Next Article in Journal
Utilization of Sugar Mill Waste Ash as Pozzolanic Material in Structural Mortar
Previous Article in Journal
Evolutionary Law of Pore Structure of Ion-Adsorbed Rare Earth Ore Leaching Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluid Components in Cordierite from the Rocks of Epidote-Amphibole Facies of the Muzkol Metamorphic Complex, Tajikistan: Pyrolysis-Free GC-MS Data

by
Ksenia Igorevna Zatolokina
*,
Anatoly Alexeyevich Tomilenko
and
Taras Alexandrovich Bul’bak
V.S. Sobolev Institute of Geology and Mineralogy, Siberian Branch of the Russian Academy of Sciences, Novosibirsk 630090, Russia
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(3), 323; https://doi.org/10.3390/min13030323
Submission received: 16 December 2022 / Revised: 15 February 2023 / Accepted: 20 February 2023 / Published: 24 February 2023
(This article belongs to the Section Mineral Deposits)

Abstract

:
The composition of volatile components of cordierite from the Muzkol metamorphic complex was studied using shock destruction with pyrolysis-free gas chromatography-mass spectrometry (GC-MS) with simultaneous IR and Raman spectroscopy. Applying the GC-MS procedure, the component relative concentrations (rel.%) and composition of different zones of cordierite crystals were determined. It was found that the substantially magnesian cordierite was formed with a predominantly aqueous-carbonic acid fluid (from 57.06 to 67.88 rel.% H2O, from 24.29 to 32.95 rel.% CO2). From the center towards the crystal periphery, the molar fraction of carbon dioxide (CO2/(H2O + CO2)) decreases from 0.36 to 0.26, whereas the alkane/alkene ratio increases from 0.80 to 0.88. At least 11 homologous series of organic compounds among the identified volatile components, in addition to water and carbon dioxide, were determined, including oxygen-free aliphatic and cyclic hydrocarbons (paraffins, olefins, cyclic alkanes and alkenes, arenes, polycyclic aromatic hydrocarbons (PAHs)), as well as oxygen-containing (alcohols, esters, aldehydes, ketones, carboxylic acids) and heterocyclic (furans, dioxanes) organic compounds.

1. Introduction

One of the major problems of modern metamorphic petrology is mineral-forming fluid composition estimation. Minerals in structural cavities and channels, in which fluid components are confined, are also prospective samplers of a mineral-forming fluid, like fluid inclusions. These minerals include microporous tectoaluminosilicate cordierite, which is widespread in nature and is stable in a wide range of temperatures and pressures [1,2,3,4,5,6,7,8,9,10,11,12]. Cordierite may be found in metamorphic rocks of epidote-amphibolite, amphibolite, and granulite facies, and in hornfels [13]. It may also be found in vulcanites and granite xenoliths, and in granite [14], pyrometamorphic rocks [15,16], granite pegmatites [17,18,19,20,21,22,23,24], lunar rocks [25,26,27], and meteorites [28,29,30,31].
The frame of the cordierite structure is based on Si-Al tetrahedra, which in the (x;y) plane are combined into hexagonal rings and, in the z direction, form vertical channels (r = 1.4 Å) with large isometric cavities (r = 2.2 Å) [32]. Volatile components (Ar, He, Ne, N2, H2, CO2, H2O, CO, O2, H2S, CH4, and molecules of other hydrocarbons) are localized in it in variable amounts [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47]. The presence of hollow channels and molecules has been proved by research using IR spectroscopy, neutronography, NMR, and other methods [37,39,41,48,49,50,51,52,53,54,55,56].
The predominant fluid components in natural cordierites are H2O [45,53,57] and CO2 [43,58,59]. Water molecules are situated in channel cavities [37,39,45,51,53,55,56,58,60,61,62] and demonstrate two types of orientations of the proton-proton vector [37,45,53,56,63]. The first type (H2O–I) is represented by molecules with a parallel orientation of the H-H vector, and the second type (H2O–II) is represented by molecules oriented perpendicular to the c axis of crystals. Water molecules of the second type may be found in the cordierite structure only in the presence of Na+ [37,63,64,65]. In the ideal case, each ion in the channel is surrounded by two H2O molecules located in adjacent cavities [52,54,59,66]. High-temperature IR spectroscopy shows that at T ≈ 400°C H2O molecules of both types transit to the unbound state, acquire gas-like characteristics, and then leave the channels during dehydration [53,55].
Cordierites with increased CO2 content and a CO2 over H2O predominance were first found in granulites from the Aldan shield [67,68] and thereafter in many other regions [33,34,35,43,56,69,70,71,72,73,74,75,76,77,78,79,80,81]. According to modern concepts, CO2 molecules, similar to water molecules, enter the same cavities of structural channels [39,41,50,82,83]. Fewer experimental works have been presented on the behavior of CO2 in cordierites than on H2O behavior [38,42,84,85]. Interpretation of the previously recorded IR spectra [49] by Farmer [86] showed that 80% carbon dioxide molecules (type A) are oriented with the O–C–O vector parallel to the a axis and oscillate near the axis at a 25° oscillation angle. In this case, the c axis is rotational. The same IR spectra demonstrate the presence of CO2 molecules (type C) that are oriented along the c axis is only ≈ 20%.
Mass spectrometric and chromatographic analyses have shown that, in addition to H2O and CO2, natural cordierites contain molecules of other gases: H2, CO, N2, He, Ne, Ar, and H2S [43,64,87,88]. It is reported in the works of Zimmerman [33,34,35] that when natural cordierites were heated to 420 °C, mass spectrometric analysis indicated the presence of hydrocarbons in addition to H2O and CO2 for the first time. A similar procedure was used by Beltrame et al. [36] to detect methane. Hydrocarbons of the alkane type were identified in Alpine cordierites [40].
The possibility of incorporation of methane-butane series alkanes into the channels of Mg-cordierite without recrystallization was determined with experiments at 700 °C and pressures ranging from 200 to 1000 MPa [89]. The incorporation of methane and ammonium as well as the components of binary mixtures H2O–CH4 and H2O–NH3 into the structural channels of cordierite at P = 50, 150, 200 MPa and temperatures ranging from 400 to 800 °C was experimentally confirmed [9]. The data of quantitative mass spectrometric analysis show that the distribution of CH4, NH3, and H2O between the fluid phase and Mg-cordierite is virtually temperature-independent, but determined by the total pressure. Experiments have been performed to study the dependence of water quantity in synthetic (Mg, Fe2+)-cordierites on the composition of solid solutions and P-T-X conditions [90,91,92,93,94].
The wide presence of cordierite in nature, its ability to entrap and retain volatiles in cavities, channels, and inclusions, and detailed experimental studies allow it to be used as a sensor of thermodynamic conditions of petrogenesis and of fluid conditions of mineral-forming processes [81,95].
Recently, the composition of volatile components in cordierite from pegmatites of the Kukhilal deposit, Tajikistan, was studied using pyrolysis-free gas chromatography-mass spectrometry (GC-MS) [96]. For the first time, the authors found aliphatic hydrocarbons (paraffins and olefins), cyclic hydrocarbons (arenes, cycloalkanes, and cycloalkenes), and organic compounds, including oxygenated, nitrogenated, and sulfonated compounds, in the cordierites they studied, along with water and carbon dioxide, the content of which varies from the center to the periphery of the crystal (H2O from 89.2 to 65.7 rel.%, and CO2 from 5.5 to 23.5 rel.%). The total amount of hydrocarbons and organic compounds, containing oxygen, nitrogen, and sulfur, reaches 5.3 to 10.9 rel.% (from center to periphery).
The GC-MS method was used to determine the compositional features of fluid components preserved in the cavities, channels, and fluid inclusions of jewelry-quality cordierite from metamorphic rocks of the epidote-amphibolite facies of the Muzkol complex. Based on the distribution of fluid components, this method allows us to determine the composition of volatiles in the mineral-forming medium from different zones of the cordierite crystal, and also to estimate the trends of changes in P and T parameters at different stages of metamorphism.
In this work we studied, for the first time, the chemical composition of volatiles in cordierite from the rocks of epidote-amphibolite facies of the Muzkol metamorphic complex using pyrolysis-free GC-MS with shock destruction. Moreover, we carried out IR and Raman spectroscopy analyses.

2. Geology

The Muzkol metamorphic complex is situated in the eastern part of the Pamir Mountains in Tajikistan (Figure 1), and it is part of the Mediterranean intracontinental mobile belt, framing the Indostan Shield from the north. The complex is represented by a belt of metamorphic rocks about 130 km long and 10–30 km wide, stretching from west to east along the northern margin of the cimmerian-alpine orogenic system of Central Pamir, extending eastwards into China. The Muzkol complex is a structurally intricate antiform (Eastern Pamir anticlinorium) made of two domes resulting from fold undulation: Shatput dome in the east and Dzhalan dome in the west. The metamorphic complex has a kyanite-sillimatite zonality, formed in the Oligocene and Miocene [97]. The shape of the zonality is thermal anticline with a gradual transition from greenschist in the periphery of the dome to amphibolite facies metamorphism in its core (Figure 1). Toward the axial part of the thermal anticline, processes of alkaline (mainly sodium) metasomatosis become more intense; ultramorphism develops in the high-temperature part of the amphibolite facies, forming migmatites and gneisses. The conformal bodies of Early Paleozoic gneiss-granites are present in the cores of the domes, the location of which is most likely related to the remobilization of the foundation and to diapirism during the formation of metamorphic zonality [98]. All the formations are ruptured by the intrusions of granitoids, pegmatites, and aplite veins of the Miocene age. Jewelry-quality and collectable cordierite manifestations appear in the horizons of cordierite-biotite, cordierite-garnet-biotite, and crystalline schists and gneisses (Figure 1) [99,100].
In the works [101,102] published earlier, based on the fluid inclusions in disthene and mineralogical thermobarometry data, the temperature and pressure of Muzkol complex metamorphism were determined. According to the data, metamorphism was shown to be 650–700 °C and pressure 5.5–7.0 kbar.

3. Materials and Methods

A plate for petrographic observations of fluid inclusions and IR and Raman spectroscopic analyses was made from the sample of jewelry-quality cordierite. From the various cordierite crystal parts, the samples suitable for pyrolysis-free GC-MS analysis were selected. The selection for GC-MS analysis was made from cordierite flat-parallel plates under the optical microscope. Cordierite areas were selected without secondary minerals, without fractures, without signs of block structures and twinnings, with uniform fading in crossed nicols, etc., and were extracted. During the preparation for the analysis, these samples were not processed, and their contact with any organic substances was avoided.
The content of the main elements in cordierite was determined on the JEOL JXA-8100 electron probe microanalyzer (with 5 wave spectrometers) using the WDS procedure at the Center of Collective Use of Scientific Equipment for Multielement and Isotope Studies in the IGM SB RAS. The measurements were performed at an accelerating voltage of 20 kV, probe current of 50 to 250 nA (depending on the measured mineral and the task), signal set time of 60–120 s at the peak of the signal and 30–60 s for background measurement, and probe size of 2 μm. Microanalyzer stability and drift were monitored using internal standards similar in composition to the studied samples. These standards were determined every 30–40 measurements, and then, if required, corrections were introduced.
IR spectroscopy is an analytical method that reliably determines the forms of location, structural position, and orientation of the molecules of fluid components in an aluminosilicate frame [45,47,58,67,77]. The studies included IR scanning in unpolarized radiation oriented both parallel and perpendicular to the c axis of cordierite plates in the transmission mode. The IR transmission spectra of natural cordierite were measured without a cuvette by vacuuming the spectrometer chamber to 10−1 Torr in the range of 1000–4000 cm−1 with a resolution of 2 cm−1 on the Bruker VERTEX 70 IR Fourier spectrometer with a diffuse reflection attachment and HYPERION 2000 IR microscope.
The gas phase composition of the cordierite channels was analyzed using Raman spectroscopy [47,82,83,104,105], using the Horiba Lab Ram HR 800 spectrometer. The excitation was performed using a solid-state Nd YAG laser with a wavelength of 532 nm and power of 75 mW. The spectra were registered with the Endor semiconductor detector using Peltier assay cooling. To localize the aim in the analyzed sample, a confocal spectrometer system based on an OLYMPUS BX-41 microscope with a 100× objective and a large numerical aperture was used. The analysis was performed using backscattering geometry. The time of signal accumulation and the size of the confocal diaphragm varied depending on the analyzed size. The minimum size of the confocal aperture was 30 nm (for 5–10 μm object size), and the maximum size was 300 nm (for objects more than 100 μm). The spectra were obtained in the range of 100–4200 cm−1. The time of signal accumulation varied from 25 sec/spectral window for large objects to 400 sec/spectral window for small objects.
Gross composition of fluid gas components from the structural cavities and channels of cordierite, and also inclusions, was determined using non-pyrolysis GC-MS with shock destruction of the sample on the Thermo Scientific (Austin, TX, USA) DSQ II MS/Focus GC chromatography-mass spectrometer [96,106,107,108,109,110,111].
A new split sample of up to 0.06 cm3 volume was placed, using tweezers, into a special device (Figure S1, node 4, in Supplementary Materials) coupled online to the gas scheme of the chromatograph, and heated to T = 140–160 °C for 133 min in the current of a carrier gas-helium (99.9999% purity, initial pressure 45 kPa). Gas compound separation into components was carried out on a Restek (Bellefonte, PA, USA) Rt-Q-BOND capillary analytical column (immobile phase–100% divinylbenzene, length 30 m, inner diameter 0.32 mm, 10 μm film thickness). The gas mixture was introduced through a thermostatically controlled (270 °C) tap (Valco Valley Tool & Die, Inc., Royalton, OH, USA) into the analytical column. The constant flow rate of He was 1.7 mL·min–1, and the temperature of the GC-MS connecting line was 300 °C. The column was kept for 2 min at T = 70 °C, then heated at a rate of 25 °C·min–1 to 150 °C and further at a rate of 5 °C·min–1 to 290 °C, and kept at this temperature for 100 min.
Electron impact ionization mass spectra on the total ion current were obtained on a quadrupole mass-selective detector in Full Scan mode. Mass spectral conditions: the electron energy was 70 eV, emission current was 100 µA, iron source temperature was 200 °C, multiplier voltage was 1350 V, polarity of the recorded ions was positive, mass scan range was 5–500 amu, and scan rate was 1 scan per second. The start time of the analysis was synchronized with the sample destruction. All the gas tracts of the chromatograph, where gas moved, including the injector, the valve, and capillaries, had sulfinert coating.
The injection of the gas mixture released from the sample by means of one-act shock destruction in the He flow was conducted in real time without concentration, including cryofocus. It is worth noting that rather than subjecting the sample to pyrolysis, it was solely heated in order to convert water in the sample into a gas phase and to release atmospheric components absorbed by the sample surface and the instrument. Therefore, the unchanged gas mixture was analyzed instead of pyrolizate, which contains more oxidized compounds (H2O, CO, CO2, etc.) resulting from reactions between the gas mixture compounds, the gas mixture and accumulator surface, the gas phase compounds, and the samples. Before and after “working” analyses, blank analyses in real time were performed. The previous analysis made it possible to control the release of gases absorbed by the sample, including atmospheric components, and to record the system blank at the end of the process (the whole analytical procedure without destruction of the sample). The results of the subsequent analysis were used to determine the degree and completeness of the elution (sequence of components release) of heavy hydrocarbons and polycyclic aromatic hydrocarbons (PAHs) from the analytical column during thermostat temperature programming. If necessary, the analytical column was thermoconditioned to achieve the required blank.
The GC-MS analysis used to determine the composition of volatile components combines the advantages of two independent quantitative analytical methods for identifying individual compounds in a gas mixture. By separating the gas mixture into components, chromatography makes it possible to determine the specific retention times for each of them using the analytical column. The peak area in the chromatogram was proportional to the concentration of the relative substance in the gas mixture. Mass spectrometry provided a set of mass spectra for each compound and information on its ionic and diagnostic fragments. Each component was identified by integrating both methods. Interpretation of mass spectrometric chromatographic data with identification of peaks and isolation of some components from overlapping peaks was performed using both the Automated Mass Spectral Deconvolution and Identification System (AMDIS 2.73) software and manually with background correction against spectra from NIST 2020 and Wiley 12 libraries (NIST MS Search 2.4), standard search parameters.
The relative concentrations of volatiles in the separated mixture were determined using the normalization method: the sum of the areas of all chromatographic peaks of the analyzed mixture was equal to 100%, and its relative percentage in the analyzed mixture was determined from the size of the area of an individual component. The peak areas in the chromatogram were determined using the ICIS algorithm Xcalibur (1.4 SR1 Qual Browser). This procedure is applicable for the detection of contents of individual volatile components from a few to tens of femtograms (10−15 g).

4. Results

A jewelry-quality cordierite crystal (Figure 2) from the rocks of epidote-amphibolite facies of the Muzkol metamorphic complex was chosen as an object for the study.
Electron microprobe analysis shows minor statistically insignificant variations in the composition of cordierite (Table 1) picked from different parts of the sample. The average crystal chemical formula of cordierite is Na0.08 (Mg1.87 Fe0.25Mn0.002)2.1Al3.9Si4.99 × H2O. The magnesian number of cordierite (XMg = Mg/(Fe + Mn + Mg)) varies from 0.880 to 0.886. The formation temperature was determined from sodium content using the following formula: T (°C) = (Na [apfu] − 0.4052)/(−0.000487) [59]. The average temperature for 12 analyses was 672 °C. The evaluated metamorphism temperature based on sodium content in cordierite almost corresponds to the above-mentioned temperature [102].
A detailed examination of doubly polished thin sections of cordierite, performed with an optical microscope, showed the absence of fluid and melt inclusions in it. However, we believe that the presence of submicroscopic inclusions in these cordierites is possible. For example, inclusions in cordierite are reported in the works of Cesar et al. [112] and Santosh et al. [113].
Raman spectroscopy data indicate the presence of molecular water and carbon dioxide in the isolated cavities of cordierite (Figure 3). The obtained data on the presence of water and carbon dioxide in the matrix of cordierite agree with the works of other authors [47,114].
Raman spectroscopy analyses of cordierite (Figure 4) also demonstrate the presence of water and carbon dioxide in the isolated cavities of the cordierite studied. In [37,45,47,53,56,62,63,89], absorption bands with frequencies of 1600 (type II) and 3597 (type II), 3634 (type II), and 3690 (type I) cm−1 correspond to the valence vibrations of types I and II in isolated cavities of cordierite, and the absorption bands with frequencies of 2349 (type A) and 2355 (type C) cm−1 correspond to the vibration of carbon dioxide molecules.
Figure 5 and Figure 6, Table 2 and Table 3, and Tables S1–S6 in Supplementary Materials show the results of GC-MS analyses from different parts of the cordierite sample (see Figure 2).
Data of GC-MS analyses of volatile components from cordierite of epidote-amphibolite facies of the Muzkol metamorphic complex reveal that the predominant volatile components are water and carbon dioxide (Table 2 and Tables S1–S6). Aliphatic, cyclic, oxygen-containing, and heterocyclic hydrocarbons and their derivatives, as well as chlorinated, nitrogenated, and sulfonated compounds were also identified. The total number of components varies from 165 to 172. Here, 57.06 to 67.88 rel.% water, 24.29 to 32.95 rel.% carbon dioxide, 7.23 to 14.36 rel.% hydrocarbons, 0.44 to 0.76 rel.% nitrogenated compounds, 0.12 to 0.41 rel.% sulfonated compounds; 0.01 to 0.13 rel.% chlorinated compounds (Table 2 and Tables S1–S6).
Aliphatic, oxygen-containing, cyclic, and heterocyclic hydrocarbons are represented by 11 homologous series of organic compounds. In turn, aliphatic hydrocarbons are represented by paraffins (from 0.96 to 1.86 rel.% CH4-C17H36) and olefins (from 1.09 to 2.48 rel.% C2H2-C17H34) (Table 2 and Tables S1–S6).
Compounds from methane to heptadecane were determined in the homologous class of paraffins. Comparative analysis (Table 2) showed that the lowest relative content among the class of paraffins was found in methane in samples 1 and 3–6, and it was less than 0.01 rel.%. No methane was found in sample 2; the minimum fraction (0.01 rel.%) was detected for ethane. The largest fraction was accounted for by 5-methyltetradecane (C15H32) in samples 1, 2, 4, and 6 (0.42 rel.% in sample 1; 0.38 rel.% in sample 2; 0.45 rel.% in sample 4; 0.22 rel.% in sample 6). In sample 3, the largest content of n-pentadecane (C15H32) was 0.27 rel.%, and in sample 5, n-hexadecane had the largest fraction (C16H34), 0.23 rel.% (Table 2 and Tables S1–S6).
Olefins were found to contain compounds from ethylene (C2H4) to 1-heptadecene (C17H34) (Table 2 and Tables S1–S6). The smallest fraction in sample 1 accounted for by 1,3–pentadiene (C5H8) and (E,E)-2,4-hexadiene (C6H10) was less than 0.01 rel.%. The largest fraction among olefins in sample 1 was 0.34 rel.% and was detected in 1-decene (C10H20). In samples 2, 3, and 4, the largest content was found in 1-pentadecene (C15H30), 0.32, 0.30 and 0.33 rel.%, respectively. In sample 2, the lowest content of 1,4–pentadiene (C5H8) and (E,E)-2,4-hexadiene (C6H10) was less than 0.01 rel.%. In sample 3, the lowest fraction was detected for (E,E)-2,4-hexadiene (C6H10) (less than 0.01 rel.%). In samples 4 and 5, the minimum value was found for 1,4–pentadiene (C5H8), which was no more than 0.01 rel.%. In sample 5, the maximum content was 0.21 rel.% for (Z)-3-octene (C8H16). In sample 6, the minimum value was identified in 4-methyl-1,3-pentadiene (C6H10) and (E,E)-2,4-hexadiene (C6H10) and was less than 0.01 rel.%. The largest content in sample 6 is 0.18 rel.% for 1-decene (C10H20) (Table 2 and Tables S1–S6).
Light (C1-C4), medium (C5-C12), and heavy (C13-C17) paraffins were identified among saturated paraffins. The light ones are represented by methane (CH4), ethane (C2H6), propane (C3H8), isobutane (C4H10), and butane (C4H10). They account for 0.07 to 0.19 rel.%. Medium paraffins include pentane (C5H12), hexane (C6H14), heptane (C7H16), 3-methyleneheptane (C8H16), octane (C8H16), nonane (C9H20), decane (C10H22), undecane (C11H24), and dodecane (C12H26). The content of medium paraffins varies from 0.41 to 0.79 rel.%, and that of heavy hydrocarbons from 0.44 to 0.98 rel.%, and these include tridecane (C13H28), tetradecane (C14H30), 5-methyltetradecane (C15H32), pentadecane (C15H32), hexadecane (C16H34), 7-methylhexadecane (C17H36), and heptadecane (C17H36) (Tables S1–S6).
Cyclic hydrocarbons (Table 2 and Tables S1–S6), the content of which ranges from 0.34 to 1.86 rel.%, were found to contain naphthenes (from 0.03 to 0.17 rel.% C6H10-C8H14), arenes (from 0.29 to 1.73 rel.% C6H6-C18H28), and polycyclic aromatic hydrocarbons (PAH) (from <0.01 to 0.03 rel.% C10H8-C11H10).
In the homologous class of naphthenes, we identified such compounds as 4-methylcyclopentene (minimum fractions in sample 6 was less than 0.01 rel.%, maximum value in sample 3 was 0.05 rel.%), and 3-propylcyclopentene (minimum content in sample 5 was less than 0.01 rel.%, and the largest fraction in sample 2 was 0.11 rel.%) (Table 2 and Tables S1–S6).
The alkane/alkene ratio within the crystal varies from 0.74 to 0.89, with the lowest value in sample 4 and the highest value in samples 2 and 3.
Among the class of arenes, compounds from benzene to 1-propyl-1-nonenyl benzene were identified (Table 2 and Tables S1–S6). In sample 1, the maximum content was observed for toluene (C7H8)—0.08 rel.%, and the minimum for nonylbenzene (C15H24). In the second sample, p-xylene (C8H10) had the highest value, 0.22 rel.% (C8H10), and heptylbenzene (C13H20) the lowest value, 0.04 rel.%. In sample 3, ethylbenzene (C8H10) had the highest content, 0.97 rel.% (C8H10), and propylbenzene (C9H12) had the lowest content, 0.01 rel.%. In sample 4, the maximum content, 0.09 rel.%, was observed for toluene, and the minimum, 0.02 rel.%, for octyl benzene (C14H22). In sample 5, the highest content, 0.05 rel.%, was found for toluene (C7H8), and the lowest for ethylbenzene (C8H10), which was no more than 0.01 rel.%. In sample 6, p-xylene (C8H10) had the highest content-0.06 rel.%, and styrene (C8H8) had the lowest content, which was less than 0.01 rel.% (Table 2 and Tables S1–S6).
The homologous series of polycyclic aromatic hydrocarbons (PAH) consist of such compounds as naphthalene, 2-methylnaphthalene, and 3-methylnaphthalene. The content of hydrocarbons is less than 0.01 rel.%, ruling out naphthalene in sample 3 (0.02 rel.%).
The fraction of oxygen-containing hydrocarbons varies from 4.65 to 9.23 rel.% (Table 2 and Tables S1–S6). Alcohols, esters (CH4O-C14H18O4—from 0.53 to 1.88 rel.%), aldehydes (CH2O-C15H30O—from 1.40 to 4.61 rel.%), ketones (C3H6O-C15H30O—from 0.56 to 1.89 rel.%), and carboxylic acids (C2H4O2-C13H26O2—from 0.60 to 1.89 rel.%) were identified.
The homologous series of alcohols and esters include compounds from methanol to dipropyl phthalate. The methyl ester of 3,4-dichlorobutanoic acid (C5H8Cl2O2) had the lowest content (less than 0.01 rel.%) in samples 1, 3, 5, and 6. The lowest fraction in sample 2 was detected for dipropyl phthalate (C14H18O4)—0.02 rel.%. In sample 4, phenol (C6H6O) had the lowest content—0.01 rel.%. The highest content in samples 1, 3, and 6 was observed for dipropyl phthalate (C14H18O4)—0.17, 0.26, and 0.09 rel.%, respectively. In sample 2, the maximum concentration was found for γ-octalactone (C8H14O2)—0.28 rel%. Methanol (CH4O) had the highest content in sample 5—0.78 rel.%. In sample 4, the maximum fraction was found for 2-ethylhexanol (C8H18O)—0.44 rel.% (Table 2 and Tables S1–S6).
The homologous series of aldehydes consists of compounds from formaldehyde to n-pentadecane. The highest content in samples 1, 2, 3, 5, and 6 was observed for 5-hydroxymethylfurfural (C6H6O3)—0.38, 0.38, 1.00, 2.26 and 0.19 rel.%, respectively. The highest content in sample 4 was found for n-hexanal (C6H12O)—1.20 rel.%. The lowest value in samples 1 and 2 was revealed for 2-oxopropanal (C3H4O2)—0.01 rel.%. In samples 3 and 4, the minimum contents were detected for 3-methylbutanal (C5H10O), in sample 5, for furfural (C5H4O2), and in sample 6, for n-tridecanal (C13H26O). The content is these elements does not exceed 0.01 rel.% (Table 2 and Tables S1–S6).
Among ketone class compounds, 2-propanone to 2-pentadecanone were detected. The lowest content in all samples was found for cyclopentanone (C5H8O) and also for 2-methyl-4-heptanone in sample 3, and it was less than 0.01 rel.%. The smallest fraction in samples 1, 2, 3, and 4 was observed for 2-nonanone (C9H18O)-0.11, 011, 0.12, and 0.89 rel.%, respectively. In sample 5, the maximum value was revealed for 2-tetradecanone (C14H28O)—0.27 rel.%, and in sample 6, for 3-methyl-2,5-furandion (C5H4O3—0.06 rel.% (Table 2 and Tables S1–S6).
In the class of carboxylic acids compounds from acetic to tridecanoic acids (Table 2 and Tables S1–S6) were identified. The highest contents in samples 1, 2, 4, 5, and 6 were found for acetic acid (C2H4O2)—0.18, 0.18, 0.26, 0.22, and 0.60 rel.%, respectively. In sample 3, the maximum fraction was shown by hexanoic acid (C6H12O2)-0.34 rel.%. The smallest fraction in samples 1, 2, 3, 4, and 6 was observed for 2,4,6-trimethyl-almond acid (C11H14O3), less than 0.01 rel.%, and in sample 5, less than 0.01 rel.% was revealed for butanoic acid (C4H8O2).
The content of heterocyclic compounds varies from 0.12 to 0.45 rel.% (Table 2 and Tables S1–S6). Dioxanes (C4H8O2—from 0.01 to 0.11 rel.%) and furans (C5H6O-C15H26O—from 0.11 to 0.34 rel.%) were identified. The homologous series of dioxanes is represented by 1,4-dioxanes (maximum and minimum value in the samples does not exceed 0.01 rel.%) and 1,3-dioxane (minimum content in sample 5 is less than 0.01 rel.%, the largest fraction in sample 2 is 0.11 rel.%). The homologous class of furans includes compounds from 2-methylfuran to 2-undecylfuran. The highest content in samples 1, 3, 4, 5, and 6 was found for 2-methoxyfuran (C5H6O2)—0.7, 0.12, 0.07, 0.06, and 0.06 rel.%, respectively. In sample 2, the maximum value was revealed for 2-pentylfuran (C9H14O)—0.11 rel.%. The lowest content was less than 0.01 rel.% and was observed for 2-methyfuran (C5H6O), 3-methylfuran (C5H6O), 2-ethylfuran (C6H8O), 2-butylfuran (C8H12O), 2-heptylfuran (C11H18O), 2-octylfuran (C12H20O), 2-nonylfuran (C13H22O), and 2-undecylfuran (C15H26O).
Sulfonated compounds are represented by hydrogen sulfide (less than 0.01 rel.% H2S), carbonyl sulfide (less than 0.01 rel.% COS), sulfur dioxide (from 0.10 to 0.27 rel.% O2S), methanethiol (0.02 rel.% CH4S), carbon disulfide (less than 0.01 rel.% CS2), dimethyl disulfide (less than 0.01 rel.% C2H6S2), 2-propylthiophene (less than 0.01 rel.% C7H10S), 2-pentylthiophen (from less than 0.01 to 0.09 rel.% C9H14S), and 2-hexylthiophen (less than 0.01 rel.% C10H16S). The highest content in all samples was found for sulfur dioxide—from 0.10 to 0.27 rel.%. The fraction of other compounds was no more than 0.01 rel.%, an exception was the fraction of 2-pentylthiophen, which reached 0.09 rel.% in sample 5, and the content of methanethiol in sample 5 reached 0.02 rel.% (Table 2 and Tables S1–S6).
Nitrogenated compounds consist of molecular nitrogen (from 0.18 to 0.25 rel.% N2), 1-isocyanobutane (from less than 0.01 to 0.03 rel.% C5H9N), formamide (less than 0.01 rel.% CH3NO), acetamide (from 0.02 to 0.07 rel.% C2H5NO), pyridine (from less than 0.01 to 0.04 rel.% C5H5N), 1,2-dimethyl-piperidine (less than 0.01 rel.% C7H15N), 2-methoxy-6-methypirazine (from 0.11 to 0.25 rel.% C6H8N2O), propylneopentylamine (0.03 rel.% C8H19N), succinimide (from 0.05 to 0.14 rel.% C4H5NO2), 1-methyl-2-pyrrolidinon (less than 0.01 rel.% C5H9NO), and 3,5,5-trimethyl-N-3-methylbutylhexanamide (0.05 rel%. C14H29NO) (Table 2 and Tables S1–S6).
The halogen-containing fluid component contained in all parts of cordierite crystal is represented by methyl ester of 3,4-dichlorobutanoic acid (C5H8Cl2O2). Its content varies from less than 0.01 to 0.13 rel.%, with the highest content in sample 2.

5. Discussion

A great number of works have been published by Russian and foreign authors in which they report their studies using Raman and IR spectroscopy, ion microprobe analysis (SIMS), mass spectrometry, and chromatography on cordierite, with the position of water molecules and carbon dioxide in the cordierite structure being well studied [37,45,47,53,56,62,63,80,81,91,115,116,117].
While there is plenty of work on water and carbon dioxide, studies on other fluid components, including hydrocarbons, are less detailed. The number of works on the study of cordierites without pyrolysis are rather limited [96]. In the works using pyrolysis, mainly light hydrocarbons (methane, ethane, propane, and butane) are discussed [33,34,35,36].
For the first time, GC-MS, along with Raman and IR spectroscopy, was used to study the volatile components in cordierites, which allowed us to considerably increase the number and diversity of volatile components (to 172) in cordierite. The data on the composition of volatiles from different parts of cordierite crystal, obtained using GC-MS, are very important (Figure 5 and Figure 6).
As mentioned by Zatolokina et al. [96], a free CO2 molecule has a C–O bond length of 1.16 Å, and in cordierite structure it is reduced to 1.05 Å (Johannes and Schreyer, 1981). The molecule of N2 also undergoes compression in the cordierite structure, from N–N = 1.0976 Å in the free state to N–N = 0.905 Å in cordierite.
The molecules of normal paraffin hydrocarbons have a zigzag shape. The carbon atoms that make up the molecule lie on the same plane. The chain of CH2 groups of saturated hydrocarbons has great mobility owing to the possibility of rotation near ordinary tetrahedrally arranged bonds [118]. The C–C distance is 1.52 Å, and the angle between the C–C–C atoms is 109°28′. The hydrogen atoms (CH2 group) are arranged in pairs on the planes perpendicular to the direction of the chain elongation, with a C–H distance of 1.17 Å and an H–C–H angle of 105°. Two neighboring CH2 groups are localized on the planes arranged at a distance of 1.265 Å. The section of the chain along the plane perpendicular to the direction of elongation provides a closed oval shape with an average radius of 2.6 Å [119]. Thus, the size and geometry of chains of normal hydrocarbons allow them to penetrate the channels of the cordierite framework.
The benzene molecule is a regular hexagon. The distance between the carbon atoms in the molecule (1.39 Å) corresponds to the arithmetic mean of ordinary and double bonds. The angle of the C–C–H bonds is about 120°. The size of the benzene ring with respect to the maximum distance between the carbon atoms is ~3.8 Å.
The formation of carbonyl groups (>C=O) in the benzene ring increases this effective size by 2.42 Å (double the length of the C=O bond in the case of two substituents). When one considers that oxygen from this group was borrowed from the framework, this estimate will decrease by at least two oxygen radii (~1.1 Å). Therefore, the molecules derived from benzene can fit well the cavities between the six-membered rings, being, most likely, localized along the c axis and, hence, along the larger diameter of this cavity. Probably, other radicals (–CH2, –CH3, and –OH) can also be present in one benzene ring in addition to carbonyl radicals. The molecules of these benzene derivates, taking into account the bond lengths, are smaller than those of the rings with a carbonyl group [89].
Thus, the ratio of the sizes of cavities and channels in the cordierite structure to the critical sizes of the simplest representatives of the detected homologous series of volatiles suggests that they are localized in these cavities. Taking into account the critical diameters of more complex homologs, an assumption can be made about the deformation of the CH2 groups of linear molecules of normal hydrocarbons and other chain molecules in the cavities of cordierite. The geometric discrepancy between the parameters of the structural cavities of cordierite and the size of large hydrocarbon molecules, as well as the proven defectiveness of the material under study, indicates that the localization of particularly large gas molecules and all nonlinear molecules should be attributed to nonstructural positions.
In cordierite, as well as in many other minerals, the presence of submicroscopic inclusions is possible. The complication is the possibility of detecting them. For example, in natural diamonds, alongside fluid macro-inclusions detected by optical microscopy, we established nanoscale fluid inclusions of a similar composition [109]. However, the detection of nanoscale fluid inclusions in diamonds was possible only using the transmission electron microscopy (TEM) because the nitrogen in the inclusions was in solid form. The application of TEM on cordierite in order to detect inclusions from nanoscale to submicron scale is of no use due to the non-solid aggregate state of the inclusion contents.
The GC-MS method for studying inclusions of mineral-forming media was used only at the end of the last century [120]. Gas chromatography (GC) and mass spectrometry (MS), by supplementing each other, make it possible to avoid the main problems each of them has: low sensitivity of GC and complex overlapping of signal lines of MS [121]. At present, the non-pyrolysis GC-MS with a shock destruction of the sample in terms of sensitivity (instrumental measurement limits), dynamic ranges, rate, and diversity of determined volatiles has no alternative. GC-MS was used to study the gas phase composition of mineral-forming media in diamonds [109,122,123], minerals of mantle rocks [106,110,124], metamorphic rocks [125,126], hydrothermal minerals from ore deposits [111,127,128], and minerals of sedimentary rocks associated with oil and gas deposits [129,130].
The GC-MS data (Figure 7, Table 2 and Tables S1–S6) show that the predominant volatile components in the cordierite studied are water and carbon dioxide. Variations in the composition of fluid components at different points of sampling of cordierite, observed in our analyses, characterize a change in the composition of volatiles of mineral-forming medium at different stages of crystal formation. The similarity in diversity of volatiles in general in all GC-MS analyses and redistribution of contents of fluid components in cordierite crystal between homologous series of organic substances, water, CO2, nitrogen, etc., allowed us to reveal the trends of variation in the fluid regime of cordierite-forming process. The molar fraction of carbon dioxide (CO2/(H2O + CO2)) decreases from the center to the periphery of the crystal, from 0.36 to 0.26, with an increase in the alkane/alkene ratio, from 0.80 to 0.88.
Based on what has been above, we can conclude that volatile components in cordierite from the Muzkol metamorphic complex change nonlinearly, depending on the central part to the periphery, but from the XCO2 parameter, we observe the following trend: in the center, the ratio reaches 0.36, and in the periphery, 0.26. The nonlinearity of the CO2 system during its introduction into Mg-cordierite was experimentally proved [85], which confirms the character of the behavior of volatiles in the studied crystal.
For the first time, it was found that volatiles in natural cordierite contain medium (pentane, hexane, heptane, 3-methylene heptane, octane, nonane, decane, undecane, and dodecane) and heavy saturated hydrocarbons (tridecane, tetradecane, 5-methyltetradecane, pentadecane, hexadecane, and heptadecane) along with light (methane, ethane, propane, isobutane and butane) saturated hydrocarbons, as well as cyclic (naphthenes, arenes, and PAHs), oxygenated (alcohols and esters, aldehydes, ketones, and carboxylic acids), and heterocyclic hydrocarbons (dioxanes and furans) (Figure 5, Figure 6 and Figure 7; Tables S1–S6). It is worth noting that the sample contains a chlorinated compound: methyl ester of 3,4-dichlorobutanoic acid (C5H8Cl2O2).
Comparison of the results of studies of peripheral and central zones allows us to estimate the evolution of fluid composition during the growth of cordierite crystals from epidote-amphibolite facies of the Muzkol metamorphic complex (Muz-1) and from pegmatite of the Kukhilal deposit (Kukh-1) [96]. A common feature of both cordierites is that they were formed with the participation of predominantly water-carbon dioxide fluid. The content of volatile components varies from the central to the peripheral part of the crystals in both cases. In cordierite from Kuh-1 pegmatite, the content of H2O decreases from the center to the periphery from 89.20 to 65.60 rel.%, and the fraction of CO2 increases from 5.50 to 23.50 rel.%; i.e., XCO2 (CO2/(H2O + CO2)) increases from the center to the periphery of the crystal (from 0.06 to 0.26). In Muz-1 cordierite, a reverse situation is observed: the content of water increases from 57.06 to 67.88 rel.% from center to periphery, and that of carbon dioxide decreases from 32.95 to 24.29 rel.%; i.e., XCO2 decreases from center to periphery (from 0.36 to 0.26).
The content of aliphatic hydrocarbons in Kukh-1 cordierite increases significantly in the peripheral part of the crystal compared to its central zone (Table 3). In turn, the amount of aliphatic hydrocarbons in Muz-1 cordierite decreases in the peripheral zone compared to the central zone. The fraction of cyclic hydrocarbons Kukh-1 (naphthenes, arenes, PAHs) increases towards the periphery. In Muz-1 cordierite, the fraction of cyclic hydrocarbons decreases from the central part towards the periphery. The content of the oxygen-containing hydrocarbons identified in the sample from Kukh-1 pegmatites, the same as in aliphatic hydrocarbons, increases considerably from the center to the periphery. In cordierite from epidote-amphibolite facies Muz-1, the fraction of oxygen-containing hydrocarbons decreases from the center to the periphery.
Molecular nitrogen (N2) in natural cordierites was first established by Armbruster using gas chromatography [41]. Lepezin and Osorgin revealed [131] that the content of molecular nitrogen can reach high values (0.067 wt.%). In our work, according to GC-MS analysis, in natural cordierites, along with molecular nitrogen, 10 more nitrogenated compounds were identified for the first time: 1-isocyanobutane (C5H9N), formamide (CH3NO), acetamide (C2H5NO), pyridine (C5H5N), 1,2-dimethyl-piperidine (C7H15N), 2-methoxy-6-methylpyrazine (C6H8N2O), propylneopentylamine (C8H19N), succinimide (C4H5NO2), 1-methyl-2-pyrrolidinone (C5H9NO), and 3,5,5-trimethyl-N-3-methylbutylhexanamide (C14H29NO) (Tables S1–S6). The concentration of nitrogenated compounds Muz-1 decreases from the center toward the periphery (from 0.56 to 0.44 rel.%).
The GC-MS analyses showed that nine sulfonated compounds were identified for the first time in natural cordierite: hydrogen sulfide (H2S), carbonyl sulfide (COS), sulfur dioxide (O2S), methanethiol (CH4S), carbon disulfide (CS2), dimethyl disulfide (C2H6S2), 2-propylthiophene (C7H10S), 2-pentylthiophene (C9H14S), and 2-hexylthiophene (C10H16S) (see Tables S1–S6). The content of sulfonated compounds Muz-1 decreases from the center to the periphery (from 0.29 to 0.16 rel.%).
For the first time, a chlorinated compound, methyl ester of 3,4-dichlorobutanoic acid (C5H8Cl2O2), was detected in natural cordierite.
Thus, the predominant volatile components in cordierites from the Muzkol metamorphic complex are water and carbon dioxide, accounting for up to 90 rel.% (Figure 8). The fraction of other volatile components (aliphatic, cyclic, oxygen-containing and heterocyclic hydrocarbons, chlorinated, nitrogenated, and sulfonated compounds) varies from 7.83 to 15.30 rel.%.

6. Conclusions

The detailed microthermometric study of cordierite from the Muzkol metamorphic complex, Tajikistan, did not reveal the presence of gas-liquid inclusions in it. At the same time, IR and Raman spectroscopy indicated that cordierite channels contain molecules of two types of water and carbon dioxide. The GC-MS data show that water and carbon dioxide are predominant volatile components.
On the basis of GC-MS as well as Raman and IR spectroscopic analyses data, it was shown that cordierite from the Muzkol complex was crystallized from mainly water-carbon dioxide fluid: from 57.06 to 67.88 rel.% H2O, from 24.29 to 32.95 rel.% CO2, from 0.26 to 0.36 rel.%, XCO2. The fraction of other volatile components (aliphatic, cyclic hydrocarbons and oxygenated, heterocyclic organic compounds, and nitrogenated and sulfonated compounds varies from 7.83 to 15.30 rel.%.
In cordierite, along with water and carbon dioxide, representatives of no less than 11 homologous series of organic compounds were identified for the first time. These include oxygen-free aliphatic and cyclic hydrocarbons (paraffins, olefins, cyclic alkanes and alkenes, arenes, and PAHs), oxygen-containing and heterocyclic hydrocarbons (alcohols and esters, aldehydes, ketones, carboxylic acids, furans, and dioxanes), as well as nitrogenated and sulfonated compounds.
Using GC-MS analysis, in natural cordierites, along with molecular nitrogen (N2), 10 more nitrogenated compounds were identified for the first time, namely 1-isocyanobutane (C5H9N), formamide (CH3NO), acetamide (C2H5NO), pyridine (C5H5N), 1,2-dimethyl-piperidine (C7H15N), 2-methoxy-6-methylpyrazine (C6H8N2O), propylneopentylamine (C8H19N), succinimide (C4H5NO2), 1-methyl-2-pyrrolidinone (C5H9NO), and 3,5,5-trimethyl-N-3-methylbutylhexanamide (C14H29NO).
Using GC-MS analyses, we also identified nine sulfonated compounds in natural cordierite for the first time: hydrogen sulfide (H2S), carbonyl sulfide (COS), sulfur dioxide (O2S), methanethiol (CH4S), carbon disulfide (CS2), dimethyl disulfide (C2H6S2), 2-propylthiophene (C7H10S), 2-pentylthiophene (C9H14S), and 2-hexylthiophene (C10H16S).
The center of the cordierite crystal of the Muzkol complex with high CO2 content formed during the metamorphosis peak, and its edges most probably formed during the retrograde state.
Thus, GC-MS as well Raman and IR spectroscopy studies of cordierites from the Muzkol metamorphic complex showed that volatile components localized in structural and nonstructural positions of cordierite can characterize the features of the chemical composition of fluid phase and the mineral acts as a sampler of fluid phase. The authors suggest that fluid submicroscopic inclusions also may be presented in cordierite, according to the data obtained in this study. Further research on the topic is required.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/min13030323/s1: Table S1. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 1 (Muzkol complex, Pamir, Tajikistan); Table S2. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 2 (Muzkol complex, Pamir, Tajikistan); Table S3. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 3 (Muzkol complex, Pamir, Tajikistan); Table S4. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 4 (Muzkol complex, Pamir, Tajikistan); Table S5. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 5 (Muzkol complex, Pamir, Tajikistan); Table S6. Results of the GC-MS analysis of the gas phase extracted during shock destruction of cordierite № 6 (Muzkol complex, Pamir, Tajikistan); Figure S1. Schematic diagram of the pneumatic circuit with the main nodes of the used gas chromatograph-mass spectrometer.

Author Contributions

Conceptualization, K.I.Z., A.A.T. and T.A.B.; methodology, T.A.B. and A.A.T.; investigation, K.I.Z., T.A.B. and A.A.T.; resources, A.A.T., T.A.B. and K.I.Z.; writing—original draft preparation, K.I.Z. and T.A.B.; writing—review and editing, K.I.Z., T.A.B. and A.A.T. 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. This research received no external funding.

Data Availability Statement

Data are available in the article or Supplementary Materials.

Acknowledgments

The study was conducted due to the initiative of Lepezin Gennady Grigor’evich, with a doctoral degree at geological and mineralogical science (https://prabook.com/web/gennady_grigoryevich.lepezin/448366), (accessed on 19 February 2023), who kindly provided the samples, began and actively participated in this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schreyer, W.; Yoder, H.S. Cordierite—H2O system. In Annual Report of the Director of the Geophysical Laboratory; Carnegie Institution: Washington, DC, USA, 1958; p. 197. [Google Scholar]
  2. Schreyer, W.; Schairer, I.F. Compositions and structural states of anhydrous Mg-cordierites: A reinvestigation of the central part of the system MgO-Al2O3-SiO2. J. Petrol. 1961, 2, 324–406. [Google Scholar] [CrossRef]
  3. Schreyer, W.; Seifert, F. Compatibility relations of the aluminium silicates in the systems MgO-Al2O3-SiO2-H2O and K2O-MgO-Al2O3-SiO2-H2O at high pressures. Am. J. Sci. 1969, 267, 371–388. [Google Scholar] [CrossRef]
  4. Seifert, F.; Schreyer, W. Lower temperature stability limit of Mg cordierite in the range 1–7 kb water pressure: A redetermination. Contr. Mineral. Petrol. 1970, 27, 225–238. [Google Scholar] [CrossRef]
  5. Seifert, F. Low-temperature compatibility relations of cordierite in haplopelites of the system K2O-MgO-Al2O3-SiO2-H2O. J. Petrol. 1970, 11, 73–99. [Google Scholar] [CrossRef]
  6. Newton, R.C. An experimental determination of the high-pressure stability limits of magnesian cordierite under wet and dry conditions. J. Geol. 1972, 80, 398–420. [Google Scholar] [CrossRef]
  7. Bul’bak, T.A.; Shvedenkova, S.V. Experimental study of the equilibrium of cordierite with potassium feldspar in water-carbon-dioxide fluid. Geol. Geophys. 1998, 39, 851–855. [Google Scholar]
  8. Bul’bak, T.A.; Shvedenkov, G.Y.; Ripinen, O.I. Kinetiks of exchange reaction water-CO2 in the structural channels of (Mg, Fe2+)-cordierite. Geokhimiya 2005, 4, 429–437. [Google Scholar]
  9. Bul’bak, T.A.; Shvedenkov, G.Y. Experimental study on incorporation of C-H-O-N fluid components in Mg-cordierite. Eur. J. Mineral. 2005, 17, 829–838. [Google Scholar] [CrossRef]
  10. Likhacheva, A.Y.; Goryainov, S.V.; Krylov, A.S.; Bul’bak, T.A.; Prasad, P.S.R. Raman spectroscopy of natural cordierite at high water pressure up to 5 GPa. J. Raman Spectrosc. 2012, 43, 559–563. [Google Scholar] [CrossRef]
  11. Likhacheva, A.Y.; Goryainov, S.V.; Bul’bak, T.A. An X-ray diffraction study of the pressure-induced hydration in cordierite at 4–5 GPa. Am. Mineral. 2013, 98, 181–186. [Google Scholar] [CrossRef]
  12. Miletich, R.; Gatta, G.D.; Willi, T.; Mirwald, P.W.; Lotti, P.; Merlini, M. Cordierite under hydrostatic compression: Anomalous elastic behavior as a precursor for a pressure-induced phase transition. Am. Mineral. 2014, 99, 479–493. [Google Scholar] [CrossRef]
  13. Kalt, A.; Altherr, R.; Ludwig, T. Contact metamorphism in pelitic rocks on the island of Kos (Greece, Eastern Aegean Sea): A test for the Na-in-cordierite thermometer. J. Petrol. 1998, 39, 663–688. [Google Scholar] [CrossRef]
  14. Gerdes, A.; Montero, P.; Bea, F.; Fershater, G.; Borodina, N.; Osipova, T.; Shardakova, G. Peraluminous granites frequently with mantle-like isotope compositions: The continental-type Murzinka and Dzhabyk batholiths of the eastern Urals. Int. J. Earth Sci. 2002, 91, 3–19. [Google Scholar] [CrossRef]
  15. Sokol, E.V.; Maksimova, N.V.; Nigmatulina, E.N.; Sharygin, V.V.; Kalugin, V.M. Combustion Metamorphism; SB RAN Publisher: Novosibirsk, Russia, 2005; 284p. [Google Scholar] [CrossRef]
  16. Grapes, R.; Korzhova, S.; Sokol, E.; Seryotkin, Y. Paragenesis of unusual Fe-cordierite (sekaninaite)-containing paralava and clinker from the Kuznetsk coal basin, Siberia, Russia. Contrib. Mineral. Petrol. 2011, 162, 253–273. [Google Scholar] [CrossRef]
  17. Heinrich, E.W. Cordierite in pegmatite near Micanite, Colorado. Am. Mineral. 1950, 35, 173–184. [Google Scholar]
  18. Newton, R.C. BeO in pegmatitic cordierite. Mineral. Mag. 1966, 35, 920–927. [Google Scholar] [CrossRef]
  19. Černý, P.; Povondra, P. Beryllian cordierite from Vĕžná: (Na,K)+Be→Al. Neues Jahrb. Miner. Monatsh. 1966, 2, 36–44. [Google Scholar]
  20. Černý, P.; Povondra, P. Cordierite in West-Moravian desilicated pegmatites. Acta Univ. Carol. Geol. 1967, 3, 203–221. [Google Scholar]
  21. Černý, P.; Chapman, R.; Schreyer, W.; Ottolini, L.; Bottazzi, P.; McCammon, C.A. Lithium in sekaninaite from the type locality, Dolní Bory, Czech Republic. Can. Mineral. 1997, 35, 167–173. [Google Scholar]
  22. Jobin-Bevans, S.; Černý, P. The beryllian cordierite + beryl + spessartine assemblage, and secondary beryl in altered cordierite, Greer Lake granitic pegmatites, southeastern Manitoba. Can. Mineral. 1998, 36, 447–462. [Google Scholar]
  23. Grew, E.S.; Yates, M.G.; Barbier, J.; Shearer, C.K.; Sheraton, J.W.; Shiraishi, K.; Motoyoshi, Y. Granulite-facies beryllium pegmatites in the Napier Complex in Khmara and Amundsen Bays, western Enderby Land, East Antarctica. Polar Geosci. 2000, 13, 1–40. [Google Scholar]
  24. Sokol, E.V.; Seryotkin, Y.V.; Bul’bak, T.A. Na-Li-Be-rich cordierite from the Murzinka pegmatite field, Middle Urals, Russia. Eur. J. Mineral. 2010, 22, 565–575. [Google Scholar] [CrossRef]
  25. Dymek, R.F.; Albee, A.L.; Chodos, A.A. Petrology and origin of Boulders #2 and #3, Apollo 17 Station 2. In Proceedings of the Seventh Lunar Science Conference, Houston, TX, USA, 15–19 March 1976; pp. 2335–2378. [Google Scholar]
  26. Herzberg, C.T.; Baker, M.B. The cordierite-to-spinel-cataclasite transition: Structure of the lunar crust. Proc. Conf. Lunar Highlands Crust. Geochim. Cosmochim. Acta 1980, 1 (Suppl. 12), 113–132. [Google Scholar]
  27. Marvin, U.B.; Carey, J.W.; Lindstrom, M.M. Cordierite-Spinel Troctolite, a New Magnesium-Rich Lithology from the Lunar Highlands. Science 1989, 243, 925–928. [Google Scholar] [CrossRef]
  28. Fuchs, L.H. Occurrence of cordierite and aluminous orthoenstatite in the Allende meteorite. Am. Mineral. 1969, 54, 1645–1653. [Google Scholar]
  29. Sheng, Y.J.; Hutcheon, I.D.; Wasserburg, G.J. Origin of plagioclase-olivine inclusions in carbonaceous chondrites. Geochim. Cosmochim. Acta 1991, 55, 581–599. [Google Scholar] [CrossRef]
  30. Petaev, M.I.; Zaslavskaya, N.I.; Clarke, R.S., Jr.; Olsen, E.J.; Jarosewich, E.; Kononkova, N.N.; Holmberg, B.B.; Davis, A.M.; Ustinov, V.I.; Wood, J.A. The Chaunskij meteorite: Mineralogical, chemical and isotope data, classification and proposed origin (abstract). Meteoritics 1992, 27, 276–277. [Google Scholar]
  31. Petaev, M.I.; Clarke, R.S.; Olsen, E.J.; Jarosewich, E.; Davis, A.M.; Steele, I.M.; Lipschutz, M.E.; Wang, M.S.; Clayton, R.N.; Mayeda, T.K.; et al. Chaunskij: The most highly metamorphosed, shock-modified and metal-rich mesosiderite (abstract). Lunar Planet. Sci. 1993, 24, 1131–1132. [Google Scholar]
  32. Schreyer, W. Experimental studies on cation substitutions and fluid incorporation in cordierite. Bull. Mineral. 1985, 108, 273–291. [Google Scholar] [CrossRef]
  33. Zimmermann, J.L. Application petrogenetique de l’etude de la liberation de l’eau et du gaz carbonique des cordierites. CR Acad. Sci. Paris D 1972, 275, 519–522. [Google Scholar]
  34. Zimmermann, J.L. Etude par spectrometric de masse de la composition des fluides dans quelques cordierites du sud de la Norvege. Soc. Geol. Fr. 1973, 418. [Google Scholar]
  35. Zimmermann, J.L. La liberation de l’eau, du gaz casrbonique et des hydrocarbures des cordierites. Cinetiques des mecanismes. Determination des sides. Interet petrogenetique. Bull. Mineral. 1981, 104, 325–338. [Google Scholar]
  36. Beltrame, R.J.; Norman, D.I.; Alexander, E.C.; Savkins, F.J. Volatiles released by step-heating a cordierite to 1200 °C. Trans. Am. Geophys. Union 1976, 57, 352. [Google Scholar]
  37. Goldman, D.S.; Rossman, G.R.; Dollase, W.A. Channel constituents in cordierite. Am. Mineral. 1977, 62, 1144–1157. [Google Scholar]
  38. Johannes, B.; Schreyer, W. Experimental introduction of CO2 and H2O into Mg- cordierite. Am. J. Sci. 1981, 281, 299–317. [Google Scholar] [CrossRef]
  39. Armbruster, T.; Bloss, F.D. Orientation and effects of channel H2O and CO2 in cordierites. Am. Mineral. 1982, 67, 284–291. [Google Scholar]
  40. Mottana, A.; Fusi, A.; Potenza, B.B.; Crespi, R.; Liborio, G. Hydrocarbon-containing cordierite from Dervio-Colico road tunnel (Como, Italy). Neues Jahrb. Miner. Abh. 1983, 148, 181–199. [Google Scholar]
  41. Armbruster, T. Ar, N2, and CO2 in the Structural Cavites of Cordierite, an Optical and X-ray Single-Crystal Study. Phys. Chem. Miner. 1985, 12, 233–245. [Google Scholar] [CrossRef]
  42. Shvedenkov, G.Y.; Lepezin, G.G.; Bul’bak, T.A.; Osorgin, N.Y. Experimental study of saturation of magnesian cordierite with the components of C-O-H fluid. Geokhimiya 1995, 2, 251–262. [Google Scholar]
  43. Lepezin, G.G.; Bul’bak, T.A.; Sokol, E.V.; Shvedenkov, G.Y. Fluid components in cordierite and their importance for metamorphic petrology. Geol. Geophys. 1999, 1, 97–112. [Google Scholar]
  44. Khomenko, V.M.; Langer, K. Aliphatic hydrocarbons in structural channels of cordierite: A first evidence from polarized single-crystal IR absorption spectroscopy. Am. Mineral. 1999, 84, 1181–1185. [Google Scholar] [CrossRef]
  45. Kolesov, E.A.; Geiger, C.A. Cordierite II: The role of CO2 and H2O. Am. Mineral. 2000, 85, 1265–1274. [Google Scholar] [CrossRef]
  46. Khomenko, V.; Langer, K.; Geiger, C.A. Structural locations of the iron ions in cordierite: A spectroscopic study. Contrib. Mineral. Petrol. 2001, 141, 381–396. [Google Scholar] [CrossRef]
  47. Kolesov, B.A. Raman spectra of single H2O molecules isolated in crystal cavities. Zhurnal Strukturnoi Khimii 2006, 47, 27–40. [Google Scholar] [CrossRef]
  48. Smith, J.V.; Schreyer, W. Location of argon and water in cordierite. Miner. Mag. 1962, 33, 226–236. [Google Scholar] [CrossRef]
  49. Farrel, E.E.; Newnham, R.E. Electronic and vibrational absorption spectra. Am. Mineral. 1967, 52, 380–389. [Google Scholar]
  50. Armbruster, T.; Bloss, F.D. Channel CO2 in cordierites. Nature 1980, 286, 140–141. [Google Scholar] [CrossRef]
  51. Carson, D.G.; Rossman, G.R.; Vaughan, R.V. Orientation and motion of water molecules in cordierite: A proton nuclear magnetic resonance study. Phys. Chem. Miner. 1982, 8, 14–19. [Google Scholar] [CrossRef]
  52. Mirwald, P. Crystal chemical effects of sodium on the incorporation of H2O and CO2 in Mg-cordierite. Terra Cogn. 1983, 3, 163. [Google Scholar]
  53. Aines, R.D.; Rossman, G.R. The high temperature behavior of water and carbon dioxide in cordierite and beryl. Am. Mineral. 1984, 69, 319–327. [Google Scholar]
  54. Armbruster, T. The role of Na in the structure of low cordierite. A syngle crystal X-ray study. Am. Mineral. 1986, 71, 746–757. [Google Scholar]
  55. Giampaolo, C.; Putnis, A. The kinetics of dehydration and order-disorder of molecular H2O in Mg-cordierite. Eur. J. Miner. 1989, 1, 193–202. [Google Scholar] [CrossRef]
  56. Vry, J.K.; Brown, P.E.; Valley, J.W. Cordierite volatile content and the role of CO2 in high grade metamorphism. Am. Mineral. 1990, 75, 71–88. [Google Scholar]
  57. Bertoldi, C.; Proyer, A.; Garbe-Schönberg, D.; Behrens, H.; Dachs, E. Comprehensive chemical analyses of natural cordierites: Implications for exchange mechanisms. Lithos 2004, 78, 389–409. [Google Scholar] [CrossRef]
  58. Majumdar, A.S.; Mathew, G. Raman-Infrared (IR) Spectroscopy Study of Natural Cordierites from Kalahandi, Odisha. J. Geol. Soc. India 2015, 86, 80–92. [Google Scholar] [CrossRef]
  59. Tropper, P.; Wyhlidal, S.; Haefeker, U.A.; Mirwald, P.W. An experimental investigation of Na incorporation in cordierite in low P/high T metapelites. Miner. Petrol. 2018, 112, 199–217. [Google Scholar] [CrossRef]
  60. Sugiura, K. The water problem of cordierite. Bull. Tokyo Inst. Technol. Ser. B 1959, 1, 1–26. [Google Scholar]
  61. Lepezin, G.G.; Melenevskii, V.N. Problem of H2O and CO2 in conrdierites. Dokl. AN SSSR 1983, 269, 920–924. [Google Scholar]
  62. Winkler, B.; Goddens, G.; Hennion, B. Movement of channel H2O in cordierite observed with quasi-elastic neutron scattering. Am. Mineral. 1994, 79, 801–808. [Google Scholar]
  63. Winkler, B.; Milman, V.; Payne, M.C. Orientation, location, and total energy of hydration of channel H2O in cordierite investigated by ab-initio total energy calculations. Am. Mineral. 1994, 79, 200–204. [Google Scholar]
  64. Wood, D.L.; Nassau, K. Infrared spectra of foreign molecules in beryl. J. Chem. Phys. 1967, 47, 2200–2228. [Google Scholar] [CrossRef]
  65. Sokol, E.V.; Stolpovskaya, V.N.; Lepezin, G.G. New data on water in cordierites (based on IR spectroscopy). Dokl. RAN 1998, 359, 671–675. [Google Scholar]
  66. Knop, E.; Mirwald, P.W. Cordierite as a monitor of fluid and melt sodium activity in metapelites, migmatites and granites: Constraints from incorporation experiments. J. Conf. Abstr. 2000, 5, 58. [Google Scholar]
  67. Suknev, V.S.; Kitsul, V.I.; Lazebnik, Y.D.; Brovkin, A.A. On the presence and qualitative estimation of CO2 in cordierites based on data of IR spectroscopy and chemical analysis. Dokl. AN SSSR 1971, 200, 950–952. [Google Scholar]
  68. Kitsul, V.I.; Lazebnik, Y.D.; Brovkin, A.A.; Suknev, V.O. Diagram for determination of the iron content of cordierites. Dokl. AN SSSR 1971, 200, 1419–1422. [Google Scholar]
  69. Schreyer, W.; Gordillo, C.E.; Werding, G. A new sodian-berillian cordierite from Soto, Argentina, and the relationship between distortion index, Be content, and state of hydration. Contrib. Miner. Petrol. 1979, 70, 421–428. [Google Scholar] [CrossRef]
  70. Hofmann, J.; Kaiser, G.; Klemm, W.; Paech, H.-J. K/Na Alter von Dolerite und Metamorphiten der Shackleton Range und der Whichaway Nunataks, Ost- und Südostumrandung des Filchner-Eisschelfs (Antarktis). Z. Geol. Wiss. Berl. 1980, 8, 1227–1232. [Google Scholar]
  71. Armbruster, T.; Schreyer, W.; Hoefs, J. Very high CO2 cordierite from Norwegian Lapland; mineralogy, petrology, and carbon isotopes. Contrib. Mineral. Petrol. 1982, 81, 262–267. [Google Scholar] [CrossRef]
  72. Lai, K.K.; Ackermand, D.; Raith, P.; Seifert, F. Sapphirine-containing assemblages from Kiranur, Southern India: A study of chemographic relationships in the Na2O-FeO-MgO-Al2O3-SiO2-H2O system. Neues Jahrb. Miner. Abh. 1984, 150, 121–152. [Google Scholar]
  73. Kurepin, V.A.; Malyuk, G.A.; Kalinichenko, A.M.; Utochkin, D.V. Volatiles in cordierite from Berdichev granites (the Ukrainian Shield). Mineral. Zhurnal 1986, 8, 70–82. [Google Scholar]
  74. Perreault, S.; Martignole, J. CO2-rich cordierites in high-temperature migmatites, northeastern Grenville province, Quebec (abs). Geol. Assoc. Can. Program Abstr. 1986, 11, 114. [Google Scholar]
  75. Perreault, S.; Martignole, J. High-temperature cordierite migmatites in the north-eastern Grenville Province. J. Metamorph. Geol. 1988, 6, 673–696. [Google Scholar] [CrossRef]
  76. Geverk’yan, S.V.; Kuripin, V.A. CO2 in cordierite structure (IR spectroscopic data). Mineral. Zhurnal 1987, 9, 49–53. [Google Scholar]
  77. Le Breton, N. Infrared investigation of CO2-containing cordierites. Some implications for the study of metapelitic granulites. Contrib. Mineral. Petrol. 1989, 103, 387–396. [Google Scholar] [CrossRef]
  78. Osorgin, N.Y. The Kinetics of Degassing of (H2O and CO2) Cordierites and Their Significance for Metamorphic Petrology: Manuscript; OIGGM SB RAN: Novosibirsk, Russia, 1991; 16p. [Google Scholar]
  79. Peck, H.W.; Valley, J.W. Genesis of cordierite–getrite gneisses, central metasedimentary Belt, boundary thrust zone, Grenville province, Ontario, Canada. Can. Mineral. 2000, 38, 511–524. [Google Scholar] [CrossRef]
  80. Harley, S.L.; Carrington, D.P. The distribution of H2O between cordierite and granitic melt: H2O incorporation in cordierite and its application to high-grade metamorphism and crustal anatexis. J. Petrol. 2001, 42, 1595–1620. [Google Scholar] [CrossRef]
  81. Harley, S.L.; Thompson, P.; Hensen, B.J.; Buick, I.S. Cordierite as a sensor of fluid conditions in high-grade metamorphism and crustal anatexis. J. Metamor. Geol. 2002, 20, 71–86. [Google Scholar] [CrossRef]
  82. Kaindl, R.; Tropper, P.; Deibla, I. A semi-quantitative technique for determination of CO2 in cordierite by Raman spectroscopy in thin sections. Eur. J. Mineral. 2006, 18, 331–335. [Google Scholar] [CrossRef]
  83. Kaindl, R.; Többens, D.; Haefeker, U. Quantum-mechanical calculations of the Raman spectra of Mg- and Fe-cordierite. Am. Mineral. 2011, 96, 1568–1574. [Google Scholar] [CrossRef]
  84. Aranovich, L.Y.; Podlessky, K.K.; Shchenochkina, N.I. Experimental determination of CO2 solubility in cordierite. Dokl. AN SSSR 1981, 261, 728–730. [Google Scholar]
  85. Le Breton, N.; Schreyer, W. Experimental CO2 incorporation in to Mg-cordierite: Nonlinear behaviour of the system. Eur. J. Mineral. 1993, 7, 427–438. [Google Scholar] [CrossRef]
  86. Farmer, V.C. Infrared Spectra of Minerals; The Mineralogical Society: London, UK, 1974; 539p. [Google Scholar]
  87. Damon, P.E.; Kulp, J.L. Excess helium and argon in beryl and other minerals. Am. Mineral. 1958, 43, 433–459. [Google Scholar]
  88. Saito, K.; Alexander, E.C., Jr.; Dragon, J.C.; Zashu, S. Rare gases in cyclosilicates and cogenetic minerals. J. Geophys. Res. 1984, 89, 7891–7901. [Google Scholar] [CrossRef]
  89. Bul’bak, T.A.; Shvedenkov, G.Y.; Lepezin, G.G. On saturation of magnesian cordierite with alkanes at high temperatures and pressures. Phys. Chem. Mineral. 2002, 29, 140–154. [Google Scholar] [CrossRef]
  90. Boberski, C.; Schreyer, W. Synthesis and water contents of Fe2+—Containing cordierites. Eur. J. Mineral. 1990, 2, 565–584. [Google Scholar] [CrossRef]
  91. Bul’bak, T.A.; Shvedenkov, G.Y.; Lepezin, G.G. Replacement of H2O molecules by D2O and CO2 in the structural channels of cordierite. Geokhimiya 1999, 1, 75–81. [Google Scholar]
  92. Shvedenkov, G.Y.; Reverdatto, V.V.; Bul’bak, T.A.; Bryksina, N.A. Bimetasomatic diffusion zonality in the CaO–MgO–SiO2–H2O–CO2 system: Experiments using natural samples. Petrology 2006, 14, 515–527. [Google Scholar] [CrossRef]
  93. Bul’bak, T.A.; Shvedenkova, S.V. Dependence of the water content in channels of the cordierite structure on the composition of its Fe-Mg solid solution. Dokl. Akad. Nauk. 2008, 419, 661–664. [Google Scholar] [CrossRef]
  94. Bul’bak, T.A.; Shvedenkova, S.V. Solid solutions of (Mg, Fe2+)-cordierite: Synthesis, water content, magnetic properties. Geokhimiya 2011, 49, 411–426. [Google Scholar] [CrossRef]
  95. Kurepin, V.A. Cordierite as an indicator of thermodynamic conditions of petrogenesis. Contrib. Mineral. Petrol. 2010, 160, 391–406. [Google Scholar] [CrossRef]
  96. Zatolokina, K.I.; Tomilenko, A.A.; Bul’bak, T.A.; Lepezin, G.G. Volatile components in cordierite and coexisting tourmaline and quartz from pegmatites of the Kuhilal deposit (Pamirs, Tajikistan). Geol. Geophys. 2021, 62, 1411–1431. [Google Scholar] [CrossRef]
  97. Dufour, M.S.; Popova, V.A.; Krivets, T.N. Alpine Metamorphic Complex of the Eastern Part of Central Pamirs; Leningrad University: Saint Petersburg, Russia, 1970; 126p. [Google Scholar]
  98. Gorokhov, I.M.; Dufour, M.S.; Neimark, D.A.; Amelin, Y.V.; Ovchinnikova, G.V.; Gorokhovsky, B.M. Early Paleozoic fragments of Gondwana in the Nappes of the Central Pamirs and Lower Himalayas: Geochemical and Isotopic Characteristics. Stratigr. Geol. Korrelyatsiya 1993, 8, 20–34. [Google Scholar]
  99. Dufour, M.S.; Popova, V.A. Processes of regional metamorphism in the Muzkol metamorphic complex in the Central Pamirs. Vestn. Leningr. Univ. 1975, 12, 14–20. [Google Scholar]
  100. Gilev, A.V. Jewelry-quality cordierite of Central Pamir. Mineral. Tadjikistana 1989, 8, 33–47. [Google Scholar]
  101. Dufour, M.S.; Kotov, N.V. Thermodynamic conditions for the manifestation of metamorphism and metasomatism in the rocks of the eastern part of the Central Pamirs. Izv. USSR Acad. Sci. Ser. Geol. 1972, 10, 24–36. [Google Scholar]
  102. Dufour, M.S.; Sedova, I.S. On the evolution of P-T parameters during the formation of metamorphic zoning of the Muzkol complex (Central Pamir) according to the study of inclusions in mineral-forming media. Izv. USSR Acad. Sci. Ser. Geol. 1975, 10, 49–58. [Google Scholar]
  103. Dufour, M.S.; Kol’tsov, A.B.; Zolotarev, A.A.; Kuznetsov, A.B. Corundum-bearing metasomatic rocks in the Central Pamirs. Petrologiya 2007, 15, 160–177. [Google Scholar] [CrossRef]
  104. Dubessy, J.; Poty, B.; Ramboz, C. Advances in C–O–H–N–S fluid geochemistry based on micro-Raman spectrometric analysis of fluid inclusions. Eur. J. Mineral. 1989, 1, 517–534. [Google Scholar] [CrossRef]
  105. Frezzotti, M.L.; Tecce, F.; Casagli, A. Raman spectroscopy for fluid inclusion analysis. J. Geochem. Explor. 2012, 112, 1–20. [Google Scholar] [CrossRef]
  106. Tomilenko, A.A.; Kuz’min, D.V.; Bul’bak, T.A.; Timina, T.Y.; Sobolev, N.V. Composition of primary fluids and melt inclusions in regenerated olivines from hypabbysal kimberlites of theMalokuonapskaya pipe (Yakutia). Dokl. Akad. Nauk. 2015, 465, 213–217. [Google Scholar] [CrossRef]
  107. Sokol, A.G.; Tomilenko, A.A.; Bul’bak, T.A.; Sokol, I.A.; Palyanov, Y.N. Carbon and nitrogen speciation in Npoor C-O-H fluids at 6.3 GPa and 1100–1400 degrees. C. Sci. Rep. 2017, 7, 1–19. [Google Scholar] [CrossRef]
  108. 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, 460, 234–243. [Google Scholar] [CrossRef]
  109. 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]
  110. 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]
  111. 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, 1535–1560. [Google Scholar] [CrossRef]
  112. Cesare, B.; Maineri, C.; Toaldo, A.B.; Pedron, D.; Vigil, A.A. Immiscibility between carbonic fluids and granitic melts during crustal anatexis: A fluid and melt inclusion study in the enclaves of the Neogene Volcanic Province of SE Spain. Chem. Geol. 2007, 237, 433–449. [Google Scholar] [CrossRef]
  113. Santosh, M.; Jackson, D.H.; Harris, N.B.W. The Significance of Channel and Fluid-Inclusion CO2 in Cordierite: Evidence from Carbon Isotopes. J. Petrol. 1993, 34, 233–258. [Google Scholar] [CrossRef]
  114. Herms, P.; Schenk, V. Fluid inclusions in granulite-facies metapelites of the Hercynian ancient lower crust of the Serre, Calabria, Southern Italy. Contrib. Mineral. Petrol. 1992, 112, 393–404. [Google Scholar] [CrossRef]
  115. Lepezin, G.G.; Kuznetsova, I.K.; Lavrent’ev, Y.G.; Chmel’nikova, O.S. Optical methods of determination of the water contents in cordierites. Contrib. Mineral. Petrol. 1976, 58, 319–329. [Google Scholar] [CrossRef]
  116. Lepezin, G.G.; Melenevsky, V.N. On the problem of water diffusion in the cordierites. Lithos 1977, 10, 49–57. [Google Scholar] [CrossRef]
  117. Mirwald, P.W.; Tropper, P. Structural disequilibrium in cordierites from long-duration experiments using crystal chemical constraints. Mitt. Osterr. Mineral. Gesellschaft. 2015, 161, 88. [Google Scholar]
  118. Zhdanova, N.V.; Khalif, A.L. Drying of Hydrocarbon Gases; Khimiya Publishing House: Moscow, Russia, 1984; 200p. (In Russian) [Google Scholar]
  119. Penkala, T. Zapys Krystalochemii; Panstwowe Wydawnictwo Naukowe: Warszawa, Poland, 1972; pp. 401–402. [Google Scholar]
  120. Andrawes, F.; Holzer, G.; Roedder, E.; Gibson, E.K.; Oro, J. Gas chromatographic analysis of volatiles in fluid and gas inclusions. J. Chromatogr. 1984, 19, 181–193. [Google Scholar] [CrossRef] [PubMed]
  121. Salvi, S.; William-Jones, A.E. Bulk analysis of volatiles in fluid inclusions. In Fluid Inclusions: Analysis and Interpretation; Samson, I., Anderson, A., Marshall, D., Eds.; Mineralogical Association of Canada: Quebec, QC, Canada, 2003; Volume 32, Chapter 10; pp. 10-1–10-30. [Google Scholar]
  122. 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. Akad. Nauk. 2018, 481, 310–314. [Google Scholar] [CrossRef]
  123. Sokol, A.G.; Tomilenko, A.A.; Bul’bak, T.A.; Sokol, I.A.; Zaikin, P.A.; Sobolev, N.V. Composition of reduced mantle fluids: Evidence from modelling experiments and fluid inclusions in natural diamond. Russ. Geol. Geophys. 2020, 61, 810–825. [Google Scholar] [CrossRef]
  124. Tomilenko, A.A.; Bul’bak, T.A.; Khomenko, M.O.; Kuz’min, D.V.; Sobolev, N.V. The composition of volatile components in olivines from Yakutian kimberlites of various ages: Evidence from gas chromatography–mass spectrometry. Dokl. Akad. Nauk. 2016, 468, 684–689. [Google Scholar] [CrossRef]
  125. Schreiber, U.; Mayer, C.; Schmitz, O.J.; Rosendahl, P.; Bronja, A.; Greule, M.; Keppler, F.; Mulder, I.; Sattler, T.; Schöler, H.F. Organic compounds in fluid inclusions of Archean quartz—Analogues of prebiotic chemistry on early Earth. PLoS ONE 2017, 12, e0177570. [Google Scholar] [CrossRef]
  126. Grozeva, N.G.; Klein, F.; Seewald, J.S.; Sylva, S.P. Chemical and isotopic analyses of hydrocarbon-containing fluid inclusions in olivine-rich rocks. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2020, 378, 20180431. [Google Scholar] [CrossRef]
  127. Ryabukha, M.A.; Gibsher, N.A.; Tomilenko, A.A.; Bul’bak, T.A.; Khomenko, M.O.; Sazonov, A.M. P-T-X parameters of metamorphogene and hydrothermal fluids, isotopy and age of the Bogunai gold deposit, southern Yenisei Ridge (Russia). Geol. Geophys. 2015, 56, 1153–1172. [Google Scholar] [CrossRef]
  128. Shaparenko, E.; Gibsher, N.; Tomilenko, A.; Sazonov, A.; Bul’bak, T.; Ryabukha, M.; Khomenko, M.; Silyanov, S.; Nekrasova, N.; Petrova, M. Ore–Bearing Fluids of the Blagodatnoye Gold Deposit (Yenisei Ridge, Russia): Results of Fluid Inclusion and Isotopic Analyses. Minerals 2021, 11, 1090. [Google Scholar] [CrossRef]
  129. Karlsen, D.A.; Nedkvitne, T.; Larter, S.R.; Bjørlykke, K. Hydrocarbon composition of authigenic inclusions: Application to elucidation of petroleum reservoir filling history. Geochim. Cosmochim. Acta 1993, 57, 3641–3659. [Google Scholar] [CrossRef]
  130. George, S.C.; Volk, H.; Dutkiewicz, A. Mass Spectrometry Techniques for Analysis of Oil and Gas Trapped in Fluid Inclusions. In Mass Spectrometry Handbook; Lee, M.S., Ed.; Wiley: Hoboken, NJ, USA, 2012; pp. 645–673. [Google Scholar] [CrossRef]
  131. Lepezin, G.G.; Osorgin, N.Y. Estimation of fluid composition of metamorphic complexes of moderate pressure. Dokl. RAN 1992, 324, 648–653. [Google Scholar]
Figure 1. Schematic map of the metamorphic zoning of the Muzkol metamorphic complex. 1—boundaries of metamorphic facies; 2—greenschist facies (zones 3 and 4); 3—epidote-amphibolite facies (zone 2); 4—amphibolite facies (zone 1); 5—Neogene granites and syenites; 6—Early Paleozoic gneiss-granites; 7—boundaries of tectonic nappes; 8—sampling site (Jalan-3) [103], with additions by the authors.
Figure 1. Schematic map of the metamorphic zoning of the Muzkol metamorphic complex. 1—boundaries of metamorphic facies; 2—greenschist facies (zones 3 and 4); 3—epidote-amphibolite facies (zone 2); 4—amphibolite facies (zone 1); 5—Neogene granites and syenites; 6—Early Paleozoic gneiss-granites; 7—boundaries of tectonic nappes; 8—sampling site (Jalan-3) [103], with additions by the authors.
Minerals 13 00323 g001
Figure 2. Photographs of the original cordierite crystal from the Muzkol metamorphic complex, Tajikistan (a), and its cut (b) with the indicated sampling sites (No. 1–6) for GC-MS analysis.
Figure 2. Photographs of the original cordierite crystal from the Muzkol metamorphic complex, Tajikistan (a), and its cut (b) with the indicated sampling sites (No. 1–6) for GC-MS analysis.
Minerals 13 00323 g002
Figure 3. Raman spectra of (a) cordierite (host mineral), (b) single water molecules (Raman lines 3581 and 3600 cm−1), and (c) carbon dioxide (Raman line 1383 cm−1) isolated in crystal cavities, from the Muzkol metamorphic complex, Tajikistan.
Figure 3. Raman spectra of (a) cordierite (host mineral), (b) single water molecules (Raman lines 3581 and 3600 cm−1), and (c) carbon dioxide (Raman line 1383 cm−1) isolated in crystal cavities, from the Muzkol metamorphic complex, Tajikistan.
Minerals 13 00323 g003
Figure 4. IR spectrum of a cordierite crystal from the Muzkol metamorphic complex, Tajikistan. Absorption bands with frequencies of 1600 (type II) and 3597 (type II), 3634 (type II), and 3690 (type I) cm−1 correspond to stretching vibrations of water molecules of types I and II in cordierite cavities; absorption bands with frequencies of 2349 (A-type) and 2355 (C-type) cm−1 correspond to vibrations of carbon dioxide molecules.
Figure 4. IR spectrum of a cordierite crystal from the Muzkol metamorphic complex, Tajikistan. Absorption bands with frequencies of 1600 (type II) and 3597 (type II), 3634 (type II), and 3690 (type I) cm−1 correspond to stretching vibrations of water molecules of types I and II in cordierite cavities; absorption bands with frequencies of 2349 (A-type) and 2355 (C-type) cm−1 correspond to vibrations of carbon dioxide molecules.
Minerals 13 00323 g004
Figure 5. Results of GC-MS analysis of volatile components extracted from the central cordierite zone of the Muzkol metamorphic complex, Tajikistan (sample 1). Chromatogram a is a total ion current (TIC) chromatogram and reconstructed ion chromatograms for ion current are presented in: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank (TIC). 1—Carbon dioxide (CO2); 2—water (H2O); 3—n-propane (C3H8); 4—acetaldehyde (C2H4O); 5—2-methyl-1-propene (C4H8); 6—isobutane (C4H10); 7—2-propanone (C3H6O); 8—n-pentane (C5H12); 9—2-methylfuran (C5H6O); 10—acetic acid (C2H4O2); 11—1-butanol (C4H10O); 12—1,4-dioxane (C2H8O2); 13—(E, E)-2,4-hexadiene (C6H10); 14—n-heptane (C7H16); 15—hexanal (C6H12O); 16—n-butanoic acid (C4H8O2); 17—octane (C8H18); 183-methylbutanoic acid (C4H8O2); 19–n-pentanoic acid (C5H10O2); 20—n-nonane (C9H20); 21—propylbenzene (C9H12); 22—n-hexanoic acid (C6H12O2); 23—1-decene (C10H20); 24—n-heptanoic acid (C7H14O2); 25—n-nonanal (C9H18O); 26—n-undecane (C11H24); 27—2-phenoxyethanol (C8H10O2); 28—n-octanoic acid (C8H16O2); 29—n-decanal (C10H20O); 30—1,3-isobenzofurandione (C8H4O3); 31—n-nonanoic acid (C9H18O2); 32—n-undecanal (C11H22O); 33—heptylbenzene (C13H20); 34—n-decanoic acid (C9H20O2); 35—5-methyltetradecane (C15H32); 36—1-pentadecene (C15H30); 37—pentadecane (C15H32); 38—2,4,6-trimethylmandelic acid (C11H14O3); 39—n-dodecanoic acid (C12H24O2); 40—n-tetradecanal (C14H28O); 41—n-hexadecane (C16H34); 42—2-pentadecanone (C15H30O); 43-dipropyl phthalate (C14H18O4).
Figure 5. Results of GC-MS analysis of volatile components extracted from the central cordierite zone of the Muzkol metamorphic complex, Tajikistan (sample 1). Chromatogram a is a total ion current (TIC) chromatogram and reconstructed ion chromatograms for ion current are presented in: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank (TIC). 1—Carbon dioxide (CO2); 2—water (H2O); 3—n-propane (C3H8); 4—acetaldehyde (C2H4O); 5—2-methyl-1-propene (C4H8); 6—isobutane (C4H10); 7—2-propanone (C3H6O); 8—n-pentane (C5H12); 9—2-methylfuran (C5H6O); 10—acetic acid (C2H4O2); 11—1-butanol (C4H10O); 12—1,4-dioxane (C2H8O2); 13—(E, E)-2,4-hexadiene (C6H10); 14—n-heptane (C7H16); 15—hexanal (C6H12O); 16—n-butanoic acid (C4H8O2); 17—octane (C8H18); 183-methylbutanoic acid (C4H8O2); 19–n-pentanoic acid (C5H10O2); 20—n-nonane (C9H20); 21—propylbenzene (C9H12); 22—n-hexanoic acid (C6H12O2); 23—1-decene (C10H20); 24—n-heptanoic acid (C7H14O2); 25—n-nonanal (C9H18O); 26—n-undecane (C11H24); 27—2-phenoxyethanol (C8H10O2); 28—n-octanoic acid (C8H16O2); 29—n-decanal (C10H20O); 30—1,3-isobenzofurandione (C8H4O3); 31—n-nonanoic acid (C9H18O2); 32—n-undecanal (C11H22O); 33—heptylbenzene (C13H20); 34—n-decanoic acid (C9H20O2); 35—5-methyltetradecane (C15H32); 36—1-pentadecene (C15H30); 37—pentadecane (C15H32); 38—2,4,6-trimethylmandelic acid (C11H14O3); 39—n-dodecanoic acid (C12H24O2); 40—n-tetradecanal (C14H28O); 41—n-hexadecane (C16H34); 42—2-pentadecanone (C15H30O); 43-dipropyl phthalate (C14H18O4).
Minerals 13 00323 g005
Figure 6. Results of GC-MS analysis of volatile components extracted from the cordierite marginal zone of the Muzkol metamorphic complex, Tajikistan (sample 4). Chromatogram a is a total ion current (TIC) chromatogram and reconstructed ion chromatograms for ion current are presented in: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank (TIC). 1—Carbon dioxide (CO2); 2—water (H2O); 3—n-propane (C3H8); 4—2-methyl-1-propene (C4H8); 5—acetaldehyde (C2H4O); 6—2-propanone (C3H6O); 7—acetic acid (C2H4O2); 8—1-butanol (C4H10O); 9—1,4-dioxane (C2H8O2); 10—(E, E)-2,4-hexadiene (C6H10); 11—n-heptane (C7H16); 12—n-hexanal (C6H12O); 13—n-butanoic acid (C4H8O2); 14—n-octane (C8H18); 15—3-methylbutanoic acid (C4H8O2); 16—n-pentanoic acid (C5H10O2); 17—n-nonane (C9H20); 18—propylbenzene (C9H12); 19—n-hexanoic acid (C6H12O2); 20—1-decene (C10H20); 21 n-heptanoic acid (C7H14O2); 22—n-nonanal (C9H18O); 23—n-undecane (C11H24); 24—2-phenoxyethanol (C8H10O2); 25—n-octanoic acid (C8H16O2); 26—n-decanal (C10H20O); 27—1,3-isobenzofurandione (C8H4O3); 28—n-nonanoic acid (C9H18O2); 29—n-undecanal (C11H22O); 30—n-decanoic acid (C9H20O2); 31—5-methyltetradecane (C15H32); 32—n-undecanoic acid (C11H22O2); 33—1-pentadecene (C15H30); 34—n-pentadecane (C15H32); 35—2,4,6-trimethylmandelic acid (C11H14O3); 36—n-dodecanoic acid (C12H24O2); 37—n-tetradecanal (C14H28O); 38—n-hexadecane (C16H34); 39—2-pentadecanone (C15H30O); 40—dipropyl phthalate (C14H18O4).
Figure 6. Results of GC-MS analysis of volatile components extracted from the cordierite marginal zone of the Muzkol metamorphic complex, Tajikistan (sample 4). Chromatogram a is a total ion current (TIC) chromatogram and reconstructed ion chromatograms for ion current are presented in: b—m/z (43 + 57 + 71 + 85); c—m/z 60; d—m/z 149; e—blank (TIC). 1—Carbon dioxide (CO2); 2—water (H2O); 3—n-propane (C3H8); 4—2-methyl-1-propene (C4H8); 5—acetaldehyde (C2H4O); 6—2-propanone (C3H6O); 7—acetic acid (C2H4O2); 8—1-butanol (C4H10O); 9—1,4-dioxane (C2H8O2); 10—(E, E)-2,4-hexadiene (C6H10); 11—n-heptane (C7H16); 12—n-hexanal (C6H12O); 13—n-butanoic acid (C4H8O2); 14—n-octane (C8H18); 15—3-methylbutanoic acid (C4H8O2); 16—n-pentanoic acid (C5H10O2); 17—n-nonane (C9H20); 18—propylbenzene (C9H12); 19—n-hexanoic acid (C6H12O2); 20—1-decene (C10H20); 21 n-heptanoic acid (C7H14O2); 22—n-nonanal (C9H18O); 23—n-undecane (C11H24); 24—2-phenoxyethanol (C8H10O2); 25—n-octanoic acid (C8H16O2); 26—n-decanal (C10H20O); 27—1,3-isobenzofurandione (C8H4O3); 28—n-nonanoic acid (C9H18O2); 29—n-undecanal (C11H22O); 30—n-decanoic acid (C9H20O2); 31—5-methyltetradecane (C15H32); 32—n-undecanoic acid (C11H22O2); 33—1-pentadecene (C15H30); 34—n-pentadecane (C15H32); 35—2,4,6-trimethylmandelic acid (C11H14O3); 36—n-dodecanoic acid (C12H24O2); 37—n-tetradecanal (C14H28O); 38—n-hexadecane (C16H34); 39—2-pentadecanone (C15H30O); 40—dipropyl phthalate (C14H18O4).
Minerals 13 00323 g006
Figure 7. Content of hydrocarbons, carbon dioxide, and water in cordierites from the Muzkol metamorphic complex of Tajikistan based on 6 samples: (a) relative content of aliphatic, cyclic, and oxygenated hydrocarbons; (b) relative content of “light” (C1–C4), “medium” (C5–C12) and “heavy” (C13–C17) hydrocarbons; (c) relative content of alcohols, ethers and esters, aldehydes, ketones, and carboxylic acids; (d) relative content of hydrocarbons, carbon dioxide, water, and (e) nitrogen- and sulfur-containing compounds; 1–6—sample numbers.
Figure 7. Content of hydrocarbons, carbon dioxide, and water in cordierites from the Muzkol metamorphic complex of Tajikistan based on 6 samples: (a) relative content of aliphatic, cyclic, and oxygenated hydrocarbons; (b) relative content of “light” (C1–C4), “medium” (C5–C12) and “heavy” (C13–C17) hydrocarbons; (c) relative content of alcohols, ethers and esters, aldehydes, ketones, and carboxylic acids; (d) relative content of hydrocarbons, carbon dioxide, water, and (e) nitrogen- and sulfur-containing compounds; 1–6—sample numbers.
Minerals 13 00323 g007
Figure 8. Contents (rel.%) of volatile components (a) of the central part (sample 1), (b) intermediate part (sample 3), and (c) marginal part (sample 6) of a cordierite crystal from the Muzkol metamorphic complex, Tajikistan. HCs are hydrocarbons.
Figure 8. Contents (rel.%) of volatile components (a) of the central part (sample 1), (b) intermediate part (sample 3), and (c) marginal part (sample 6) of a cordierite crystal from the Muzkol metamorphic complex, Tajikistan. HCs are hydrocarbons.
Minerals 13 00323 g008
Table 1. Chemical analysis of cordierite from the Muzkol metamorphic complex, Tajikistan (wt.%).
Table 1. Chemical analysis of cordierite from the Muzkol metamorphic complex, Tajikistan (wt.%).
ComponentCordierite from the Muzkol Complex
123456789101112
SiO249.4449.5349.4049.3449.3749.5649.4649.4949.5449.4449.5549.44
TiO20.010.000.000.000.000.000.000.010.000.010.000.00
Al2O332.6832.6132.7032.7132.7532.7332.8632.8032.6932.7932.9332.79
FeO2.952.872.872.942.912.892.912.932.892.922.952.90
MnO0.020.030.030.030.030.040.030.020.030.030.030.04
MgO12.2112.3812.3912.3812.4912.5612.5212.5512.5312.4612.3812.40
CaO0.010.010.010.010.010.000.010.010.010.010.000.02
Na2O0.400.350.430.390.430.400.440.420.390.380.380.38
K2O0.000.000.000.000.000.000.000.000.000.010.000.00
Ʃ97.7297.7897.8397.8097.9998.1898.2398.2398.0898.0598.2297.97
Formula for 18 O atoms
Si5.005.014.994.994.995.004.994.995.004.994.994.99
Ti0.000.000.000.000.000.000.000.000.000.000.000.00
Al3.903.893.893.903.903.8903.903.903.893.903.913.90
Fe0.250.240.240.250.250.2400.250.250.240.250.2500.25
Mn0.0020.0030.0020.0020.0020.0030.0020.0020.0030.0030.0020.003
Mg1.841.871.8701.871.881.8901.881.891.881.881.861.87
Ca0.000.000.000.000.000.000.0010.0010.0010.0010.000.002
Na0.0780.0690.0840.0770.0840.0780.0850.0820.0770.0740.0740.075
K0.000.000.000.000.000.000.000.000.000.000.000.00
Ʃ11.07011.08211.07611.08911.10611.09711.10811.11411.0911.10111.08911.094
XMg0.8800.8850.8850.8810.8820.8860.8820.8820.8860.8810.8810.881
Temperature672690660674660672657664674680680678
Table 2. Composition of volatile components (based on 6 samples) released during mechanical destruction of cordierite from the Muzkol metamorphic complex, according to GC-MS analysis (rel.%).
Table 2. Composition of volatile components (based on 6 samples) released during mechanical destruction of cordierite from the Muzkol metamorphic complex, according to GC-MS analysis (rel.%).
NameMWSample Number (Numbering Corresponds to Figure 2)
123456
Aliphatic hydrocarbons: 3.543.963.694.312.372.05
Paraffins (CH4-C17H36)32–2401.571.861.741.831.100.96
Olefins (C2H2-C17H34)28–2381.972.101.952.481.281.09
Cyclic hydrocarbons: 0.681.761.860.710.340.38
Naphthenes (C6H10-C8H14)82–1100.060.170.110.070.030.03
Arenes (C6H6-C18H28)78–2040.621.581.730.620.290.35
PAH (C10H8-C11H10)128–142<0.010.010.030.010.01<0.01
Oxygenated organic compounds: 4.975.787.329.236.404.65
Alcohols, Esters and ethers (CH4O-C14H18O4)32–2500.841.721.041.881.180.53
Aldehydes (CH2O-C15H30O)44–2262.502.463.724.613.751.40
Ketones (C3H6O- C15H30O)58–2261.031.010.971.890.790.56
Carboxylic acids (C2H4O2-C13H26O2)60–2140.610.601.590.840.672.15
Heterocyclic organic compounds: 0.240.450.350.310.120.15
Dioxanes (C4H8O2)880.050.110.040.070.010.03
Furans (C5H6O-C15H26O)82–2220.190.340.310.240.110.12
Nitrogenated compounds:
(N2-C14H29NO)
28–1240.560.550.560.760.580.44
Sulfonated compounds:
(H2S-C10H16S)
34–1680.290.290.120.180.410.16
Chlorinated compounds:
(C5H8Cl2O2)
170<0.010.13<0.010.02<0.01<0.01
CO24432.0530.1229.0425.8026.3124.29
H2O1857.6857.0957.0658.9163.4967.88
Total number of components 165169166170172165
CO2/(H2O + CO2) 0.360.350.340.300.290.26
Alkanes/Alkenes 0.800.890.890.740.860.88
Note: MW—nominal mass.
Table 3. Composition of volatile components released during mechanical destruction of cordierite from the Muzkol metamorphic complex (Muz-1) and from pegmatite of the Kukhilal deposit (Kukh-1), Tajikistan [96], according to GC-MS analysis (rel.%).
Table 3. Composition of volatile components released during mechanical destruction of cordierite from the Muzkol metamorphic complex (Muz-1) and from pegmatite of the Kukhilal deposit (Kukh-1), Tajikistan [96], according to GC-MS analysis (rel.%).
NameMWMuz-1cMuz-1kKukh-1cKukh-1k
Aliphatic hydrocarbons: 3.542.051.273.44
Paraffins (CH4-C17H36)32–2401.570.960.521.34
Olefins (C2H2-C17H34)28–2381.971.090.752.10
Cyclic hydrocarbons: 0.680.380.170.29
Naphthenes (C6H10-C8H14)82–1100.060.030.020.04
Arenes (C6H6-C18H28)78–2040.620.350.140.24
PAH (C10H8-C11H10)128–142<0.01<0.010.010.01
Oxygenated organic compounds: 4.974.653.706.68
Alcohols, Esters and ethers (CH4O-C14H18O4)32–2500.840.530.601.04
Aldehydes (CH2O-C15H30O)44–2262.501.402.223.75
Ketones (C3H6O-C15H30O)58–2261.030.560.320.89
Carboxylic acids (C2H4O2-C13H26O2)60–2140.612.150.561.00
Heterocyclic organic compounds: 0.240.150.030.05
Dioxanes (C4H8O2)880.050.030.010.01
Furans (ethers) (C5H6O-C15H26O)82–2220.190.120.020.04
Nitrogenated compounds:
(N2-C14H29NO)
28–1240.560.440.110.40
Sulfonated compounds:
(H2S-C10H16S)
34–1680.290.160.020.04
Chlorinated compounds:
(C5H8Cl2O2, C16H33Cl)
170–240<0.01<0.010.030.06
CO24432.0524.295.523.50
H2O1857.6867.8889.265.60
Total number of components 165165166170
CO2/(H2O + CO2) 0.360.260.060.26
Alkanes/Alkenes 0.800.880.690.64
Note: Muz-1c is the center of cordierite crystal (sample 1); Muz-1k is the edge of a cordierite crystal (sample 6); Kukh-1c is the center of cordierite crystal; Kukh-1k is the edge of a cordierite crystal; MW is the nominal mass.
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

Zatolokina, K.I.; Tomilenko, A.A.; Bul’bak, T.A. Fluid Components in Cordierite from the Rocks of Epidote-Amphibole Facies of the Muzkol Metamorphic Complex, Tajikistan: Pyrolysis-Free GC-MS Data. Minerals 2023, 13, 323. https://doi.org/10.3390/min13030323

AMA Style

Zatolokina KI, Tomilenko AA, Bul’bak TA. Fluid Components in Cordierite from the Rocks of Epidote-Amphibole Facies of the Muzkol Metamorphic Complex, Tajikistan: Pyrolysis-Free GC-MS Data. Minerals. 2023; 13(3):323. https://doi.org/10.3390/min13030323

Chicago/Turabian Style

Zatolokina, Ksenia Igorevna, Anatoly Alexeyevich Tomilenko, and Taras Alexandrovich Bul’bak. 2023. "Fluid Components in Cordierite from the Rocks of Epidote-Amphibole Facies of the Muzkol Metamorphic Complex, Tajikistan: Pyrolysis-Free GC-MS Data" Minerals 13, no. 3: 323. https://doi.org/10.3390/min13030323

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