively (Table 1; Figure 7d and Figure 8). However, at the same time, the presence of water in olivines in the form of OH-groups is observed in all experiments. CO2 content in the inclusions in olivines from all the experiments is very insignificant and amounts to 0.4, 0.5, and 0.3 rel. %, respectively (Table 1; Figure 12d and Figure 13).
Sulfonated compounds (H2S-C12H26O3S) were found in the inclusions of all studied olivines (0.2–1.0 rel. %): hydrogen sulfide (H2S), sulfur dioxide (SO2), sulfur carbonyl (COS), carbon disulfide (CS2), dimethyl disulfide (C2H6S2), methanethiol (CH4S) and various thiophenes (thiophene C4H4S—nonylthiophene C13H22S) (Supplementary Tables S1–S3). Sulfur was initially present in insignificant amounts in natural serpentine in the form of sulfides.
Nitrogen in fluid inclusions was present both in molecular form (N2) and as nitrogenated compounds. It was established that there were from 21 to 24 nitrogenated compounds (from acetonitrile C2H3N to decanamide C10H21NO) (Supplementary Tables S1–S3). Nitrogen was present in the reaction volume as an adsorption impurity captured during the HPC assembly.
Thus, according to the conducted studies, the fluid phase in the metal FeNi–olivine–anthracene system with the addition of serpentine under high pressure and temperature conditions is composed of a complex mixture of components where not only water, carbon dioxide, and hydrogen were found, but also hydrocarbons and their derivatives (aliphatic, cyclic and oxygenated) as well as nitrogenated and sulfonated compounds. Moreover, with an increase in temperature from 1100 to 1400 °C and pressure from 2.0 to 4.5 GPa, the content of hydrocarbons in the system became overpowering (88.4 rel.%).

4. Discussion

Experiments conducted in the olivine–serpentine–anthracene–metal (FeNi) system have shown that the recrystallization of olivines occurs under substantially reduced conditions with active involvement of hydrocarbons, especially paraffins; moreover, their amount increases with increasing pressure and temperature. The results obtained in this study confirm the data on the stability of hydrocarbon compounds (up to C18) in fluid that forms during the decomposition of anthracene at 3 GPa and 1500 °C in the presence of olivine and FeNi melt [43]. Moreover, when an additional amount of hydrogen was introduced into the system in the form of an OH-group (from serpentine), that led to a significant increase in the proportion of paraffin hydrocarbons, mainly of light components (C1-C4). A very high content of H2O at 1100 °C and 2.0 GPa (experiment 4-36-21) is also noteworthy. At higher P-T parameters, the amount of H2O decreases dramatically with an increase in the proportion of hydrocarbon compounds. The relative amount of CO2 is rather insignificant (0.3–0.5 rel. %); that is, carbon in the fluid is concentrated in organic compounds.
At 1400 °C and 4.5 GPa (experiment 4-38-21), there was a decrease identified in the amount of oxygenated and nitrogenated components in fluid inclusions in olivines relative to other experiments in the series conducted at lower temperatures and pressures. This effect is probably related to the dissolution of oxygen and nitrogen in the partially molten metal FeNi. At normal pressure (0.1 MPa), the melting point of Fe is 1538 °C; for Ni, 1455 °C, and for FeNi alloy (36 wt. %), 1455 °C [50]. However, the eutectic temperature in the Fe–Ni–C system is quite low, being about 1077 °C at 4.7 GPa [51]. Therefore, when carbon is dissolved, a melt of Fe–Ni–C should appear on contact with the FeNi alloy. The higher the temperature, the more carbon is able to dissolve and the larger the volume of molten metal, respectively.
Despite the presence of medium and heavy hydrocarbons (up to C16H34), the main components in the fluid of the graphite–serpentine–metallic iron system (without the initial olivine and solid hydrocarbons) at 2–4 GPa and 1200 °C are light hydrocarbons (methane -butane) and inorganic gases H2O, CO2 [46].
Decomposition of serpentine leads to the release of H2O, and thus the fluid is oxidized [39,52]. However, this is the case when other redox-sensitive components are absent. Despite the main trend of the oxidized state of fluid during serpentine dehydration in the slab, the presence of sulfides and especially avaruite, on the other hand, indicates the possibility of relatively low oxygen fugitivity [8,18]; that is, there is a high heterogeneity in the redox state.
Metamorphosed sedimentary rocks falling into subduction zones contain kerogen, an organic polymer material consisting mainly of hydrogen and carbon. Organic hydrocarbon compounds can make a significant contribution to fluid balance during subduction transformations of rocks [53,54,55]. Thus, the decomposition of kerogen should introduce significant portions of hydrocarbon compounds into the component composition of the fluid already at the early stages of the oceanic crust subduction. Moreover, this phenomenon should increase with increasing pressure and temperature, especially when reaching the stability conditions of the metal phases.
The interaction of aqueous fluids with ultrabasic rocks generates a large amount of hydrogen already at the stage of the oceanic lithosphere hydrothermal transformation (serpenitinization). Hydrogen release during serpentinization of ultrabasic rock may occur in accordance with the following reaction [56]:
olivine + H2O → serpentine + brucite + magnetite +H2
or [57]:
olivine ± orthopyroxene + H2O → serpentine ± brucite + magnetite +H2
FTT (Fischer–Tropsch type) reactions are believed to be the main mechanism for the formation of methane and light hydrocarbons [56,58]. This process consists of surface-catalyzed reduction of CO2 and CO by H2. Light hydrocarbons are formed in several stages on the surface of the catalyst by substituting a carbonyl unit (-CO) with carbide (-C), methylene (-CH2), and methyl (-CH3) groups. Catalysts can be compounds of iron (sulfides or oxides) and other transition metals, but FeNi alloy is the most active catalyst of the FTT process [57]. Another possibility considered for generating CH4 is a Sabatier type reaction [56,58]. The appearance of higher molecular weight hydrocarbon compounds is due to the polymerization of CH4 in accordance with the following reaction [56]:
nCH4 → CnHn+2 + (n − 1)H2
or polycondensation on catalysts with the formation of different classes of hydrocarbons in accordance with [59]:
nCO + (2n + 1)H2 → CnH2n+2 + nH2O
Fe compounds are active catalysts in the polycondensation reaction [60].
The appearance of brucite in inclusions in olivines is associated with cooling of samples, since it is stable at low temperatures under high pressure conditions [61,62]. The presence of carbonates in inclusions in olivines is due to their ability to dissolve in an aqueous fluid, which was established experimentally. In such systems, carbon is present in the form of HCO3 and CO32− anions [63,64,65,66]. As the temperature decreases, the solubility of the fluid components decreases as well, so carbonates deposit as a solid phase. The fluid inclusions in ultrahigh-pressure silicate minerals of subducted rocks contain dissolved carbonate samples including CO2, aqueous carbonate- and bicarbonate-ion as well as CaCO3 crystals [6,67,68].

5. Conclusions

Fluid inclusions in olivine crystals were obtained in experimental modelling of olivine recrystallization in the olivine–serpentine–anthracene–metal FeNi system at high pressures and temperatures. Their composition was studied using Raman spectroscopy and the pyrolysis-free GC–MS method. It was established that the major volatile components in the composition of fluid inclusions in olivines are hydrocarbons and their derivatives: aliphatic (paraffins, olefins), cyclic (naphthenes, arenes, PAHs), oxygenated (alcohols and esters, aldehydes, ketones, carboxylic acids) and heterocyclic compounds (dioxanes, furans). Sulfonated and nitrogenated compounds, water, carbon dioxide, and hydrogen were also found in the inclusions of all analysed olivines. During the decomposition of serpentine, a large amount of water is released; therefore, the fluid at relatively low P-T parameters (2 GPa, 1100 °C) has a mainly water–hydrocarbon composition. With an increase in pressure up to 3–4.5 GPa and temperature up to 1300–1400 °C, the composition of the fluid changes greatly towards an increase in the relative amount of hydrocarbons, while the main share is occupied by light (C1-C4) aliphatic hydrocarbons. Therefore, a biogenic material with a carbon–hydrogen composition can make a certain contribution to the carbon budget in subduction processes when falling into the subduction zones and may affect the oxygen fugitivity in the subducted slab.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min13050618/s1. Table S1. Results of GC–MS analysis of volatiles extracted by mechanical shock crushing of olivines from experiment 4-36-21 in the olivine–serpentine–anthracene–FeNi system at 1100 °C and 2.0 GPa. Table S2. Results of GC–MS analysis of volatiles extracted by mechanical shock crushing of olivines from experiment 4-37-21 in the olivine–serpentine–anthracene–FeNi system at 1300 °C and 3.0 GPa. Table S3. Results of GC–MS analysis of volatiles extracted by mechanical shock crushing of olivines from experiment 4-38-21 in the olivine–serpentine–anthracene–FeNi system at 1400 °C and 4.5 GPa.

Author Contributions

Conceptualization, A.T. and V.S.; Investigation, T.B., T.T. and E.S.; Methodology, T.B., E.Z., A.C. (Aleksey Chepurov) and A.C. (Anatoly Chepurov); Project administration, A.T.; Writing—original draft, A.T.; Writing—review and editing, V.S. All authors have read and agreed to the published version of the manuscript.

Funding

The work is done on state assignment of IGM SB RAS and also was funded by the Russian Science Foundation Grant No. 21-17-00082.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Molina, J.F.; Poli, S. Carbonate stability and fluid composition in subducted oceanic crust: Experimental study on H2O-CO2-bearing basalts. Earth Planet. Sci. Lett. 2000, 176, 295–310. [Google Scholar] [CrossRef]
  2. Kerrick, D.M.; Connoly, J.A.D. Metamorphic devolatilization of subducted oceanic metabasalts: Implication for seismicity, arc magmatism and votilate recycling. Erath Planet. Sci. Lett. 2001, 189, 19–29. [Google Scholar] [CrossRef]
  3. Presnall, D.C.; Gudfinnsson, G.H. Carbonate-rich in oceanic low-velocity zone and deep mantle. Geological Society of America. Spec. Pap. 2005, 388, 207–216. [Google Scholar]
  4. Thomsen, T.B.; Schmidt, M.W. Melting of carbonated pelites at 2.5-5.0 GPa, silicate-carbonatite liquid immiscibility, and potassium-carbon metasomatism of the mantle. Erath Planet. Sci. Lett. 2008, 267, 17–31. [Google Scholar] [CrossRef]
  5. Dasgupta, R.; Hirschmann, M.M. The deep carbon cycle and melting in Earth’s interior. Earth Planet. Sci. Lett. 2010, 298, 1–13. [Google Scholar] [CrossRef]
  6. Ague, J.J.; Nicolescu, S. Carbon dioxide released from subduction zones by fluid-mediated reactions. Nat. Geosci. 2014, 7, 355–360. [Google Scholar] [CrossRef]
  7. Kelement, P.B.; Manning, C.E. Reevaluating carbon fluxes in subduction zones, what goes down, mostly comes up. Proc. Natl. Acad. Sci. USA 2015, 112, E3997–E4006. [Google Scholar] [CrossRef]
  8. Debret, B.; Sverjensky, D. Highly oxidizing fluids generated during serpentinite breakdown in subduction zones. Sci. Rep. 2017, 7, 10351. [Google Scholar] [CrossRef]
  9. Sverjensky, D.; Stagno, V.; Huang, F. Important role for organic carbon in subduction-zone fluids in the deep carbon cycle. Nat. Geosci. 2014, 7, 909–913. [Google Scholar] [CrossRef]
  10. Malvoisin, B.; Chopin, C.; Brunet, F.; Galvez, M.E. Low-temperature wollastonite formed by carbonate reduction: A marker of serpentinite redox conditions. J. Petrol. 2012, 53, 159–176. [Google Scholar] [CrossRef]
  11. Galvez, M.E.; Beyssac, O.; Martinez, I.; Benzerara, K.; Chaduteau, C.; Malvoisin, B.; Malavieille, J. Graphite formation by carbonate reduction during subduction. Nat. Geosci. 2013, 6, 473–477. [Google Scholar] [CrossRef]
  12. Buseck, P.R.; Beyssak, O. From organic matter to graphite: Graphitization. Elements 2014, 10, 421–426. [Google Scholar] [CrossRef]
  13. Vitale Brovarone, A.V.; Tumiati, S.; Piccoli, F.; Ague, J.J.; Connolli, J.A.D.; Beyssac, O. Fluid-mediated selective dissolution of subducting carbonaceous material: Implication for carbon recycling and fluid fluxes at forearc depths. Chem. Geol. 2020, 549, 119682. [Google Scholar] [CrossRef]
  14. Nakamura, Y.; Yoshino, T.; Satish-Kumar, M. Pressure dependence of graphitization: Implications for rapid recrystallization of carbonaceous material in subduction zone. Contrib. Mineral. Petrol. 2020, 175, 32. [Google Scholar] [CrossRef]
  15. Duncan, M.S.; Dasgupta, R. Rise of Earth’s atmospheric oxygen controlled by efficient subduction of organic carbon. Nat. Geosci. 2017, 10, 387–392. [Google Scholar] [CrossRef]
  16. Plank, T.; Manning, G.E. Subducting carbon. Nature 2019, 574, 343–352. [Google Scholar] [CrossRef]
  17. Chanyshev, A.D.; Litasov, K.D.; Shatskiy, A.F.; Sharygin, I.S.; Higo, Y.; Ohtani, E. Transition from melting to carbonization of naphthaline, antracene, pyrene and coronene at high pressure. Phys. Earth Planet. Inter. 2017, 270, 29–39. [Google Scholar] [CrossRef]
  18. Evans, K.A.; Reddy, S.M.; Tomkins, A.G.; Crossley, R.J.; Frost, B.R. Effects of geodynamic setting on the redox state of fluids released by subducted mantle lithosphere. Lithos 2017, 278-281, 26–42. [Google Scholar] [CrossRef]
  19. Vitale Brovarone, A.V.; Sverjensky, D.A.; Piccoli, F.; Ressico, F.; Giovannelli, D.; Daniel, I. Subduction hides high-pressure sources of energy that may feed the deep subsurface biosphere. Nat. Commun. 2020, 11, 3880–3911. [Google Scholar] [CrossRef]
  20. Tao, R.; Zhang, L.; Tian, M.; Zhu, J.; Liu, X.; Liu, J.; Höfer, H.E.; Stagno, V.; Fei, Y. Formation of abiotic hydrocarbon from reduction of carbonate in subduction zones: Constraints from petrological observation and experimental simulation. Geochim. Cosmochim. Acta 2018, 239, 390–408. [Google Scholar] [CrossRef]
  21. Vacquand, C.; Deville, E.; Beaumont, V.; Guyot, F.; Sissmann, O.; Pillot, D.; Arcilla, C.; Prinzhofer, A. Reduced gas seepages in ophiolitic complexes: Evidences for multiple origins of the H2-CH4-N2 gas mixture. Geochim. Cosmochim. Acta 2018, 223, 437–461. [Google Scholar] [CrossRef]
  22. Maffeis, A.; Ferrando, S.; Connolly, J.A.D.; Groppo, C.; Frezzotti, M.L.; Castelli, D. Thermodynamic analysis of HP-UHP fluid inclusions: The solute load and chemistry of metamorphic fluids. Geochim. Cosmochim. Acta 2021, 315, 207–229. [Google Scholar] [CrossRef]
  23. Zhang, L.; Wang, Q.; Ding, X.; Li, W.-C. Divers serpentinization and associated abiotic methanogenesis within multiple types of olivine-hosted fluid inclusions in orogenic peridotite from northern Tibet. Geochim. Cosmochim. Acta 2021, 296, 1–17. [Google Scholar] [CrossRef]
  24. McCollom, T.M.; Seewald, J.S.; German, C.R. Investigation of extractable organic compounds in deep-sea hydrothermal vent fluids along the Mid-Atlantic Ridge. Geochem. Cosmochim. Acta 2015, 156, 122–144. [Google Scholar] [CrossRef]
  25. Ménez, B.; Pisapia, C.; Andreani, M.; Jamme, F.; Vanbellingen, Q.; 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] [PubMed]
  26. Schwandner, F.M.; Seward, T.M.; Gize, A.P.; Hall, K.; Dietrich, V.J. Halocarbons and other trace heteroatomic organic compounds in volcanic gases from Vulcano (Aeolian Islands, Italy). Geochim. Cosmochim. Acta 2013, 101, 191–221. [Google Scholar] [CrossRef]
  27. Tomilenko, A.A.; Bul’bak, T.A.; Logvinova, A.M.; Sonin, V.M.; Sobolev, N.V. The composition features of volatile components in diamonds from the placers in the northeastern part of the Siberian platform by gas chromatography–mass spectrometry. Dokl. Earth Sci. 2018, 481, 953–957. [Google Scholar] [CrossRef]
  28. Sobolev, N.V.; Sobolev, A.V.; Tomilenko, A.A.; Kuz’min, D.V.; Grakhanov, S.A.; Batanova, V.G.; Logvinova, A.M.; Bul’bak, T.A.; Kostrovitskii, S.I.; Yakovlev, D.A.; et al. Prospects of searching for diamondiferous kimberlites in the northeastern Siberian Platform. Russ. Geol. Geophys. 2018, 59, 1385–1399. [Google Scholar] [CrossRef]
  29. Sobolev, N.V.; Logvinova, A.M.; Tomilenko, A.A.; Wirth, R.; Bul’bak, T.A.; Luk’yanova, L.I.; Fedorova, E.N.; Reutsky, V.N.; Efimova, E.S. Mineral and fluid inclusions in diamonds from the Urals placers, Russia: Evidence for solid molecular N2 and hydrocarbons in fluid inclusions. Geochim. Cosmochim. Acta 2019, 266, 197–219. [Google Scholar] [CrossRef]
  30. Sobolev, N.V.; Tomilenko, A.A.; Bul’bak, T.A.; Logvinova, A.M. Composition of hydrocarbons in diamonds, garnet, and olivine from diamondiferous peridotites from the Udachnaya pipe in Yakutia, Russia. Engineering 2019, 5, 471–478. [Google Scholar] [CrossRef]
  31. Nizametdinov, I.R.; Kuzmin, D.V.; Smirnov, S.Z.; Bul’bak, T.A.; Tomilenko, A.A.; Maksimovich, I.A.; Kotov, A.A. Hydrocarbons in Magmatic Fluid in Phenocrysts of Eruption Products of Men’shii Brat Volcano (Iturup Island): Data from Pyrolysis-Free Gas Chromatography–Mass Spectrometry of Melt and Fluid Inclusions. Russ. Geol. Geophys. 2022, 63, 890–900. [Google Scholar] [CrossRef]
  32. Tomilenko, A.A.; Bul’bak, T.A.; Timina, T.Y.; Shaparenko, E.O.; Simonov, V.A.; Laptev, Y.V. Composition of volatiles of sulfide deposits and carbonate structures in submarine hydrothermal fields of the Mid-Atlantic Ridge. Marine Geol. 2022, 444, 106713. [Google Scholar] [CrossRef]
  33. Zhimulev, E.I.; Sonin, V.M.; Bil’bak, T.A.; Chepurov, A.I.; Tomilenko, A.A.; Pokhilenko, N.P. Volatile Compounds of Sulfur in the Fe-C-S System at 5.3 GPa and 1300 °C. Dokl. Earth Sci. 2015, 462, 528–533. [Google Scholar] [CrossRef]
  34. Poli, S.; Schmidt, M.W. H2O transport and release in subduction zones—Experimental constraints on basaltic and andesitic systems. J. Geophys. Res. 1995, 100, 22299–22314. [Google Scholar] [CrossRef]
  35. Huang, R.; Sun, W.; Ding, X.; Zhao, Y.; Song, M. Effect of pressure on the kinetics of peridotite serpentinization. Phys. Chem. Minerals 2020, 47, 33. [Google Scholar] [CrossRef]
  36. Ulmer, P.; Trommsdorff, V. Serpentine stability to mantle depths and subduction-related magmatism. Science 1995, 268, 858–861. [Google Scholar] [CrossRef]
  37. Wunder, B.; Schreyer, W. Antigorite: High-pressure stability in the system MgO-SiO2-H2O (MSH). Lithos 1997, 41, 213–227. [Google Scholar] [CrossRef]
  38. Chepurov, A.I.; Tomilenko, A.A.; Zhimulev, E.I.; Sonin, V.M.; Chepurov, A.A.; Surkov, N.V.; Kovyazin, S.V. Problem of water in the upper mantle: Antigorite breakdown. Dokl. Earth Sci. 2010, 434, 1275–1278. [Google Scholar] [CrossRef]
  39. Chepurov, A.I.; Tomilenko, A.A.; Zhimulev, E.I.; Sonin, V.M.; Chepurov, A.A.; Kovyazin, S.V.; Timina, T.Y.; Surkov, N.V. The conservation of an aqueous fluid in inclusions in minerals and their interstices at high pressures and temperatures during the decomposition of antigorite. Russ. Geol. Geophys. 2012, 53, 234–246. [Google Scholar] [CrossRef]
  40. Stagno, V.; Frost, D.J. Carbon speciation in the asthenosphere: Experimental measurements of the redox conditions at carbonate-bearing melts coexist with graphite or diamonds in peridotite assemblages. Earth Planet. Sci. Lett. 2010, 300, 72–84. [Google Scholar] [CrossRef]
  41. Stagno, V.; Frost, D.J.; McCammon, C.A.; Mohseni, H.; Fei, Y. The oxygen fugacity at which graphite or diamond forms from carbonate-bearing melts in eclogitic rocks. Contrib. Mineral. Petrol. 2015, 169, 16. [Google Scholar] [CrossRef]
  42. Smith, E.M.; Shirey, S.B.; Nestola, F.; Bullock, E.S.; Wang, J.; Richardson, S.H.; Wang, W. Large gem diamonds from metallic liquid in Earth’s deep mantle. Science 2016, 35, 1403–1405. [Google Scholar] [CrossRef]
  43. Chepurov, A.I.; Tomilenko, A.A.; Sonin, V.M.; Zhimulev, E.I.; Bul’bak, T.A.; Chepurov, A.A.; Sobolev, N.V. Interaction of an Fe-Ni melt with anthracene (C14H10) in presence of olivine at 3 GPa: Fluid phase composition. Dokl. Earth Sci. 2020, 492, 333–337. [Google Scholar] [CrossRef]
  44. Turkin, A.I. Lead selenide as a continuous internal indicator of pressure in solid-media cells of high-pressure apparatus in the range of 4–6.8 GPa. High Temp. High Press. 2004, 36, 371–376. [Google Scholar] [CrossRef]
  45. Chepurov, A.; Zhimulev, E.; Chepurov, A.; Sonin, V. Where did the largest diamonds grow? The experiments on percolation of Fe-Ni melt through olivine matrix in the presence of hydrocarbons. Lithos 2021, 404–405, 106437. [Google Scholar] [CrossRef]
  46. Tomilenko, A.A.; Chepurov, A.I.; Sonin, V.M.; Bul’bak, T.A.; Zhimulev, E.I.; Chepurov, A.A.; Timina, T.Y.; Pokhilenko, N.P. The synthesis of methane and neavier hydrocarbons in the system graphite-iron-serpentine at 2 and 4 GPa and 1200 °C. High Temp. High Press. 2015, 44, 451–465. [Google Scholar]
  47. Sokol, A.G.; Palyanov, Y.N.; Tomilenko, A.A.; Bul’bak, T.A.; Palyanova, G.A. Carbon and nitrogen speciation in nitrogen-rich C-O–H–N fluids at 5.5–7.8 GPa. Earth Planet. Sci. Lett. 2017, 60, 234–243. [Google Scholar] [CrossRef]
  48. Bul’bak, T.A.; Tomilenko, A.A.; Gibsher, N.A.; Sazonov, A.M.; Shaparenko, E.O.; Ryabukha, M.A.; Khomenko, M.O.; Sil’yanov, S.A.; Nekrasova, N.A. Hydrocarbons in Fluid Inclusions from Native Gold, Pyrite, and Quartz of the Sovetskoe Deposit (Yenisei Ridge, Russia) According to Pyrolysis-Free Gas Chromatography-Mass Spectrometry Data. Russ. Geol. Geophys. 2020, 61, 1260–1282. [Google Scholar] [CrossRef]
  49. Sonin, V.; Tomilenko, A.; Zhimulev, E.; Bul’bak, T.; Chepurov, A.; Babich, Y.; Logvinova, A.; Timina, Y.; Chepurov, A. The composition on the fluid phase in inclusions in synthetic HPHT diamonds grown in system Fe-Ni-Ti-C. Sci. Rep. 2022, 12, 1246. [Google Scholar] [CrossRef]
  50. Swartzendruber, L.J.; Itkin, V.P.; Alcock, C.B. The Fe-Ni (Iron-Nickel) System. J. Phase Equilibria 1991, 12, 288–312. [Google Scholar] [CrossRef]
  51. Kocherzhinskii, Y.A.; Kulik, O.G.; Turkevich, V.Z. Phase equilibria in the Fe-Ni-C and Fe-Co-C systems under high temperatures and high pressures. High Temp. High Pres. 1993, 25, 113–116. [Google Scholar]
  52. Lakovino, K.; Guild, M.R.; Till, C.B. Aqueous fluids are effective oxidizing agents of the mantle in subduction zones. Contrib. Mineral. Petrol. 2020, 175, 36. [Google Scholar] [CrossRef]
  53. Sawada, Y.; Sampei, Y.; Hada, O.; Taguchi, S. Thermal degradation and polymerization of carbonaceous materials in a metapelite-granitiod magma system in the Ryoke metamorphic belt, SW Japan. J. Asian Earth Sci. 2008, 33, 91–105. [Google Scholar] [CrossRef]
  54. Manning, C.E.; Shock, E.L.; Sverjensky, D.A. The chemistry of carbon in aqueous fluids at crustal and upper-mantle conditions: Experimental and theoretical constraints. Rew. Mineral. Geochem. 2013, 75, 109–148. [Google Scholar] [CrossRef]
  55. Stephton, M.A.; Hazen, R.M. On the origins of deep hydrocarbons. Rew. Mineral. Geochem. 2013, 75, 449–465. [Google Scholar] [CrossRef]
  56. McCollom, T.M. Laboratory simulations of abiotic hydrocarbon formation in Earth’s deep subsurface. Rev. Mineral. Geochem. 2013, 75, 467–494. [Google Scholar] [CrossRef]
  57. McCollom, T.M. Abiotic methane formation during experimental serpentinization of olivine. Proc. Natl. Acad. Sci. USA 2016, 113, 13965–13970. [Google Scholar] [CrossRef]
  58. Etiope, G.; Sherwood Lollar, B. Abiotic methane on Earth. Rev. Geophys. 2013, 51, 276–299. [Google Scholar] [CrossRef]
  59. Zubkov, V.S. Tendencies in distribution and hypotheses of the genesis of condensed naphthides in magmatic rocks from various geodynamic environments. Geochem. Int. 2009, 47, 741–757. [Google Scholar] [CrossRef]
  60. Savel’ev, V.S.; Ogloblina, A.I.; Florovskaya, V.N.; Rudenko, A.P.; Kulakova, I.I. Polycondensation of carbon monoxide with hydrogen is a possible source of PAHs in nature. Dokl. Akad. Nauk SSSR 1984, 275, 733–736. [Google Scholar]
  61. Ellis, D.E.; Wyllie, P.J. Hydration and melting reactions in the system MgO-SiO2-H2O at pressure up to 100 kbar. Am. Mineral. 1979, 64, 41–48. [Google Scholar]
  62. Johnson, M.C.; Walker, D. Brucite [Mg(OH)2] dehydration and molar volume of H2O to 15 GPa. Am. Mineral. 1993, 78, 271–284. [Google Scholar]
  63. Newton, R.C.; Manning, C.E. Experimental determination of calcite solubility in H2O-NaCl solutions at deep-crust/upper mantle pressures and temperatures: Implications for metasomatic processes in shear zones. Am. Mineral. 2002, 87, 1401–1409. [Google Scholar] [CrossRef]
  64. Caciagli, N.C.; Manning, C.E. The solubility of calcite in water at 6–16 kbar and 500–800 °C. Contrib. Mineral. Petrol. 2003, 146, 355–360. [Google Scholar] [CrossRef]
  65. Facq, S.; Daniel, I.; Montagnac, G.; Cardon, H.; Sverjensky, D.A. In situ Raman study and thermodynamic model of aqueous carbonate speciation in equilibrium with aragonite under subduction zone conditions. Geochim. Cosmochim. Acta 2014, 132, 375–390. [Google Scholar] [CrossRef]
  66. Farsang, S.; Louvel, M.; Zhao, C.; Mezouar, M.; Rosa, A.D.; Widmer, R.N.; Feng, X.; Liu, J.; Redfern, S.A.T. Deep carbon cycle constrained by carbonate solubility. Nat. Commun. 2021, 12, 4311–4319. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, Z.; Shen, K.; Xiao, Y.; Hoefs, J.; Liou, J.G. Mineral and fluid inclusions in zircon of UHP metamorphic rocks from CCSD-main drill hole: A record of metamorphism and fluid activity. Lithos 2006, 92, 378–398. [Google Scholar] [CrossRef]
  68. Frezzotti, M.; Selverstone, J.; Sharp, Z.; Compagnoni, R. Carbonate dissolution during subduction revealed by diamond-bearing rocks from the Alps. Nat. Geosci. 2011, 4, 703–706. [Google Scholar] [CrossRef]
Figure 1. The assembly scheme of the HPC reaction volume.
Figure 1. The assembly scheme of the HPC reaction volume.
Minerals 13 00618 g001
Figure 2. Distribution of fluid inclusions in olivine (b) and enlarged FIs (a). FI—fluid inclusions.
Figure 2. Distribution of fluid inclusions in olivine (b) and enlarged FIs (a). FI—fluid inclusions.
Minerals 13 00618 g002
Figure 3. Elemental maps (bd) of a fluid inclusion (a) in olivine obtained experimentally in the olivine-serpentine-anthracene-metal (FeNi) system (No. 4-37-21) demonstrate the presence of solid phases containing Ca, Si, and Fe. According to Raman spectroscopy, these solid phases can be calcite (Raman lines 152, 279, 712, 1086, and 1436 cm−1), brucite (Raman lines 278, 442, 3645, and 3652 cm−1), and an unidentified solid phase (Raman lines 640 and 1114 cm−1). Photo (a) was taken in backscattered electrons.
Figure 3. Elemental maps (bd) of a fluid inclusion (a) in olivine obtained experimentally in the olivine-serpentine-anthracene-metal (FeNi) system (No. 4-37-21) demonstrate the presence of solid phases containing Ca, Si, and Fe. According to Raman spectroscopy, these solid phases can be calcite (Raman lines 152, 279, 712, 1086, and 1436 cm−1), brucite (Raman lines 278, 442, 3645, and 3652 cm−1), and an unidentified solid phase (Raman lines 640 and 1114 cm−1). Photo (a) was taken in backscattered electrons.
Minerals 13 00618 g003
Figure 4. (ad)—Raman spectra of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene-metal FeNi system (No. 4-37-21). Raman lines 2911 and 2912 cm−1 correspond to methane; Raman lines 4127, 4144, 4155, and 4162 cm−1 correspond to hydrogen; Raman line 2329 cm−1 corresponds to nitrogen; Raman lines 2873, 2934, 2970 cm−1 correspond to C-H vibrations in (CH2)n and (CH3)n groups.
Figure 4. (ad)—Raman spectra of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene-metal FeNi system (No. 4-37-21). Raman lines 2911 and 2912 cm−1 correspond to methane; Raman lines 4127, 4144, 4155, and 4162 cm−1 correspond to hydrogen; Raman line 2329 cm−1 corresponds to nitrogen; Raman lines 2873, 2934, 2970 cm−1 correspond to C-H vibrations in (CH2)n and (CH3)n groups.
Minerals 13 00618 g004
Figure 5. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman lines 223, 303, 588, 607, 824, 856, 920, and 962 cm−1 correspond to olivine; Raman lines 278 and 442 cm−1 correspond to brucite; Raman line 1086 cm−1 corresponds to calcite; Raman lines 640 and 1114 cm−1 are for an unidentified solid phase.
Figure 5. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman lines 223, 303, 588, 607, 824, 856, 920, and 962 cm−1 correspond to olivine; Raman lines 278 and 442 cm−1 correspond to brucite; Raman line 1086 cm−1 corresponds to calcite; Raman lines 640 and 1114 cm−1 are for an unidentified solid phase.
Minerals 13 00618 g005
Figure 6. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman lines 824 and 856 cm−1 correspond to olivine; Raman lines 152, 279, 712, 1086, and 1436 cm−1 correspond to calcite.
Figure 6. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman lines 824 and 856 cm−1 correspond to olivine; Raman lines 152, 279, 712, 1086, and 1436 cm−1 correspond to calcite.
Minerals 13 00618 g006
Figure 7. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman line 3582 cm−1 corresponds to OH-group in olivine; Raman lines 3645 and 3652 cm−1 correspond to OH-group in brucite. A box (a) shows an increase fragment of the Raman spectrum of olivine in the frequency range of 3560–3600 cm−1.
Figure 7. Raman spectrum of a fluid inclusion in olivine obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). Raman line 3582 cm−1 corresponds to OH-group in olivine; Raman lines 3645 and 3652 cm−1 correspond to OH-group in brucite. A box (a) shows an increase fragment of the Raman spectrum of olivine in the frequency range of 3560–3600 cm−1.
Minerals 13 00618 g007
Figure 8. Raman spectra of graphite inclusions in olivine (a) and graphite in a fluid inclusion in olivine (b), obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). a—Raman lines 1355, 1579, and 1580 cm−1 correspond to well-structured graphite; b—Raman line 1608 cm−1 corresponds to amorphized graphite.
Figure 8. Raman spectra of graphite inclusions in olivine (a) and graphite in a fluid inclusion in olivine (b), obtained in the olivine–serpentine–anthracene–metal FeNi system (No. 4-37-21). a—Raman lines 1355, 1579, and 1580 cm−1 correspond to well-structured graphite; b—Raman line 1608 cm−1 corresponds to amorphized graphite.
Minerals 13 00618 g008
Figure 9. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-36-21 (at 1100 °C and at 2.0 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Water (H2O); 3. 2-Methyl-1-propene (C4H8); 4. (Z)-2-Butene (C4H8); 5. 2-Propanone (C3H6O); 6. Butanal (C4H8O); 7. Acetic acid (C2H4O2); 8. Benzene (C6H6); 9. 2-Pentanone (C5H10O); 10. Heptane (C7H16); 11. Butanoic acid (C4H8O2); 12. Pentanoic acid (C5H10O2); 13. Phenol (C6H6O); 14. 2-Ethylhexanal (C8H16O); 15. Hexanoic acid (C6H12O2); 16. Heptanoic acid (C7H14O2); 17. Nonanal (C9H18O); 18. Cyclodecanol (C10H20O); 19. Octanoic acid (C8H16O2); 20. Decanal (C10H20O); 21. 1,3-Isobenzofurandione (C8H4O3); 22. Nonanoic acid (C9H18O2); 23. 2-Methyl-4-methoxybenzaldehyde (C9H10O2); 24. n-Decanoic acid (C10H20O2); 25. (5E)-6,10-Dimethyl-5,9-undecadien-2-one (C13H22O); 26. Pentadecane (C15H32); 27. Monomethyl phthalate (C12H24O2); 28. Dodecanoic acid (C12H24O2); 29. Hexadecane (C16H34); 30. γ-Dodecalactone (C12H22O2); 31. 2-Pentadecanone (C15H30O); 32. Heptadecane (C17H36). 33. Octadecane (C18H38).
Figure 9. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-36-21 (at 1100 °C and at 2.0 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Water (H2O); 3. 2-Methyl-1-propene (C4H8); 4. (Z)-2-Butene (C4H8); 5. 2-Propanone (C3H6O); 6. Butanal (C4H8O); 7. Acetic acid (C2H4O2); 8. Benzene (C6H6); 9. 2-Pentanone (C5H10O); 10. Heptane (C7H16); 11. Butanoic acid (C4H8O2); 12. Pentanoic acid (C5H10O2); 13. Phenol (C6H6O); 14. 2-Ethylhexanal (C8H16O); 15. Hexanoic acid (C6H12O2); 16. Heptanoic acid (C7H14O2); 17. Nonanal (C9H18O); 18. Cyclodecanol (C10H20O); 19. Octanoic acid (C8H16O2); 20. Decanal (C10H20O); 21. 1,3-Isobenzofurandione (C8H4O3); 22. Nonanoic acid (C9H18O2); 23. 2-Methyl-4-methoxybenzaldehyde (C9H10O2); 24. n-Decanoic acid (C10H20O2); 25. (5E)-6,10-Dimethyl-5,9-undecadien-2-one (C13H22O); 26. Pentadecane (C15H32); 27. Monomethyl phthalate (C12H24O2); 28. Dodecanoic acid (C12H24O2); 29. Hexadecane (C16H34); 30. γ-Dodecalactone (C12H22O2); 31. 2-Pentadecanone (C15H30O); 32. Heptadecane (C17H36). 33. Octadecane (C18H38).
Minerals 13 00618 g009
Figure 10. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-37-21 (at 1300 °C and at 3.0 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Ethane (C2H6); 3. Propane (C3H8); 4. Isobutane (C4H10); 5. Butane (C4H10); 6. Propanal (C3H6O); 7. Pentane (C5H12); 8. Butanal (C4H8O); 9. Acetic acid (C2H4O2); 10. 1-Butanol (C4H10O); 11. 2-Pentanone (C5H10O); 12. Heptane (C7H16); 13. Butanoic acid (C4H8O2); 14. Octane (C8H18); 15. Pentanoic acid (C5H10O2); 16. Nonane (C9H20); 17. Phenol (C6H6O); 18. Hexanoic acid (C6H12O2); 19. Heptanoic acid (C7H14O2); 20. Octanoic acid (C8H16O2); 21. Dodecane (C12H26); 22. Nonanoic acid (C9H18O2); 23. Tridecane (C13H28); 24. n-Decanoic acid (C10H20O2); 25. Tetradecane (C14H30); 26. Phenylpropanamide (C9H11NO); 27. Pentadecane (C15H32); 28. Hexadecane (C16H34); 29. 3-tert-Butyl-2-benzopyran-1-one (C13H14O2); 30. Pentadecanal (C15H30O); 31. γ-Tridecalactone (C13H24O2).
Figure 10. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-37-21 (at 1300 °C and at 3.0 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Ethane (C2H6); 3. Propane (C3H8); 4. Isobutane (C4H10); 5. Butane (C4H10); 6. Propanal (C3H6O); 7. Pentane (C5H12); 8. Butanal (C4H8O); 9. Acetic acid (C2H4O2); 10. 1-Butanol (C4H10O); 11. 2-Pentanone (C5H10O); 12. Heptane (C7H16); 13. Butanoic acid (C4H8O2); 14. Octane (C8H18); 15. Pentanoic acid (C5H10O2); 16. Nonane (C9H20); 17. Phenol (C6H6O); 18. Hexanoic acid (C6H12O2); 19. Heptanoic acid (C7H14O2); 20. Octanoic acid (C8H16O2); 21. Dodecane (C12H26); 22. Nonanoic acid (C9H18O2); 23. Tridecane (C13H28); 24. n-Decanoic acid (C10H20O2); 25. Tetradecane (C14H30); 26. Phenylpropanamide (C9H11NO); 27. Pentadecane (C15H32); 28. Hexadecane (C16H34); 29. 3-tert-Butyl-2-benzopyran-1-one (C13H14O2); 30. Pentadecanal (C15H30O); 31. γ-Tridecalactone (C13H24O2).
Minerals 13 00618 g010
Figure 11. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-38-21 (at 1400 °C and at 4.5 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Ethane (C2H6); 3. Propane (C3H8); 4. Isobutane (C4H10); 5. Butane (C4H10); 6. 2-Propanone (C3H6O); 7. 2-Methylbutane (C5H12); 8. Pentane (C5H12); 9. Acetic acid (C2H4O2); 10. Butanoic acid (C4H8O2); 11. Pentanoic acid (C5H10O2); 12. Phenol (C6H6O); 13. Hexanoic acid (C6H12O2); 14. Hexanoic acid propyl ester (C9H18O2); 15. Octanoic acid (C8H16O2); 16. Nonanoic acid (C9H18O2); 17. n-Decanoic acid (C10H20O2); 18. 4-Acetylbenzoic acid (C9H8O3); 19. 2,3,4,5,6-Pentamethylphenol (C11H16O); 20. Dodecanoic acid (C12H24O2).
Figure 11. Results of GC-MS analysis of volatile compounds released at mechanical crushing of fluid inclusions in olivines from experiment 4-38-21 (at 1400 °C and at 4.5 GPa). Chromatogram (a) by total ion current (TIC) and reconstructed ion chromatograms by ion current: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank. 1. Methane (CH4); 2. Ethane (C2H6); 3. Propane (C3H8); 4. Isobutane (C4H10); 5. Butane (C4H10); 6. 2-Propanone (C3H6O); 7. 2-Methylbutane (C5H12); 8. Pentane (C5H12); 9. Acetic acid (C2H4O2); 10. Butanoic acid (C4H8O2); 11. Pentanoic acid (C5H10O2); 12. Phenol (C6H6O); 13. Hexanoic acid (C6H12O2); 14. Hexanoic acid propyl ester (C9H18O2); 15. Octanoic acid (C8H16O2); 16. Nonanoic acid (C9H18O2); 17. n-Decanoic acid (C10H20O2); 18. 4-Acetylbenzoic acid (C9H8O3); 19. 2,3,4,5,6-Pentamethylphenol (C11H16O); 20. Dodecanoic acid (C12H24O2).
Minerals 13 00618 g011
Figure 12. Relative abundance of volatile compounds released during mechanical destruction of fluid inclusions in olivines obtained in the olivine–serpentine–anthracene–FeNi system: (a)—aliphatic, cyclic and oxygenated hydrocarbons; (b)—light (C1-C4), medium (C5-C12), and heavy (C13-C17) saturated hydrocarbons (paraffins); (c)—oxygenated hydrocarbons (alcohols, ethers, esters, aldehydes, ketones, carboxylic acids); (d)—sum total of hydrocarbons, carbon dioxide, water, nitrogenated and sulfonated compounds. Olivines from 4-36-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1100 °C and 2.0 GPa; olivines from 4-37-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1300 °C and 3.0 GPa; olivines from 4-38-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1400 °C and 4.5 GPa.
Figure 12. Relative abundance of volatile compounds released during mechanical destruction of fluid inclusions in olivines obtained in the olivine–serpentine–anthracene–FeNi system: (a)—aliphatic, cyclic and oxygenated hydrocarbons; (b)—light (C1-C4), medium (C5-C12), and heavy (C13-C17) saturated hydrocarbons (paraffins); (c)—oxygenated hydrocarbons (alcohols, ethers, esters, aldehydes, ketones, carboxylic acids); (d)—sum total of hydrocarbons, carbon dioxide, water, nitrogenated and sulfonated compounds. Olivines from 4-36-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1100 °C and 2.0 GPa; olivines from 4-37-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1300 °C and 3.0 GPa; olivines from 4-38-21 obtained in the olivine–serpentine–anthracene–FeNi system at 1400 °C and 4.5 GPa.
Minerals 13 00618 g012
Figure 13. Relative content of hydrocarbons, carbon dioxide, water and nitrogenated and sulfonated compounds in fluid inclusions in olivines obtained in the olivine–serpentine–anthracene-metal (FeNi) system.
Figure 13. Relative content of hydrocarbons, carbon dioxide, water and nitrogenated and sulfonated compounds in fluid inclusions in olivines obtained in the olivine–serpentine–anthracene-metal (FeNi) system.
Minerals 13 00618 g013
Table 1. Composition (relative %) of volatiles released during mechanical shock destruction of olivines from experiments in the olivine–serpentine–anthracene–FeNi system (GC-MS analysis).
Table 1. Composition (relative %) of volatiles released during mechanical shock destruction of olivines from experiments in the olivine–serpentine–anthracene–FeNi system (GC-MS analysis).
NameMW *4-36-214-37-214-38-21
Aliphatic hydrocarbons: 26.460.482.9
Paraffins (CH4-C19H40)16–26825.860.182.6
Olefins (C2H4-C15H30)28–2100.60.30.3
Cyclic hydrocarbons: 0.720.750.6
Cycloalkanes (naphthenes)) (C5H8-C12H20)68–164---
Arenes (C6H6-C16H26)78–2180.670.740.07
Polycyclic aromatic hydrocarbons (C10H8-C14H10)128–1780.050.010.53
Oxygenated hydrocarbons: 10.311.14.8
Alcohols, esters and ethers (CH4O-C14H18O4)32–2503.05.20.9
Aldehydes (C2H4O-C15H30O)44–2262.42.71.3
Ketones (C3H6O-C15H30O)58–2261.91.21.2
Carboxylic acids (C2H4O2-C14H28O2)60–2283.02.01.4
Heterocyclic compounds: 0.180.250.1
Dioxanes (C4H8O2)880.010.010.01
Furans (ethers) (C5H6O-C12H20O)82–1800.170.240.09
Nitrogenated compounds (N2-C10H21NO)28–17114.022.06.7
Sulfonated compounds (H2S-C12H26O3S)34–2501.00.50.2
CO2 440.40.50.3
H2O1847.04.54.4
The number of identified components 248213218
H/(O+H) 0.820.970.98
Alkanes/Alkenes 43.0200.0275.0
* MW—nominal mass. Note: 4-36-21 is for olivines obtained in the experiment in the olivine–serpentine–anthracene–FeNi system at 1100 °C and 2.0 GPa; 4-37-21 is for olivines obtained in the experiment in the olivine–serpentine–anthracene–FeNi system at 1300 °C and 3.0 GPa; 4-38-21 is for olivines obtained in the experiment in the olivine–serpentine–anthracene–FeNi system at 1400 °C and 4.5 GPa.
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

Tomilenko, A.; Sonin, V.; Bul’bak, T.; Zhimulev, E.; Timina, T.; Chepurov, A.; Shaparenko, E.; Chepurov, A. Impact of Solid Hydrocarbon on the Composition of Fluid Phase at the Subduction (Experimental Simulation). Minerals 2023, 13, 618. https://doi.org/10.3390/min13050618

AMA Style

Tomilenko A, Sonin V, Bul’bak T, Zhimulev E, Timina T, Chepurov A, Shaparenko E, Chepurov A. Impact of Solid Hydrocarbon on the Composition of Fluid Phase at the Subduction (Experimental Simulation). Minerals. 2023; 13(5):618. https://doi.org/10.3390/min13050618

Chicago/Turabian Style

Tomilenko, Anatoly, Valeriy Sonin, Taras Bul’bak, Egor Zhimulev, Tatiana Timina, Aleksey Chepurov, Elena Shaparenko, and Anatoly Chepurov. 2023. "Impact of Solid Hydrocarbon on the Composition of Fluid Phase at the Subduction (Experimental Simulation)" Minerals 13, no. 5: 618. https://doi.org/10.3390/min13050618

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