Next Article in Journal
Blending Characterization for Effective Management in Mining Operations
Previous Article in Journal
Effect of Mixing Water Temperature on the Thermal and Microstructural Evolution of Cemented Paste Backfill in Underground Mining
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluid Components in Cordierites from Granulite- and Amphibolite-Facies Rocks of the Aldan Shield and Yenisei Ridge, Russia: Evidence from Pyrolysis-Free GC-MS, Raman, and IR Spectroscopy

1
Sobolev Institute of Geology and Mineralogy, Siberian Branch of the Russian Academy of Sciences, Novosibirsk 630090, Russia
2
Trofimuk Institute of Petroleum Geology and Geophysics, Siberian Branch of the Russian Academy of Sciences, Novosibirsk 630090, Russia
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(9), 890; https://doi.org/10.3390/min15090890
Submission received: 17 June 2025 / Revised: 18 August 2025 / Accepted: 19 August 2025 / Published: 22 August 2025
(This article belongs to the Section Mineral Geochemistry and Geochronology)

Abstract

This study provides the first comprehensive characterization of fluid components in cordierites from both moderate- to high-pressure granulite facies of the Aldan Shield (Sutam and Nimnyr blocks), and granulite–amphibolite facies of the Yenisei Ridge (Kan and Yenisei series of the Angara–Kan complex), Russia, using integrated infrared and Raman spectroscopy coupled with pyrolysis-free gas chromatography–mass spectrometry (GC-MS). Granulite-facies cordierites record CO2-dominated fluids (XCO2 = CO2/(H2O + CO2) = 0.74–0.99) with elevated values (XCO2 = 0.89–0.99) in high-pressure, high-temperature (high-P-T) samples from the Sutam block and Kan series compared to moderate-P-T samples from the Nimnyr block (XCO2 = 0.74–0.84). Amphibolite-facies cordierites (Yenisei series) show significantly lower CO2 contents (XCO2 = 0.51–0.57) and higher H2O concentrations relative to high-pressure granulites. Critically, we report the first identification in cordierites of at least 12 homologous series of organic compounds and nitrogenated, sulfonated, and halogenated compounds. These results provide new constraints on fluid behavior across metamorphic facies transitions.

1. Introduction

Nowadays, the determination of fluid composition and regime in the formation of metamorphic minerals is one of the key aspects in metamorphic petrology [1,2,3]. During the last few decades, cordierite has been used to study this issue since it is a widespread mineral with a high thermobaric stability over a wide range of P-T conditions [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18]. Cordierite is a typical mineral for epidote–amphibolite [7,19,20], amphibolite [20,21], and granulite [22,23] facies, hornfels [24], granites [25], pyrometamorphic rocks [26,27], granitic pegmatites [28,29,30,31,32,33,34,35,36], lunar rocks [37,38,39], and meteorites [40,41,42,43].
The structure of cordierite features a framework of Si-Al tetrahedrons that form hexagonal rings in the (x;y) plane and vertical channels with a radius of 1.4 Å and large isometric cavities with a radius of 2.2 Å along the z direction [44]. These cavities harbor volatile components such as Ar, He, Ne, N2, H2, CO2, H2O, CO, O2, H2S, CH4, and other hydrocarbon molecules in a wide range [1,7,36,45,46,47,48,49,50,51,52,53,54,55,56,57]. The presence of hollow channels and the emplacement of molecules within them have been confirmed by various spectroscopic and diffraction techniques, including IR spectroscopy, neutronography, NMR, etc. [4,49,51,58,59,60,61,62,63,64,65].
Until recently, the compositions of volatiles in structural voids and channels of cordierites were mainly determined using Fourier-transform spectroscopy and Raman spectroscopy in combination with gas chromatography. It has been shown that water and carbon dioxide are the dominant fluid components in cordierites [1,4,5,6,45,46,47,55,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81]. In addition to water and carbon dioxide, other gases have also been detected in natural cordierites using mass spectrometric and chromatographic methods: H2, CO, N2, He, Ne, Ar, and H2S [1,82,83,84]. Zimmermann’s studies [45,46,47] revealed hydrocarbons upon heating natural cordierites up to 420 °C using mass spectrometry. Similarly, methane was first discovered at T = 1100 °C by Beltrame et al. [48]. Alpine cordierites also contain alkane-type hydrocarbons [52].
Recently, in addition to Raman and IR spectroscopy, non-pyrolytic gas chromatography–mass spectrometry has been applied to study the composition of volatile components preserved in cavities, channels, and fluid inclusions in cordierites. Channels and fluid inclusions in cordierites are from the pegmatites of the Kuhilal deposit [36], the epidote–amphibolite facies of the Muzkol metamorphic complex, Tajikistan [7], the pegmatites of the Middle Urals (Russia), and Dolni Bory (Czech Republic) [85]. In addition to water and carbon dioxide, aliphatic hydrocarbons (paraffins and olefins), cyclic hydrocarbons (arenes, cycloalkanes, and cycloalkenes), oxygenated hydrocarbons (alcohols, esters and ethers, aldehydes, ketones, and carboxylic acids), and nitrogenated and sulfonated compounds were detected for the first time in these studied cordierites.
The primary objective of this study was to perform a comprehensive investigation of volatile components in cordierites from the Aldan–Stanovoy Shield and the South Yenisei Ridge using advanced analytical techniques, including infrared (IR) and Raman spectroscopy, as well as pyrolysis-free gas chromatography–mass spectrometry.

2. Geology

2.1. Aldan–Stanovoy Shield

The Aldan–Stanovoy Shield is the largest area of the early Precambrian basement of the Siberian Platform located along its southern margin. It extends over a distance of more than 1200 km with a width of up to 600 km from the Sea of Okhotsk in the east to the Vitim River in the west. To the north, it is covered by sedimentary formations of the platform cover. To the south, the Shield is bounded by the Mongol-Okhotsk Fold Belt with tectonic unconformity. Two granite–greenstone regions are distinguished on the Shield: the eastern Batomg region (Paleoproterozoic age) and the western Olekminsk region (Mesoproterozoic age). The Aldan granulite–gneiss terrane is exposed between these regions, having been comprehensively described in key publications [86,87,88,89,90,91]. This terrane is subdivided into several lower-rank blocks that differ from each other both in their assemblages of metamorphic formations and metamorphic regimes (Figure 1) [92]. Major components of granulite sequences include undivided formations represented by granitoid, enderbite, and charnockite gneisses (infra-crustal complex) and stratified volcano–sedimentary rocks (supracrustal complex) [93].
In modern understanding [89,90], these blocks are combined into the Central Aldan and Eastern Aldan superterranes (Figure 1) [94]. The boundaries of the blocks are tectonic, often thrust-faulted, with the displacement plane participating in folding, making it difficult to decipher [86]. The petrological aspects of the formation of strata in the Aldan Shield have been extensively discussed in numerous studies [88,95,96,97,98,99], which emphasize the complex interrelationships of ongoing metamorphic processes.
The Central Aldan superterrane incorporates the Nimnyr, Melemken, Seym (Uchur), and Sutam blocks [94]. For our investigations, we selected the Nimnyr and Sutam blocks due to their most contrasting P-T regimes of metamorphism. The Nimnyr block is characterized by an approximately equal distribution of supracrustal and infracrustal complex formations. The Sutam block is predominantly composed of supracrustal sequences. A characteristic feature of these sequences is the presence of interlayered aluminous gneisses, representing metamorphosed pelitic analogues. Samarium–neodymium (Sm-Nd) model ages for metapelite protoliths (source rocks) from the Kurumkan Formation (Nimnyr block) range from 2.85 to 3.06 Ga, while those for paragneisses of the Sutam block yield 2.5–2.9 Ga [94]. Typical cordierite-bearing assemblages in the Sutam block correspond to the hypersthene–sillimanite subfacies, whereas the Nimnyr block exhibits cordierite–hypersthene subfacies of granulite-facies metamorphism, with P-T conditions of T = 825–950 °C at P = 8–11 kbar [88,96,100] and T = 770–830 °C at P = 5–6 kbar [88,99,100], respectively.
Cordierite samples were collected from bedrock outcrops along the Sutam River (Sutam block: samples 99–77, 173–77, 174–77, 176–77) (Figure 1) and the Aldan–Timpton Rivers (Nimnyr block: samples 39–80, 42–80, 78–80).

2.2. South Yenisei Ridge

The South Yenisei Ridge (Angara–Kan block) is located in the central part of the southwestern margin of the Siberian Platform [101,102]. Crystalline rocks of the South Yenisei Ridge were first systematically studied by Yu.A. Kuznetsov between 1937 and 1941. Summarizing these studies [103,104], he identified both metamorphic and magmatic complexes whose geological significance remains valid until today. The close spatial association of lithological units with tectonic contacts and sharp gradients of metamorphic transformations indicates an extremely complex tectonic evolution of the region. The Early Precambrian strata are grouped into two major metamorphic complexes (Figure 2): the Kan complex (dominant), with granulite-facies metamorphism, and the Yenisei complex, with amphibolite-facies metamorphism. Structural–geological reconstructions by E.K. Kovrigina [105], based on geological survey data synthesis, demonstrate that the South Yenisei Ridge comprises a series of seven imbricated blocks with submeridional to north–northwestern strikes. Later, A.D. Nozhkin [106] developed a geological scheme subdividing the granulite complexes into three geoblocks distinguished by their architecture, lithological composition, and charnockitoid distribution patterns [107]. The metamorphic conditions of the Kan and Yenisei complexes correspond to T = 850–900 °C at P = 8.5–9.0 kbar and T = 630–770 °C at P ≤ 7.0 kbar, respectively [99,100,108,109,110,111,112,113].
Rocks of the Kan complex predominate in the block’s architecture. Their metapelites contain rare spinel + quartz and corundum + quartz assemblages, whose stable formation (particularly the latter) under granulite-facies conditions remains debated in the literature.
For example, G.G. Lepezin and coauthors [109] propose three possible explanations: (1) several stages of metamorphism, comparable in temperature but different in pressure, (2) the block structure of the region with varying formation depths of metamorphic assemblages, and (3) partial reversibility of displaced-equilibrium reactions. They concluded that the primary cause of this heterogeneity is the third factor—reversibility of displaced-equilibrium reactions. L.L. Perchuk and colleagues [110] explain the pressure differences by the region’s block structure, showing that the original thickness of the rock sequences and subsequent retrograde metamorphic processes sufficiently account for their calculated pressure gradient of 10 to 7 kbar.

3. Materials and Methods

3.1. Sampling and Preparation

Polished sections were prepared from cordierite samples from moderate- to high-pressure granulite facies (Aldan Shield) and granulite–amphibolite facies (Yenisei Ridge) for comprehensive analysis, including infrared (IR) spectroscopy, Raman spectroscopy, and fluid inclusion studies. To ensure analytical purity for GC-MS measurements, all cordierite samples were prepared without any processing steps involving organic compounds or chemical treatments.
A representative area of the cordierite plate—free of alteration products, twinning, and structural blockiness, with uniform extinction under cross-polarized light—was carefully selected using an optical microscope and extracted. One aliquot was subjected to gas chromatography–mass spectrometry (GC-MS) analysis, while another aliquot was used for Raman and infrared (IR) spectroscopy, as well as X-ray spectral analysis.
Detailed microscopic examination of cordierite polished sections revealed no detectable fluid or melt inclusions in any analyzed samples.

3.2. Electron Microprobe Method

The major element content in cordierite was determined using electron-probe microanalyzer JEOL JXA-8100 (Japan Electronics Optics Laboratory, Tokio, Japan) equipped with five wavelength spectrometers via the WDS method at the Analytical Center for multi-elemental and isotope research SB RAS. Measurements were conducted under an accelerating voltage of 20 kV, probe current ranging from 50 to 250 nA depending on the mineral being measured and the task requirements, signal accumulation time of 60–120 s per peak, and background measurement duration of 30–60 s, with a probe size of 2 microns. Internal standards similar in composition to analyzed samples were used for monitoring stability and drift correction. These standards were remeasured every 30–40 measurements, followed by necessary corrections if required.

3.3. Infrared Spectroscopy

Infrared spectra of cordierite samples were recorded using a Vertex 70 FTIR Fourier-transform infrared spectrometer (Bruker Optics, Ettlingen, Germany) with diffuse reflection attachment and a Hyperion 2000 microscope. Samples were prepared as pressed pellets mixed with KBr powder. The KBr spectrum was subtracted from sample spectra using the OPUS version 5.5 software package. The spectral range covered was 370–4000 cm−1.

3.4. Raman Spectroscopy

Gas phase composition inside channels of cordierite crystals was investigated using Raman spectroscopy performed on a Horiba LabRam HR 800 spectrometer (Torus, Laser Quantum, Telford, UK). A solid-state Nd-YAG laser was used with a wavelength of 532 nm and power output of 75 mW. Signal detection utilized an Endor semiconductor detector cooled according to the Peltier method. A confocal microscopy system based on the Olympus BX-41 optical microscope with high numerical aperture objective lens × 100 was employed for precise localization of target points within the specimen. Backscattering geometry was adopted during analysis. The confocal diaphragm size varied depending on object dimensions: minimum size—30 nm (for objects measuring 5–10 μm); maximum size—300 nm (for larger than 100 μm objects). Raman spectra were obtained across the range of 100–4200 cm−1. Signal accumulation times ranged from 25 s/spectral window for large objects up to 400 s/spectral window for smaller ones.

3.5. Pyrolysis-Free Gas Chromatography–Mass Spectrometry with Mechanical Sample Destruction

The bulk composition of gas components from structural cavities and channels in cordierite as well as its inclusions was determined by a pyrolysis-free gas chromatography–mass spectrometry method (GC-MS) using mechanical sample destruction on the DSQ II MS/Focused GC instrument (Thermo Scientific, Austin, TX, USA). A freshly cleaved sample volume of up to 0.06 cm3 was placed into a special device connected online to the gas path of the chromatograph before the analytical column. The sample was then heated at T = 140–160 °C for 133 min under helium carrier gas flow (purity 99.9999%, initial pressure 45 kPa). Separation of gaseous mixture components took place on a capillary analytical column Rt-Q-BOND (Restek Corporation, Bellefonte, PA, USA) (stationary phase—100% divinylbenzene, length—30 m, internal diameter—0.32 mm, stationary phase thickness—10 μm). The gas mixture was introduced through a thermostated valve (270 °C) (VICI Valco Instruments, Houston, TX, USA) into the analytical column, with the constant He flow rate being 1.7 mL/min and temperature of the GC-MS connection line set at 300 °C. The column was held for 2 min at T = 70 °C, followed by heating at a rate of 25 °C/min until reaching 150 °C, further heating at 5 °C/min until 290 °C, and holding this temperature for 100 min.
Mass spectra obtained via electron impact ionization were recorded in full scan mode by a quadrupole mass-selective detector. Mass spectral conditions included electron energy—70 eV, emission current—100 µA, ion source temperature—200 °C, amplifier voltage—1350 V, positive ion polarity, scanning range—5–500 amu, scanning speed—one scan per second. The analysis start synchronized with the moment of sample destruction.
Relative concentrations of volatile components within the separated mixture were established by normalization: the total area of all chromatographic peaks equated to 100%, while the individual component content was calculated based on peak area size. This methodology is detailed in works [7,36,114].

4. Results

Cordierites selected for the study originated from granulite-facies rocks of the Sutam block (shore outcrops along Sutam River) and Nimnyr block (shore outcrops along Aldan and Timpton Rivers) of the Aldan Shield (Figure 1) and granulite (Kan series) and amphibolite (Yenisei series) facies of Yenisei Ridge (Figure 2). Using an optical microscopy, we selected samples without fluid or melt inclusions.

4.1. Cordierites from Granulite-Facies Rocks of Sutam and Nimnyr Blocks of the Aldan Shield

Electron probe microanalysis data showed minor statistically insignificant variations in chemical compositions of cordierites sampled from Sutam and Nimnyr blocks of the Aldan Shield (Figure 1, Table 1). The values of XMg (XMg = Mg/(Fe + Mn + Mg)) in cordierites vary from 0.86 (Sutam block) to 0.59 (Nimnyr block).
Raman spectroscopy revealed molecular water, carbon dioxide, nitrogen, and heavier-than-methane aliphatic hydrocarbons in isolated voids of cordierites from both Sutam and Nimnyr blocks (Figure 3).
Infrared spectroscopic studies of cordierites from these blocks (Table 2, Figure 4 and Figure 5) indicate the presence of water and carbon dioxide in isolated cavities. Moreover, bands corresponding to valence vibrations of C-H bonds in CH2- and CH3- groups were detected at maxima of 2854, 2872, 2926, and 2956 cm−1.
The results of the gas chromatography–mass spectrometry (GC-MS) demonstrate that cordierites from the Sutam block contain between 226 and 261 compounds (Figure 6, Table 3, Supplementary Tables S1–S4). Carbon dioxide and water are predominant volatile components in cordierites of the Sutam block (Table 3, Supplementary Tables S1–S4). The concentration of the carbon dioxide ranges from 86.7 to 97.5 rel.%, while water content varies between 1.3 and 10.9 rel.%. The values of hydrocarbon derivatives (aliphatic, cyclic, oxygenated hydrocarbons) and chlorinated, nitrogenated, and sulfonated compounds span from 1.0 to 4.6 rel.% (Table 3, Supplementary Tables S1–S4).
Aliphatic hydrocarbons consist of paraffins (CH4-C17H36) and olefins (C2H2-C17H34), ranging from 0.1 to 0.5 rel.% and from 0.1 to 0.8 rel.%, respectively. The amount of light (C1-C4) paraffins is less than 0.1 rel.%, medium (C5-C12) 0.1–0.3 rel.%, and heavy (C13-C17) up to 0.2 rel.% (Table 3, Supplementary Tables S1–S4).
The value of the cyclic hydrocarbons is around 0.1–0.3 rel.%. The maximum concentration of the naphthenes and cycloalkenes reaches 0.1 rel.%. The concentration of the arenes is in the range of 0.1–0.2 rel.%, while for polycyclic aromatic hydrocarbons (PAHs) it is less than 0.1 rel.% (Table 3, Supplementary Tables S1–S4).
The concentration of the oxygenated hydrocarbons varies from 0.7 to 3.1 rel.%, whereas the content of the alcohol is less than 0.3 rel.%, esters and ethers from 0.1 to 0.2 rel.%, aldehydes and ketones ranging from 0.1 to 1.4 rel.%, and carboxylic acids from 0.3 to 0.8 rel.% (Table 3, Supplementary Tables S1–S4).
The amount of the sulfonated compounds (H2S-C13H22S) is up to 0.1 rel.%, while nitrogenated compounds (N2-C13H16F3NO) are in a wide range from 0.1 to 0.9 rel.% (Table 3, Supplementary Tables S1–S4).
The GC-MS analysis of cordierites from the granulite-facies rocks in the Nimnyr block revealed the presence of between 246 and 261 distinct compounds. (Figure 7, Table 4, Supplementary Tables S5–S8). The predominant volatiles are carbon dioxide and water (Table 3, Supplementary Tables S1–S4). However, the concentration of the carbon dioxide is significantly lower compared to samples from the Sutam block (ranging from 66.9 to 80.0 rel.%), while water content is higher (from 14.8 to 23.9 rel.%) (XCO2 = 0.74 − 0.84). As in Sutam block cordierites, nitrogenated, sulfonated,, and halogenated compounds were also detected. The content of the hydrocarbon derivatives in cordierites from the Nimnyr block is slightly higher than in the Sutam block, ranging from 4.0 to 10.9 rel.% and 1.0 to 4.6 rel.%, respectively. Nitrogenated compounds vary from 1.0 to 1.8 rel.%, while sulfonated compounds remain in the range from 0.1 to 0.3 rel.% (Table 4, Supplementary Tables S5–S8).
Aliphatic, oxygenated, cyclic, and heterocyclic hydrocarbons comprise 12 homologous organic compounds. Aliphatic hydrocarbons include paraffins (CH4-C17H36) (from 0.2 to 3.3 rel.%) and olefins (C2H2-C17H34) (from 0.6 to 1.5 rel.%) (Table 4, Supplementary Tables S5–S8). The concentrations of the paraffin and olefin compounds in cordierites from the Nimnyr block are higher than those in the Sutam block. Saturated hydrocarbons include light, medium, and heavy paraffins. The amount of light paraffins is less than 0.1 rel.%, medium paraffins vary from 0.2 to 0.4 rel.%, and heavy paraffins span from 0.1 to 3.1 rel.% (Table 4, Supplementary Tables S5–S8). We conclude that the content of the cyclic hydrocarbon in the cordierites of the Nimnyr block is generally higher compared to those from the Sutam block (from 0.3 to 0.8 rel.%).
Oxygenated hydrocarbons consist of five homologous series: (1) alcohols (CH4O-C12H22O2), (2) esters and ethers (C4H4O2-C16H19FO2), (3) aldehydes (CH2O-C16H32O), (4) ketones (C3H6O-C16H32O), and (5) carboxylic acids (C2H4O2-C14H28O2) (Table 4, Supplementary Tables S5–S8). In general, the content of the oxygenated hydrocarbons in the cordierites of the Nimnyr block is slightly higher than those in the Sutam block, ranging from 2.4 to 7.9 rel.%. The concentrations of the oxygenated hydrocarbons show relatively wide variations: alcohols (0.3–2.5 rel.%), ethers and esters (0.5–0.9 rel.%), aldehydes (0.7–2.4 rel.%), ketones (0.3–1.0 rel.%), and carboxylic acids (0.4–1.8 rel.%) (Table 4, Supplementary Tables S5–S8).
Sulfonated compounds in the Nimnyr block mainly include hydrogen sulfide and thiophenes (H2S-C13H22S), ranging from 0.1 to 0.3 rel.% (Table 4, Supplementary Tables S5–S8).
Nitrogenated compounds (N2-C13H16F3NO) in the Nimnyr block consist of molecular nitrogen, benzonitrile, pyridine, and related derivatives, ranging from 1.0 to 1.8 rel.% (Table 4, Supplementary Tables S5–S8).

4.2. Cordierites from Granulite-Facies Metamorphism of Kan Series, Yenisei Ridge

According to electron-probe microanalysis data obtained from granulites of the Kan series (Figure 2), we recognized minor statistical inconsistencies in their chemical composition (Table 5). The values of XMg are less than 0.72 (Table 5).
Raman spectroscopy showed that isolated cavities in cordierites from granulites of the Yenisei Ridge possess molecular water, carbon dioxide, nitrogen, and heavier-than-methane aliphatic hydrocarbons (Figure 8).
IR spectroscopy of cordierites from granulite-facies rocks of the Yenisei Ridge (Table 6, Figure 9 and Figure 10) proved the presence of water and carbon dioxide in isolated cavities. The additional IR-bands peaked at 2854, 2872, 2926, and 2956 cm−1, which is typical for the C–H bond in the functional groups of CH2- and CH3 (Figure 10).
According to GC-MS analysis, dominant volatile components in cordierites from granulite facies of the Kan series (Yenisei Ridge) were similar to those found in Sutam and Nimnyr blocks of the Aldan Shield, namely carbon dioxide and water (Figure 11, Table 7, Supplementary Table S9). The concentration of the carbon dioxide is 84.2 rel.%, which is comparable to the Sutam block (XCO2 = 0.94 vs. XCO2 = 0.89–0.94). In total, we found two hundred and fifty-seven volatile components. Moreover, we detected aliphatic, cyclic, oxygenated, and heterocyclic hydrocarbons along with their derivatives (9.3 rel.%) and chlorinated, nitrogenated, and sulfonated compounds (Table 7, Supplementary Table S9). Aliphatic hydrocarbons are predominantly paraffins (0.4 rel.%) and olefins (0.8 rel.%). The concentration of the light paraffins is negligible (<0.1 rel.%), medium paraffins—0.3 rel.%, and heavy hydrocarbons—0.2 rel.% (Table 7, Supplementary Table S9).
The content of the cyclic hydrocarbons is 0.8 rel.%, naphthene and cycloalkene—less than 0.1 rel.%, arenes—0.7 rel.%, and PAHs—0.1 rel.% (Table 7, Supplementary Table S9). The concentrations of the oxygenated hydrocarbons are 7.2 rel.%, alcohol about 0.8 rel.%, esters and ethers—1.5 rel.%, aldehyde—1.8 rel.%, ketone—1.5 rel.%, carboxylic acids—1.6 rel.%. The amount of the sulfonated compounds is 0.3 rel.%, while nitrogenated compounds is a bit higher at up to 1.1 rel.% (Table 7, Supplementary Table S9).

4.3. Cordierites from Amphibolite-Facies Metamorphism of Yenisei Series, Yenisei Ridge

According to electron-probe microanalysis data obtained from amphibolites of the Yenisei series (Figure 2), we detected minor statistical inconsistencies in their chemical composition (Table 8). The values of XMg are less than 0.72 (Table 8).
The results of the Raman spectroscopy indicate that isolated cavities of cordierites from amphibolite facies have molecular water, carbon dioxide, nitrogen, and heavier-than-methane aliphatic hydrocarbons (Figure 12).
IR spectroscopy (Table 6, Figure 9 and Figure 10) shows the presence of water and carbon dioxide in isolated cavities of studied cordierites from the Yenisei Ridge.
Results of GC-MS analysis of cordierites from amphibolite facies are presented in Figure 13, Table 7, and Supplementary Tables S10 and S11. Overall, we identified from 263 to 273 compounds. According to obtained data, dominant volatile components in cordierites from amphibolite facies resembled those from granulite facies of the Kan series, i.e., carbon dioxide and water (Table 7, Supplementary Tables S10 and S11). The content of water is considerably higher compared to granulite-facies cordierites (up to 42.3 rel.%), while the content of carbon dioxide is markedly lower, ranging from 43.6 to 45.1 rel.%. Furthermore, aliphatic, cyclic, oxygenated, and heterocyclic hydrocarbons along with their derivatives and chlorinated, nitrogenated, and sulfonated compounds were discovered. Specifically, the concentrations of hydrocarbon vary from 12.6 to 16.9 rel.%, nitrogenated compounds from 0.4 to 3.7 rel.%, and sulfonated compounds from 0.4 to 1.1 rel.% (Table 7, Supplementary Tables S10 and S11).
Aliphatic, oxygenated, cyclic, and heterocyclic hydrocarbons belong to 12 series of homologous organic compounds. Aliphatic hydrocarbons contain paraffins (1.1–1.8 rel.%) and olefins (2.6–3.2 rel.%) (Table 7, Supplementary Tables S10 and S11). The amount of light paraffin is less than 0.1 rel.%, medium from 0.18 to 0.44 rel.%, heavy up to 3.1 rel.% (Supplementary Tables S10 and S11).
The content of cyclic hydrocarbons (Table 7, Supplementary Tables S10 and S11) varies from 1.3 to 1.3 rel. %. Cyclic hydrocarbons include naphthenes and cycloalkenes, arenes, and PAHs. The total amount of naphthenes and cycloalkenes is less than 0.1 rel. %. The content of aromatic compounds is in the narrow range, from 1.1 to 1.2 rel. %. The proportion of PAHs changes from 0.1 to 0.2 rel. %.
Oxygenated hydrocarbons constitute 7.4–10.4 rel.% of the total volatile components, with alcohols accounting for ~0.9 rel.%, ethers and esters for 0.6–0.9 rel.%, aldehydes for 3.3–3.7 rel.%, ketones for 1.2–1.9 rel.%, and carboxylic acids for 1.5–3.1 rel.% (Table 7; Supplementary Tables S10 and S11).
Cordierites from the amphibolite facies of the Yenisei Ridge possess sulfonated compounds mainly represented by hydrogen sulfide and thiophenes. The proportion of sulfonated compounds (H2S-C12H20S) varies from 0.4 to 1.1 rel.%. The ratio of nitrogenated compounds (N2-C16H24FNO) changes from 0.4 to 3.7 rel.% (Table 7, Supplementary Tables S10 and S11).

5. Discussion

In this paper, we applied a comprehensive approach (GC-MS, Raman, and IR spectroscopy) to analyze volatile components in cordierites from granulite and amphibolite facies of the Aldan Shield and Yenisei Ridge. According to Raman and IR spectroscopy data, isolated cavities in studied cordierites contain molecular water and carbon dioxide. Based on Raman and IR spectroscopy data, we identified nitrogen and aliphatic hydrocarbons in cordierites. Gas chromatography–mass spectrometry analysis allowed us to recognize a wide variety of volatile components (up to 273 compounds). Using the data obtained by GC-MS analysis, we determined carbon dioxide, water, non-oxygenated aliphatic and cyclic hydrocarbons (paraffins, olefins, naphthenes and cycloalkenes, arenes, and PAHs), oxygenated hydrocarbons (alcohols, esters and ethers, aldehydes, ketones, and carboxylic acids), and heterocyclic compounds (furans, dioxins, and dioxanes), as well as nitrogenated and sulfonated compounds hosted by studied cordierites. Carbon dioxide and water are major volatile components and their concentrations can vary significantly depending on the conditions of metamorphism.
Carbon dioxide and water. Cordierites from high-temperature–pressure granulite facies (Sutam block of the Aldan Shield and Kan series of the Angara–Kan complex of the Yenisei Ridge) are characterized mainly by carbonic fluid (Figure 14) with CO2 reaching up to 97.6 rel.% (XCO2 = CO2/(H2O + CO2) from 0.89 to 0.99), while the water concentration remains low (1.3–10.8 rel.%).
Moderate-to-low-temperature–pressure granulites (Nimnyr block of the Aldan Shield) are characterized by a gradual change in the composition of fluids (Figure 13): CO2 decreases to 80 rel.% (XCO2 = 0.74–0.84) whereas water increases to 14.8–24.1 rel.%.
Significant changes occur in the amphibolite facies (Yenisei series of the Yenisei Ridge), where CO2 drops sharply to 45.1 rel.% (XCO2 = 0.51–0.57) and water rises to 42.3 rel.% (Figure 14).
Cordierites from epidote–amphibolite facies (Muzkol metamorphic complex) are dominated mainly by aqueous fluids: H2O = 57.1–67.9 rel.%, CO2 = 24.3–32.9 rel.%, XCO2 = 0.26–0.36 [7].
Hydrocarbons and their derivatives. Besides water and carbon dioxide, among the detected volatiles in cordierites from granulite and amphibolite facies of the Aldan Shield and Yenisei Ridge, we found at least 12 homologous organic compounds. These compounds include non-oxygenated aliphatic and cyclic hydrocarbons (paraffins, olefins, naphthenes and cycloalkenes, arenes, and PAHs), oxygenated hydrocarbons (alcohols, esters and ethers, aldehydes, ketones, and carboxylic acids), and heterocyclic compounds (furans and dioxanes). Previous studies [54,77,115] also reported hydrocarbons, but only light species (C1–C5) were identified.
The structure of cordierite contains channels along axis [001] with isometric cavities whose maximum dimensions reach approximately 5.4 Å along [010] and 6.0 Å along [99], narrowing down to 2.5–2.8 Å [44,51]. The ratio of critical diameters (CDs) of molecules to structural cavity sizes suggests localization of simpler compounds (for instance, CD of CH4—3.8 Å, CO—2.8 Å, CO2—3.1 Å, C2H4—4.0 Å, 1-butene—4.9 Å, and acetone—5.6 Å) within mineral channels. It has been determined that water, argon, CO2, and N2 can localize in these cavities undergoing deformation [53,58]. Given the size ratios between cavities and channels in cordierite structures and critical sizes of simpler members of discovered homologous volatile rows, it is suggested they may be localized in these cavities and channels, while larger organic compounds likely reside in nonstructural positions in submicron inclusions [7,36,85,116,117,118].
At high temperatures and pressures corresponding to peak metamorphic events, the fluid consists not of individual substances but of ionized particles, including H+, OH, CH+, CH2+, CH3+, N+, C=O2−, RCOO, etc. The growing cordierite captures these ionized particles in structural channels and cavities, as well as in non-structural nano-sized primary inclusions. The degree of filling of channels and cavities with volatiles is maximal at peak temperature and pressure. As temperature decreases, compounds stable at corresponding parameters form from these ionized particles. One might assume, based on the ratio of volatile molecule sizes to channel diameter (2.8 Å), that only water could exchange in cordierite during retrograde metamorphic stages or throughout a sample’s geological history. However, this is prevented by large alkali metal cations (Na+, K+, Li+) that block these channels, along with high activation energies and low water diffusion coefficients in cordierite [3,119].
Hydrocarbon chains cannot penetrate through the openings, but they can be assembled in the cordierite channel during its crystallization from CH2+ and CH3+ fragments. The resulting hydrocarbon molecule would be deformed, like all known molecules in the cordierite structure (H2O, CO2, Ar, N2). The second form of localization for higher-molecular-weight hydrocarbons is nano-inclusions in cordierite. Here, we rely on works [116,117,118], where such inclusions were discovered and their composition was studied. We maintain that submicroscopic inclusions in cordierite must be present, as evidenced by low-temperature maxima on kinetic curves of natural cordierites (epidote–amphibolite, amphibolite, and granulite facies) obtained by us previously using mass spectrometry [1]. We draw on previous work by our colleagues that discovered solid nitrogen in nano-inclusions in diamond [120], which was possible using transmission electron microscopy (TEM) precisely because the nitrogen in inclusions was solid. Applying TEM to cordierite to detect inclusions ranging from nano-sized to submicron is impractical due to the non-solid state of inclusion contents.
High-temperature–high-pressure granulites (Sutam block of the Aldan Shield and Kan Series, Yenisei Ridge) are characterized by relatively low contents of hydrocarbon compounds (from 1.0 to 4.6 rel.%). In contrast, low-temperature-and-pressure granulites (Nimnyr block, Aldan Shield) show a significantly higher hydrocarbon concentration (3.5–10.9 rel.%) (Figure 13). Cordierites from the amphibolite facies (Yenisei series, Yenisei Ridge) exhibit the highest hydrocarbon content (from 12.6 to 16.9 rel.%). Thus, the hydrocarbon content decreases systematically from 16.9 to 1.0 rel.% with an increasing metamorphism grade.
The hydrocarbon content of the cordierites from the epidote–amphibolite facies (Muzkol metamorphic complex) varies from 7.2 to 14.4 rel.% (Figure 14).
Nitrogenated and sulfonated compounds. In addition to dominant volatiles, cordierites from granulite- and amphibolite-facies rocks contain nitrogenated (molecular nitrogen and other nitrogenated compounds like acetonitrile, propanenitrile, pyridine, acetamide, etc.) and sulfonated compounds (hydrogen sulfide, carbonyl sulfide, sulfur dioxide, carbon disulfide, dimethyl sulfide, thiophene, etc.). We also detected halogenated compounds such as 1-chlorobutane, 1-chloroundecane, 2-chloroethoxybenzene, fluorocyclohexane, p-fluoroethylbenzene, 3-fluoro-o-xylene, 5-fluoro-m-xylene, and others. The studied cordierites contain halogenated components, particularly C8H9F isomers (3-fluoro-o-xylene, 5-fluoro-m-xylene). The presence of other F- and Cl-bearing volatiles reflects individual characteristics of their mineral-forming environment.
Cordierites from high-temperature–high-pressure granulites (Sutam block, Aldan Shield; Kan series, Yenisei Ridge) contain trace nitrogen-bearing compounds (0.1–1.1 rel.%) and only negligible concentrations of sulfonated components (<0.3 rel.%) (Figure 13). In moderate-to-low-temperature–pressure granulites (Nimnyr block of the Aldan Shield), nitrogenated compounds show increased concentrations (0.8–1.8 rel.%).
Cordierites from the amphibolite facies (Yenisei series, Yenisei Ridge) show maximum concentrations of both nitrogenated (up to 3.7 rel.%) and sulfonated (0.4–1.1 rel.%) compounds, while those from the epidote–amphibolite facies (Muzkol metamorphic complex) contain lower amounts of nitrogenated compounds (0.4–0.8 rel.%) and sulfonated compounds (0.1–0.4 rel.%) [7].
Our study, combined with previous published results [7], demonstrates a systematic evolution in fluid composition with an increasing metamorphic grade: from aqueous-dominated fluids (XCO2 = 0.26) in epidote–amphibolite facies to mixed water–carbonate in amphibolite facies (XCO2 = 0.51), and finally to granulite facies rich in carbon dioxide (XCO2 = 0.99) (Figure 14 and Figure 15). As the degree of metamorphism increases, oxidative–reductive conditions of mineral formation shift toward highly oxidized states [H/(O + H)—from 0.59 for rocks of the amphibolite facies to 0.05 for rocks of the high-temperature–high-pressure granulite facies) (Table 3, Table 4, and Table 7).
Our data confirmed that cordierite serves as a unique natural “fluid-phase sample collector” reliably capturing the evolution of volatile component compositions during variations in P-T conditions of mineral formation. The application of pyrolysis-free gas chromatography–mass spectrometry with shock sample destruction provides new opportunities for reconstructing fluid regimes of metamorphism for different geodynamic settings.
This study also indicates that previous analytical determinations of H2O content in cordierite (Penfield, DTA, Karl Fischer titration, and ion-probe techniques) may require reevaluation. Previous calibrations of water content in cordierite did not account for multiple volatile phases revealed in our investigation or their contribution to thermal behavior of the mineral, leading to substantial distortions in interpreting mass loss upon heating.

6. Conclusions

Comprehensive GC-MS, Raman, and IR spectroscopy studies of cordierites demonstrated that metamorphic fluids of the Aldan Shield (Sutam and Nimnyr blocks) and Yenisei Ridge (Kan and Yenisei series of the Angara–Kan complex) formed a complex multicomponent mixture of C-H-O-S-N±Hal elements dominated either by H2O or CO2 depending on metamorphic conditions.
We proved that cordierites from granulite-facies rocks developed primarily under the influence of carbon-dioxide-dominated fluid (XCO2 = CO2/(H2O + CO2) ranging from 0.74 to 0.99). Thus, cordierites from high-temperature–high-pressure granulites (Sutam block of the Aldan Shield and Kan series of the Angara–Kan complex, Yenisei Ridge) possess higher CO2 contents (XCO2 = 0.89–0.99) compared to those from moderate-to-low-temperature–pressure granulites of the Nimnyr block of the Aldan Shield (XCO2 = 0.74–0.84).
The dominant fluid components in cordierites from amphibolite-facies rocks (Yenisei series, Angara–Kan complex, Yenisei Ridge) are carbon dioxide and water. However, the CO2 content (XCO2 = 0.51–0.57) is significantly lower, while the water content is correspondingly higher compared to cordierites from granulite-facies rocks of the Aldan Shield and Yenisei Ridge.
For the first time, we have detected no fewer than twelve homologous series of organic compounds in studied cordierites, in addition to water and carbon dioxide. These compounds are non-oxygenated aliphatic and cyclic hydrocarbons (paraffins, olefins, cyclic alkanes and alkenes, arenes, and PAHs), oxygenated hydrocarbons (alcohols, esters and ethers, aldehydes, ketones, and carboxylic acids), and heterocyclic hydrocarbons (furans, dioxins, and dioxanes).
Additionally, nitrogenated compounds such as acetonitrile, propanenitrile, pyridine, acetamide, etc. (0.1–3.7 rel.%), sulfonated compounds including hydrogen sulfide, carbonyl sulfide, sulfur dioxide, carbon disulfide, dimethyl sulfide, thiophene, etc. (0.04–1.1 rel.%), and halogenated compounds such as 1-chlorobutane, 1-chloroundecane, 2-chloroethoxybenzene, fluorocyclohexane, p-fluoroethylbenzene, 3-fluoro-o-xylene, 5-fluoro-m-xylene, etc., were established in volatile components of cordierites from granulite and amphibolite facies. Widespread halogenated components for all studied cordierites are C8H9F isomeres (3-fluoro-o-xylene, 5-fluoro-m-xylene); other detected F- and Cl-containing volatiles reflect specific features of the mineral-forming medium.
We demonstrated that as P-T parameters of metamorphism increase from epidote-amphibolite through amphibolite to granulite facies, the composition of mineralizing fluid evolves from predominantly aqueous for the epidote–amphibolite facies (XCO2 = 0.26) [7] to water–carbonate-dominant for the amphibolite facies (XCO2 = 0.51), and ultimately to granulite facies of metamorphism that are largely carbon-dioxide-rich (XCO2 = 0.99). Simultaneously, with increasing degrees of metamorphism, redox conditions of mineral formation also change toward more oxidized environments [H/(O + H)—from 0.59 for amphibolite-facies rocks to 0.05 for high-temperature–high-pressure granulite-facies rocks].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min15090890/s1, Table S1: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 99–77 from granulites of the Sutam block of the Aldan Shield (upper reaches of the Sutam River, species diversity of 226 components); Table S2: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 173–77 from granulites of the Sutam block of the Aldan shield (upper reaches of the Sutam River, species diversity of 237 components); Table S3: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 174–77 from granulites of the Sutam block of the Aldan shield (Sutam River, species diversity of 261 components); Table S4: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 176–77 from granulites of the Sutam block of the Aldan shield (Sutam River, species diversity of 254 components); Table S5: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 39–80 from granulites of the Nimnyr block of the Aldan shield (Aldan River, Kurumkan, species diversity of 258 components); Table S6: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 42–80 from granulites of the Nimnyr block of the Aldan shield (Aldan River, species diversity of 246 components); Table S7: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 78–80 from granulites of the Nimnyr block of the Aldan shield (Aldan River, species diversity of 261 components); Table S8: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 77–79 from granulites of the Nimnyr block of the Aldan shield (lower reaches of Timpton River, species diversity of 245 components); Table S9: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 16–21 from granulites of the Kan series of the Yenisei Ridge (species diversity of 257 components); Table S10: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite PN-1_02-06 from the amphibolite-facies rocks of the Yenisei series of the Yenisei Ridge (species diversity of 263 components); Table S11: Results of the GC-MS analysis of volatile components extracted by mechanical shock crushing from cordierite 02-06 from the amphibolite-facies rocks of the Yenisei series of the Yenisei Ridge (species diversity of 273 components).

Author Contributions

Conceptualization, K.Z.; methodology, T.B. and A.T.; investigation, K.Z. and T.B.; resources, A.T. and N.P.; data curation, K.Z.; writing—original draft preparation, K.Z.; writing—review and editing, T.B., N.P. and A.T.; visualization, K.Z. and A.T.; supervision, A.T. All authors have read and agreed to the published version of the manuscript.

Funding

The work is performed on state assignment of IGM SB RAS (№ 122041400312-2) and on state assignment of IPGG SB RAS (FWZZ-2022-0002).

Data Availability Statement

Data are contained within the article and 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, who began and actively participated in this work. We also sincerely thank the four anonymous reviewers whose insightful comments and constructive suggestions significantly improved the quality of this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lepezin, G.G.; Bul’bak, T.A.; Sokol, E.V.; Shvedenkov, G.Y. Fluid components in cordierites and their significance for metamorphic petrology. Russ. Geol. Geophys. 1999, 40, 99–116. [Google Scholar]
  2. 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]
  3. Bul’bak, T.A.; Shvedenkov, G.Y.; Ripinen, O.I. Kinetics of H2O–CO2 Molecular Exchange in the Structural Channels of (Mg, Fe2+)-Cordierite. Geochem. Int. 2005, 43, 386–394. [Google Scholar]
  4. Vry, J.K.; Brown, P.E.; Valley, J.W. Cordierite volatile content and the role of CO2 in hight–grade metamorphism. Am. Mineral. 1990, 75, 71–88. [Google Scholar]
  5. 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]
  6. 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]
  7. 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. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  10. 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]
  11. 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]
  12. 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]
  13. 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]
  14. Bul’bak, T.A.; Shvedenkova, S.V. Experimental study of the equilibrium of cordierite with potassium feldspar in water-carbon-dioxide fluid. Russ. Geol. Geophys. 1998, 39, 851–855. [Google Scholar]
  15. 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]
  16. 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]
  17. 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]
  18. 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]
  19. 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]
  20. Lepezin, G.G.; Melenevsky, V.N. On the problem of water diffusion in the cordierites. Lithos 1977, 10, 49–57. [Google Scholar] [CrossRef]
  21. Palin, R.M.; Palmer, Z.; Holm-Denoma, C.; Hernández-Uribe, D.; Hernández-Montenegro, J.D. On the occurrence of amphibolite-facies sapphirine, spinel, phlogopite, anorthite, and corundum in the Wet Mountains, Colorado, USA. Lithos 2023, 440–441, 107024. [Google Scholar] [CrossRef]
  22. Holdaway, M.J.; Lee, S.M. Fe-Mg cordierite stability in high-grade pelitic rocks based on experimental, theoretical, and natural observations. Contrib. Mineral. Petrol. 1977, 63, 175–198. [Google Scholar] [CrossRef]
  23. Winter, J.D. Principles of Igneous and Metamorphic Petrology, 2nd ed.; Pearson: Harlow, UK, 2014; 739p. [Google Scholar]
  24. 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]
  25. Phillips, G.N.; Wall, V.J.; Clemens, J.D. Petrology of the Strathbogie Batholith: A Cordierite-Bearing Granite. Can. Mineral 1981, 19, 47–63. [Google Scholar]
  26. Sokol, E.V.; Maksimova, N.V.; Nigmatulina, E.N.; Sharygin, V.V.; Kalugin, V.M. Combustion Metamorphism; SB RAN Publisher: Novosibirsk, Russia, 2005. [Google Scholar] [CrossRef]
  27. 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]
  28. Heinrich, E.W. Cordierite in pegmatite near Micanite, Colorado. Am. Mineral. 1950, 35, 173–184. [Google Scholar]
  29. Newton, R.C. BeO in pegmatitic cordierite. Mineral. Mag. 1966, 35, 920–927. [Google Scholar] [CrossRef]
  30. Černý, P.; Povondra, P. Beryllian cordierite from Vĕžná: (Na,K)+Be→Al. Neues Jahrb. Miner. Monatsh. 1966, 2, 36–44. [Google Scholar]
  31. Černý, P.; Povondra, P. Cordierite in West-Moravian desilicated pegmatites. Acta Univ. Carol. Geol. 1967, 3, 203–221. [Google Scholar]
  32. 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]
  33. 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]
  34. 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]
  35. 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]
  36. 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). Russ. Geol. Geophys. 2021, 62, 1411–1431. [Google Scholar] [CrossRef]
  37. 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]
  38. Herzberg, C.T.; Baker, M.B. The cordierite-to-spinel-cataclasite transition: Structure of the lunar crust. In Proceedings of the Conference on the Lunar Highlands Crust, Houston, TX, USA, 14–16 November 1979; Pergamon Press: New York, NY, USA, 1980; Volume 1, pp. 113–132. [Google Scholar]
  39. 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]
  40. Fuchs, L.H. Occurrence of cordierite and aluminous orthoenstatite in the Allende meteorite. Am. Mineral. 1969, 54, 1645–1653. [Google Scholar]
  41. 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]
  42. 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]
  43. 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]
  44. Schreyer, W. Experimental studies on cation substitutions and fluid incorporation in cordierite. Bull Mineral. 1985, 108, 273–291. [Google Scholar] [CrossRef]
  45. Zimmermann, J.L. Application petrogenetique de l’etude de la liberation de l’eau et du gaz carbonique des cordierites. CR Acad. Sci. 1972, 275, 519–522. [Google Scholar]
  46. Zimmermann, J.L. Etude par spectrometric de masse de la composition des fluides dans quelques cordierites du sud de la Norvege. In Reunion Annuelle des Sciences de la Terre; Societe Geologique de France: Paris, France, 1973. [Google Scholar]
  47. 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]
  48. 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]
  49. Goldman, D.S.; Rossman, G.R.; Dollase, W.A. Channel constituents in cordierite. Am. Mineral. 1977, 62, 1144–1157. [Google Scholar]
  50. Johannes, B.; Schreyer, W. Experimental introduction of CO2 and H2O into Mg- cordierite. Am. J. Sci. 1981, 281, 299–317. [Google Scholar] [CrossRef]
  51. Armbruster, T.; Bloss, F.D. Orientation and effects of channel H2O and CO2 in cordierites. Am. Mineral. 1982, 67, 284–291. [Google Scholar]
  52. Mottana, A.; Fusi, A.; Bianchi Potenza, 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]
  53. 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]
  54. 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]
  55. Kolesov, E.A.; Geiger, C.A. Cordierite II: The role of CO2 and H2O. Am. Mineral. 2000, 85, 1265–1274. [Google Scholar] [CrossRef]
  56. 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]
  57. Kolesov, B.A. Raman spectra of single H2O molecules isolated in crystal cavities. Zhurnal Strukturnoi Khimii 2006, 47, 27–40. [Google Scholar] [CrossRef]
  58. Smith, J.V.; Schreyer, W. Location of argon and water in cordierite. Miner. Mag. 1962, 33, 226–236. [Google Scholar] [CrossRef]
  59. Farrel, E.E.; Newnham, R.E. Electronic and vibrational absorption spectra. Am. Mineral. 1967, 52, 380–389. [Google Scholar]
  60. Armbruster, T.; Bloss, F.D. Channel CO2 in cordierites. Nature 1980, 286, 140–141. [Google Scholar] [CrossRef]
  61. 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]
  62. Mirwald, P. Crystal chemical effects of sodium on the incorporation of H2O and CO2 in Mg-cordierite. Terra Cogn. 1983, 3, 163. [Google Scholar]
  63. 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]
  64. 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]
  65. 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]
  66. 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. Doklady Akademii Nauk SSSR 1971, 200, 950–952. [Google Scholar]
  67. Kitsul, V.I.; Lazebnik, Y.D.; Brovkin, A.A.; Suknev, V.O. Diagram for determination of the iron content of cordierites. Doklady Akademii Nauk SSSR 1971, 200, 1419–1422. [Google Scholar]
  68. 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]
  69. 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]
  70. 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]
  71. 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]
  72. 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]
  73. 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]
  74. Perreault, S.; Martignole, J. High-temperature cordierite migmatites in the north-eastern Grenville Province. J. Metamorph. Geol. 1988, 6, 673–696. [Google Scholar] [CrossRef]
  75. Geverk’yan, S.V.; Kuripin, V.A. CO2 in cordierite structure (IR spectroscopic data). Mineral. Zhurnal 1987, 9, 49–53. [Google Scholar]
  76. 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]
  77. 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]
  78. 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]
  79. 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]
  80. Majumdar, A.S.; Mathew, G. Raman-Infrared (IR) Spectroscopy Study of Natural Cordierites from Kalahandi, Odisha. J. Geol. Soc. 2015, 86, 80–92. [Google Scholar] [CrossRef]
  81. 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]
  82. Damon, P.E.; Kulp, J.L. Excess helium and argon in beryl and other minerals. Am. Mineral. 1958, 43, 433–459. [Google Scholar]
  83. Wood, D.L.; Nassau, K. Infrared spectra of foreign molecules in beryl. J. Chem. Phys. 1967, 47, 2200–2228. [Google Scholar] [CrossRef]
  84. 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]
  85. Bul’bak, T.A.; Tomilenko, A.A.; Zatolokina, K.I.; Shaparenko, E.O. Fluid components in cordierites from the Middle Urals (Russia) and Dolní Bory (Czech Republic) pegmatites according to pyrolysis-free gas chromatography-mass spectrometry data. Bull. Tomsk Polytech. Univ. Geo Assets Eng. 2025, in press. [Google Scholar]
  86. Dobretsov, N.L. (Ed.) Early Precambrian of Southern Yakutia; Nauka: Moscow, Russia, 1986; p. 280. [Google Scholar]
  87. Rundqvist, D.V.; Mitrofanov, F.P. (Eds.) Precambrian Geology of the USSR; Academy of Sciences USSR, The Institute of Precambrian Geology and Geochronology, Nauka: Moscow, Russia, 1988; p. 441. [Google Scholar]
  88. Popov, N.V.; Smelov, A.P. Metamorphic Formations of the Aldan Shield. Russ. Geol. Geophys. 1996, 37, 148–161. [Google Scholar]
  89. Smelov, A.P.; Timofeev, V.F. Terrane Analysis and Geodynamic Model of the North Asian Craton in the Early Precambrian. Pac. Geol. 2003, 22, 42–54. [Google Scholar]
  90. Smelov, A.P.; Timofeev, V.F. The Age of the North Asian Cratonic Basement: An Overview. Gondwana Res. 2007, 12, 279–288. [Google Scholar] [CrossRef]
  91. Berezkin, V.I.; Smelov, A.P.; Zedgenizov, A.N.; Kravchenko, A.A.; Popov, N.V.; Timofeev, V.F.; Toropova, L.I. Geological Structure of the Central Part of the Aldan-Stanovoy Shield and Chemical Compositions of Early Precambrian Rocks (Southern Yakutia); Smelov, A.P., Ed.; IGABM SO RAN, INGG SO RAN: Novosibirsk, Russia, 2015; p. 459. [Google Scholar]
  92. Drogova, G.M.; Glebovitsky, V.A.; Duk, V.L.; Kitsul, V.I.; Savelyeva, T.L.; Sedova, I.S.; Semenov, A.P. High-Gradient Metamorphic Regimes in Crustal Evolution; Sokolov, Y.M., Ed.; Nauka: Leningrad, Russia, 1982; p. 229. [Google Scholar]
  93. Kitsul, V.I.; Zedgenizov, A.N. On the Binary Division of the Aldan Complex. In Stratigraphy of the Archean and Lower Proterozoic of the USSR; Nauka: Leningrad, Russia, 1979; pp. 125–130. [Google Scholar]
  94. Parfenov, L.M.; Kuzmin, M.I. (Eds.) Tectonics, Geodynamics and Metallogeny of the Territory of the Sakha Republic (Yakutia); MAIK “Nauka/Interperiodica”: Moscow, Russia, 2001; p. 572. [Google Scholar]
  95. Perchuk, L.L.; Kitsul, V.I.; Podlesskii, K.K.; Gerasimov, V.Y.; Aranovich, L.Y.; Fed’kin, V.V. Evolution of Metamorphism of the Aldan Shield. Preprint, Report. In Proceedings of the the VI All-Union Petrographic Meeting, Leningrad, Russia, 28–29 May 1981; Yakutsk Branch of the Siberian Branch of the USSR Academy of Sciences: Yakutsk, Russia, 1981; p. 63. [Google Scholar]
  96. Perchuk, L.L.; Aranovich, L.Y.; Podlesskii, K.K.; Lavrant’eva, I.V.; Gerasimov, V.Y.; Fed’kin, V.V.; Kitsul, V.I.; Karsakov, L.P.; Berdnikov, N.V. Precambrian granulites of the Aldan shield, eastern Siberia, USSR. J. Metamorph. Geol. 1985, 3, 265–310. [Google Scholar] [CrossRef]
  97. Perchuk, L.L.; Kitsul, V.I.; Aranovich, L.Y.; Podlessky, K.K.; Lavrentieva, I.V.; Gerasimov, V.Y.; Fedkin, V.V.; Berdnikov, N.V.; Karsakov, L.P. Mineralogy and Petrology of Granulites of the Aldan-Stanovoy Region; Yakutsk Branch of the USSR Academy of Sciences: Yakutsk, Russia, 1986; p. 40. [Google Scholar]
  98. Perchuk, L.L.; Kitsul, V.I.; Aranovich, L.Y. Petrology of Granulites of the Aldan Shield; Yakutsk Branch of the USSR Academy of Sciences: Yakutsk, Russia, 1987; p. 80. [Google Scholar]
  99. Tomilenko, A.A.; Chupin, V.P. Thermobarogeochemistry of Metamorphic Complexes; Nauka: Novosibirsk, Russia, 1983; p. 201. [Google Scholar]
  100. Tomilenko, A.A. Fluid Regime of Mineral Formation in the Continental Lithosphere at High–Moderate Pressures Based on the Study of Fluid and Melt Inclusions in Minerals. Ph.D. Thesis, Institute of Geology and Mineralogy SB RAS, Novosibirsk, Russia, 2006; p. 31. [Google Scholar]
  101. Kosygin, Y.A.; Luchitsky, I.V. Structures of the Siberian Platform Margins. In Tectonics of Siberia; SB USSR Academy of Sciences: Novosibirsk, Russia, 1963; Volume 2, pp. 9–12. [Google Scholar]
  102. Kirichenko, G.I. Tectonics of the Yenisei Ridge. In Tectonics of Siberia; SB USSR Academy of Sciences: Novosibirsk, Russia, 1963; Volume 2, pp. 65–82. [Google Scholar]
  103. Kuznetsov, Y.A. Petrology of the Precambrian of the South Yenisei Ridge (Materials on the Geology of Western Siberia, Issue 15(57)); West Siberian Geological Survey: Tomsk, Russia, 1941; p. 250. [Google Scholar]
  104. Kuznetsov, Y.A. Selected Works, Volume 1; Nauka: Novosibirsk, Russia, 1988; p. 217. [Google Scholar]
  105. Kovrigina, E.K. Experience in Formation Analysis of Metamorphic Sequences Using the Example of the Angara-Kan Part of the Yenisei Ridge. In Proceedings of Annual and Anniversary Sessions of the Scientific Council of VSEGEI; VSEGEI: Leningrad, Russia, 1971; pp. 116–127. [Google Scholar]
  106. Nozhkin, A.D. Early Precambrian Gneiss Complexes of the Yenisei Ridge and Their Geochemical Features. Geol. Geophys. 1983, 9, 3–10. [Google Scholar]
  107. Nozhkin, A.D.; Turkina, O.M. Geochemistry of Granulites from the Kan and Sharyzhalgay Complexes; OIGGM SO RAN: Novosibirsk, Russia, 1993; p. 223. [Google Scholar]
  108. IGG SB USSR Academy of Sciences. Precambrian Crystalline Complexes of the Yenisei Ridge: Guidebook for the Yenisei Excursion; IGG SB USSR Academy of Sciences: Novosibirsk, Russia, 1986; p. 117. [Google Scholar]
  109. Lepezin, G.G.; Nozhkin, A.D.; Gerya, T.V. Thermodynamic Parameters of Metamorphism of the Kan Series (Yenisei Ridge). Russ. Geol. Geophys. 1986, 9, 11–19. [Google Scholar]
  110. Perchuk, L.L.; Gerya, T.V.; Nozhkin, A.D. Petrology and retrograde P-T path in granulites of the Kan formation, Yenisei range, Eastern Siberia. J. Metamorphic Geol. 1989, 7, 599–617. [Google Scholar] [CrossRef]
  111. Popov, N.V. Tectonic Model of the Early Precambrian Evolution of the South Yenisei Ridge. Russ. Geol. Geophys. 2001, 42, 1028–1041. [Google Scholar]
  112. Nozhkin, A.D.; Turkina, O.M.; Likhanov, I.I.; Dmitrieva, N.V. Late Paleoproterozoic Volcanic Associations in the Southwest of the Siberian Craton (Angara-Kan block). Russ. Geol. Geophys. 2016, 57, 312–332. [Google Scholar] [CrossRef]
  113. Nozhkin, A.D.; Turkina, O.M.; Likhanov, I.I.; Savko, K.A. Paleoproterozoic Metavolcanosedimentary Sequences of the Yenisei Metamorphic Complex, Southwestern Siberian Craton (Angara–Kan Block): Subdivision, Composition, And U–Pb Zircon Age. Russ. Geol. Geophys. 2019, 60, 1384–1406. [Google Scholar] [CrossRef]
  114. 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]
  115. Khomenko, V.M.; Langer, K. Carbon oxides in cordierite channels: Determination of CO2 isotopic species and CO by single crystal IR spectroscopy. Am. Mineral. 2005, 90, 1913–1917. [Google Scholar] [CrossRef]
  116. 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]
  117. 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]
  118. Cesare, B.; Maineri, C.; Baron Toaldo, A.; Pedron, D.; Acosta Vigil, 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]
  119. Mirwald, P.W.; Jochum, C.; Maresch, W.V. Rate Studies on Hydration and Dehydration of Synthetic Mg-Cordierite. Mater. Sci. Forum 1986, 7, 113–122. [Google Scholar] [CrossRef]
  120. 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]
Figure 1. Tectonic zoning scheme of the central Aldan Shield (modified after [92]). Blocks (I–XII): I—Nimnyr; II—Melemken; III—Seym; IV—Kyurikan; V—Tyrkanda; VI—Sunnagin; VII—Nizhnegonam; VIII—Zverev; IX—Kabaktin; X—Sutam; XI—Nuyam; XII—Alvanar. Iyengra complex (1–6): 1—quartzite–gneiss, 2—gneiss, 3—enderbite–gneiss (a), enderbite–granulite (b), 4—enderbite, 5—granulite–enderbite, 6—granulite–schist strata. Dzheltula complex (7–9): 7—carbonate–gneiss, 8—carbonate–schist–gneiss, 9—carbonate–schist strata. Facies and subfacies (10–16): 10–12—biotite-cordierite–garnet–quartz–two-feldspar subfacies of granulite facies (10—low, 11—moderate, 12—high pressure); 13, 14—biotite–sillimanite–garnet–quartz–two-feldspar subfacies of granulite facies (13—low and moderate, 14—high pressure); 15—hypersthene–sillimanite–garnet–quartz–two-feldspar subfacies of granulite facies; 16—almandine amphibolite facies. 17—Stanovoy marginal suture (South Aldan fault), 18—major faults: (A) Amgin, (T) Timpton, (IN) Idzhek–Nuyam; 19—block-boundary faults: (1) Tyrkanda, (2) Gonam, (3) Sunnangin, (4) Anamzhak; 20—numbers and locations of the studied samples.
Figure 1. Tectonic zoning scheme of the central Aldan Shield (modified after [92]). Blocks (I–XII): I—Nimnyr; II—Melemken; III—Seym; IV—Kyurikan; V—Tyrkanda; VI—Sunnagin; VII—Nizhnegonam; VIII—Zverev; IX—Kabaktin; X—Sutam; XI—Nuyam; XII—Alvanar. Iyengra complex (1–6): 1—quartzite–gneiss, 2—gneiss, 3—enderbite–gneiss (a), enderbite–granulite (b), 4—enderbite, 5—granulite–enderbite, 6—granulite–schist strata. Dzheltula complex (7–9): 7—carbonate–gneiss, 8—carbonate–schist–gneiss, 9—carbonate–schist strata. Facies and subfacies (10–16): 10–12—biotite-cordierite–garnet–quartz–two-feldspar subfacies of granulite facies (10—low, 11—moderate, 12—high pressure); 13, 14—biotite–sillimanite–garnet–quartz–two-feldspar subfacies of granulite facies (13—low and moderate, 14—high pressure); 15—hypersthene–sillimanite–garnet–quartz–two-feldspar subfacies of granulite facies; 16—almandine amphibolite facies. 17—Stanovoy marginal suture (South Aldan fault), 18—major faults: (A) Amgin, (T) Timpton, (IN) Idzhek–Nuyam; 19—block-boundary faults: (1) Tyrkanda, (2) Gonam, (3) Sunnangin, (4) Anamzhak; 20—numbers and locations of the studied samples.
Minerals 15 00890 g001
Figure 2. Geological map of the South Yenisei Ridge showing metamorphic and magmatic formations (modified after [111]). 1, 2—Overlying deposits of Phanerozoic cover (1) and Late Proterozoic cover (2) age; 3—Porozhin complex: syenites, syenite porphyries, and alkaline granites (V?); 4—Yukseev complex: island-arc ophiolite formations (R); 5—Allochthonous granitoid complex: (a) granites, (b) diorites and granodiorites (PR2?); 6—complex of rheomorphic granites (PR2?); 7—Tarak gneiss–granite complex (PR1); 8—migmatite–gneiss complex (PR1); 9—Kimbir complex of layered gabbro-norites (PR1?); 10—Yenisei amphibolite–gneiss complex (PR1): (a) biotite–muscovite gneisses with interlayers of carbonate rocks and amphibolites, (b) cordierite gneisses; 11—Atamanov granulite–gneiss complex (AR?); 12—charnockite–gneiss complex (AR?); 13—Kuzeev granulite–metabasite–gneiss complex (AR?); 14—locations of critical parageneses: (a) hypersthene–sillimanite–quartz, (b) corundum–quartz, (c) eclogite-like rocks, (d) andalusite-bearing gneisses; 15—tectonic structures: (a) fault zones with blastomylonites and cataclasites, (b) thrusts, (c) major faults.
Figure 2. Geological map of the South Yenisei Ridge showing metamorphic and magmatic formations (modified after [111]). 1, 2—Overlying deposits of Phanerozoic cover (1) and Late Proterozoic cover (2) age; 3—Porozhin complex: syenites, syenite porphyries, and alkaline granites (V?); 4—Yukseev complex: island-arc ophiolite formations (R); 5—Allochthonous granitoid complex: (a) granites, (b) diorites and granodiorites (PR2?); 6—complex of rheomorphic granites (PR2?); 7—Tarak gneiss–granite complex (PR1); 8—migmatite–gneiss complex (PR1); 9—Kimbir complex of layered gabbro-norites (PR1?); 10—Yenisei amphibolite–gneiss complex (PR1): (a) biotite–muscovite gneisses with interlayers of carbonate rocks and amphibolites, (b) cordierite gneisses; 11—Atamanov granulite–gneiss complex (AR?); 12—charnockite–gneiss complex (AR?); 13—Kuzeev granulite–metabasite–gneiss complex (AR?); 14—locations of critical parageneses: (a) hypersthene–sillimanite–quartz, (b) corundum–quartz, (c) eclogite-like rocks, (d) andalusite-bearing gneisses; 15—tectonic structures: (a) fault zones with blastomylonites and cataclasites, (b) thrusts, (c) major faults.
Minerals 15 00890 g002
Figure 3. Raman spectra: (a) cordierite (host mineral; sample 99–77) from rocks of the granulite facies of the Sutam block, Aldan shield (Raman lines 125, 158, 238, 260, 294, 330, 427, 455, 555, 577, 621, 670, 712, 723, 752, 901, 926, 974, 1010, 1113, and 1184 cm−1), (b) single water molecules (Raman lines 3578 and 3600 cm−1), (c) carbon dioxide (Raman lines 1268 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm−1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2933 cm−1, νsymCH3 2912 cm−1, νsymCH2 2873 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Aldan Shield.
Figure 3. Raman spectra: (a) cordierite (host mineral; sample 99–77) from rocks of the granulite facies of the Sutam block, Aldan shield (Raman lines 125, 158, 238, 260, 294, 330, 427, 455, 555, 577, 621, 670, 712, 723, 752, 901, 926, 974, 1010, 1113, and 1184 cm−1), (b) single water molecules (Raman lines 3578 and 3600 cm−1), (c) carbon dioxide (Raman lines 1268 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm−1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2933 cm−1, νsymCH3 2912 cm−1, νsymCH2 2873 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Aldan Shield.
Minerals 15 00890 g003aMinerals 15 00890 g003b
Figure 4. IR spectra of (a) cordierite (sample 99–77) from rocks of the granulite facies, Sutam block, Aldan shield; (b) cordierite (sample 78–80) of the granulite facies, Nimnyr block, Aldan Shield, cordierites in the range of 370–1800 cm−1 and 1800–3800 cm−1 (insert).
Figure 4. IR spectra of (a) cordierite (sample 99–77) from rocks of the granulite facies, Sutam block, Aldan shield; (b) cordierite (sample 78–80) of the granulite facies, Nimnyr block, Aldan Shield, cordierites in the range of 370–1800 cm−1 and 1800–3800 cm−1 (insert).
Minerals 15 00890 g004
Figure 5. IR spectra of cordierites: (a) the wide absorption band is due to OH absorption of the absorbed water molecules hydrogen-bonded to the surface, (b) the stretching vibration of CO2 molecules, (c) the bands typical of the antisymmetric and symmetric stretching modes of CH3- and CH2- groups of aliphatic hydrocarbons, CnH2n+2asCH3 2958 cm−1, νasCH2 2926cm−1, νsymCH3 2872 cm−1, νsymCH2 2854 cm−1); 1, 2—Nimnyr block (samples 78–80 and 77–79); 3—Sutam block (sample 99–77).
Figure 5. IR spectra of cordierites: (a) the wide absorption band is due to OH absorption of the absorbed water molecules hydrogen-bonded to the surface, (b) the stretching vibration of CO2 molecules, (c) the bands typical of the antisymmetric and symmetric stretching modes of CH3- and CH2- groups of aliphatic hydrocarbons, CnH2n+2asCH3 2958 cm−1, νasCH2 2926cm−1, νsymCH3 2872 cm−1, νsymCH2 2854 cm−1); 1, 2—Nimnyr block (samples 78–80 and 77–79); 3—Sutam block (sample 99–77).
Minerals 15 00890 g005
Figure 6. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 173–77) from granulite-facies rocks of the Sutam block, Aldan Shield. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—butane (C4H10); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—1-butanol (C4H10O); 8—acetamide (C2H5NO); 9—2-oxopropionamide (C3H5NO2); 10—hexanal (C6H12O); 11—n-butanoic acid (C4H8O2); 12—n-octane (C8H18); 13—n-pentanoic acid (C5H10O2); 14—5-methylfurfural (C6H6O2); 15—n-nonane (C9H20); 16—n-hexanoic acid (C6H12O2); 17—octanal (C8H16O); 18—n-decane (C10H22); 19—n-heptanoic acid (C7H14O2); 20—nonanal (C9H18O); 21—n-octanoic acid (C8H16O2); 22—decanal (C10H20O); 23—n-nonanoic acid (C9H18O2); 24—n-tridecane (C13H28); 25—n-decanoic acid (C10H20O2); 26—dodecanal (C12H24O); 27—1-pentadecene (C15H30); 28—n-pentadecane (C15H32); 29—n-dodecanoic acid (C12H24O2); 30—2-tetradecanone (C14H28O); 31—1,3-benzenedicarboxylic acid (C8H6O4); 32—2-pentadecanone (C15H30O); 33—n-heptadecane (C17H36).
Figure 6. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 173–77) from granulite-facies rocks of the Sutam block, Aldan Shield. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—butane (C4H10); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—1-butanol (C4H10O); 8—acetamide (C2H5NO); 9—2-oxopropionamide (C3H5NO2); 10—hexanal (C6H12O); 11—n-butanoic acid (C4H8O2); 12—n-octane (C8H18); 13—n-pentanoic acid (C5H10O2); 14—5-methylfurfural (C6H6O2); 15—n-nonane (C9H20); 16—n-hexanoic acid (C6H12O2); 17—octanal (C8H16O); 18—n-decane (C10H22); 19—n-heptanoic acid (C7H14O2); 20—nonanal (C9H18O); 21—n-octanoic acid (C8H16O2); 22—decanal (C10H20O); 23—n-nonanoic acid (C9H18O2); 24—n-tridecane (C13H28); 25—n-decanoic acid (C10H20O2); 26—dodecanal (C12H24O); 27—1-pentadecene (C15H30); 28—n-pentadecane (C15H32); 29—n-dodecanoic acid (C12H24O2); 30—2-tetradecanone (C14H28O); 31—1,3-benzenedicarboxylic acid (C8H6O4); 32—2-pentadecanone (C15H30O); 33—n-heptadecane (C17H36).
Minerals 15 00890 g006
Figure 7. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 77–79) of the Nimnyr block, Aldan Shield. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—isocyanic acid (CHNO); 4—2-propanone (C3H6O); 5—2-butanone (C4H8O); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—n-octane (C8H18); 10—3-methylbutanoic acid (C5H10O2); 11—n-pentanoic acid (C5H10O2); 12—heptanal (C7H14O); 13—n-nonane (C9H20);14—n-hexanoic acid (C6H12O2); 15—octanal (C8H16O); 16—n-decane (C10H22), 17—n-heptanoic acid (C7H14O2); 18—decanal (C10H20O); 19—n-octanoic acid (C8H16O2); 20—decanal (C10H20O); 21—1,3-isobenzofurandione (C8H4O3); 22—n-nonanoic acid (C9H18O2); 23—n-decanoic acid (C10H20O2); 24—tridecanal (C13H26O); 25—1-pentadecene (C15H30); 26—n-dodecanoic acid (C12H24O2); 27—tetradecanal (C14H28O); 28—3-methyladamantane-1-carboxylic acid (C12H18O2); 29—2-hexadecanone (C16H32O); 30—hexadecanal (C16H32O); 31—dipropyl phthalate (C14H18O4); 32—n-tetradecanoic acid (C14H28O2).
Figure 7. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 77–79) of the Nimnyr block, Aldan Shield. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—isocyanic acid (CHNO); 4—2-propanone (C3H6O); 5—2-butanone (C4H8O); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—n-octane (C8H18); 10—3-methylbutanoic acid (C5H10O2); 11—n-pentanoic acid (C5H10O2); 12—heptanal (C7H14O); 13—n-nonane (C9H20);14—n-hexanoic acid (C6H12O2); 15—octanal (C8H16O); 16—n-decane (C10H22), 17—n-heptanoic acid (C7H14O2); 18—decanal (C10H20O); 19—n-octanoic acid (C8H16O2); 20—decanal (C10H20O); 21—1,3-isobenzofurandione (C8H4O3); 22—n-nonanoic acid (C9H18O2); 23—n-decanoic acid (C10H20O2); 24—tridecanal (C13H26O); 25—1-pentadecene (C15H30); 26—n-dodecanoic acid (C12H24O2); 27—tetradecanal (C14H28O); 28—3-methyladamantane-1-carboxylic acid (C12H18O2); 29—2-hexadecanone (C16H32O); 30—hexadecanal (C16H32O); 31—dipropyl phthalate (C14H18O4); 32—n-tetradecanoic acid (C14H28O2).
Minerals 15 00890 g007
Figure 8. Raman spectra of cordierite (sample 16–21) from the granulite-facies rocks of the Kan series, Yenisei Ridge: (a) cordierite (host mineral) (Raman lines 120, 157, 238, 260, 294, 330, 330, 365, 425, 455, 555, 577, 620, 670, 712, 723, 901, 922, 973, 1010, 1111, and 1183 cm−1), (b) a single water molecule (Raman line 3600 cm−1), (c) carbon dioxide (Raman lines 1268 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm−1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2932 cm−1, νsymCH3 2913 cm−1, νsymCH2 2872 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Yenisei Ridge.
Figure 8. Raman spectra of cordierite (sample 16–21) from the granulite-facies rocks of the Kan series, Yenisei Ridge: (a) cordierite (host mineral) (Raman lines 120, 157, 238, 260, 294, 330, 330, 365, 425, 455, 555, 577, 620, 670, 712, 723, 901, 922, 973, 1010, 1111, and 1183 cm−1), (b) a single water molecule (Raman line 3600 cm−1), (c) carbon dioxide (Raman lines 1268 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm−1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2932 cm−1, νsymCH3 2913 cm−1, νsymCH2 2872 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Yenisei Ridge.
Minerals 15 00890 g008
Figure 9. IR spectra of (a) cordierite (sample 16–21) from the granulite-facies rocks, Kan series, Yenisei Ridge; (b) cordierite (sample 02–06) from the amphibolite-facies rocks, Yenisei series, Yenisei Ridge in the range of 370–1800 cm−1 and 1800–3800 cm−1 (insert).
Figure 9. IR spectra of (a) cordierite (sample 16–21) from the granulite-facies rocks, Kan series, Yenisei Ridge; (b) cordierite (sample 02–06) from the amphibolite-facies rocks, Yenisei series, Yenisei Ridge in the range of 370–1800 cm−1 and 1800–3800 cm−1 (insert).
Minerals 15 00890 g009
Figure 10. IR spectra of cordierites showing (a) that the wide absorption band is due to OH absorption of the absorbed water molecules hydrogen-bonded to the surface, (b) the stretching vibration of CO2 molecules, (c) the bands typical of the antisymmetric and symmetric stretching modes of CH3- and CH2- groups of aliphatic hydrocarbons, CnH2n+2asCH3 2958 cm−1, νasCH2 2926cm−1, νsymCH3 2872 cm−1, νsymCH2 2854 cm−1); 1—cordierite from granulite-facies rocks of the Kan series (sample 16–21); 2—cordierite from amphibolite-facies rocks of the Yenisei Series (sample 02–06).
Figure 10. IR spectra of cordierites showing (a) that the wide absorption band is due to OH absorption of the absorbed water molecules hydrogen-bonded to the surface, (b) the stretching vibration of CO2 molecules, (c) the bands typical of the antisymmetric and symmetric stretching modes of CH3- and CH2- groups of aliphatic hydrocarbons, CnH2n+2asCH3 2958 cm−1, νasCH2 2926cm−1, νsymCH3 2872 cm−1, νsymCH2 2854 cm−1); 1—cordierite from granulite-facies rocks of the Kan series (sample 16–21); 2—cordierite from amphibolite-facies rocks of the Yenisei Series (sample 02–06).
Minerals 15 00890 g010
Figure 11. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 16–21) from granulite-facies rocks of the Kan series, Yenisei Ridge. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—acetaldehyde (C2H4O); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—octane (C8H18); 10—n-pentanoic acid (C5H10O2); 11—n-nonane (C9H20); 12—n-hexanoic acid (C6H12O2); 13—n-heptanoic acid (C7H14O2); 14—nonanal (C9H18O); 15—n-octanoic acid (C8H16O2); 16—decanal (C10H20O); 17—n-nonanoic acid (C9H18O2); 18—1,3-isobenzofurandione (C8H4O3); 19—n-decanoic acid (C10H20O2); 20—n-pentadecane (C15H32); 21—n-dodecanoic acid (C12H24O2); 22—tetradecanal (C14H28O); 23—2-pentadecanone (C15H30O); 24—t-butyl hydrogen phthalate (C12H14O4); 25—pentadecanal (C15H30O); 26—2-amino-6-methoxybenzoic acid (C8H9NO3); 27—n-tetradecanoic acid (C14H28O2).
Figure 11. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 16–21) from granulite-facies rocks of the Kan series, Yenisei Ridge. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—acetaldehyde (C2H4O); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—octane (C8H18); 10—n-pentanoic acid (C5H10O2); 11—n-nonane (C9H20); 12—n-hexanoic acid (C6H12O2); 13—n-heptanoic acid (C7H14O2); 14—nonanal (C9H18O); 15—n-octanoic acid (C8H16O2); 16—decanal (C10H20O); 17—n-nonanoic acid (C9H18O2); 18—1,3-isobenzofurandione (C8H4O3); 19—n-decanoic acid (C10H20O2); 20—n-pentadecane (C15H32); 21—n-dodecanoic acid (C12H24O2); 22—tetradecanal (C14H28O); 23—2-pentadecanone (C15H30O); 24—t-butyl hydrogen phthalate (C12H14O4); 25—pentadecanal (C15H30O); 26—2-amino-6-methoxybenzoic acid (C8H9NO3); 27—n-tetradecanoic acid (C14H28O2).
Minerals 15 00890 g011
Figure 12. Raman spectra of cordierite (sample PN-1) from the amphibolite-facies rocks of the Yenisei series of the Yenisei Ridge: (a) cordierite (host mineral) (Raman lines 125, 157, 238, 260, 294, 304, 329, 367, 469, 488, 555, 577, 619, 668, 712, 735, 901, 924, 974, 1009, and 1183 cm−1), (b) single water molecules (Raman lines 3580 and 3600 cm−1), (c) carbon dioxide (Raman lines 1269 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm-1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2932 cm−1, νsymCH3 2913 cm−1, νsymCH2 2872 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Yenisei Ridge.
Figure 12. Raman spectra of cordierite (sample PN-1) from the amphibolite-facies rocks of the Yenisei series of the Yenisei Ridge: (a) cordierite (host mineral) (Raman lines 125, 157, 238, 260, 294, 304, 329, 367, 469, 488, 555, 577, 619, 668, 712, 735, 901, 924, 974, 1009, and 1183 cm−1), (b) single water molecules (Raman lines 3580 and 3600 cm−1), (c) carbon dioxide (Raman lines 1269 and 1381 cm−1), (d) nitrogen (molecular nitrogen—Raman line 2326 cm−1; atmospheric nitrogen—Raman line 2330 cm-1), (e) CH2- and CH3- groups of aliphatic hydrocarbons (νasCH3 2969 cm−1, νasCH2 2932 cm−1, νsymCH3 2913 cm−1, νsymCH2 2872 cm−1) and CH- groups of benzene ring (νC-H at C=C 3069 cm−1) isolated in crystal cavities of cordierites from the Yenisei Ridge.
Minerals 15 00890 g012aMinerals 15 00890 g012b
Figure 13. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 02–06) from the amphibolite-facies rocks of the Yenisei series, Yenisei Ridge. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—acetaldehyde (C2H4O); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—n-octane (C8H18); 10—n-pentanoic acid (C5H10O2); 11—n-nonane (C9H20); 12—n-hexanoic acid (C6H12O2); 13—n-heptanoic acid (C7H14O2); 14—nonanal (C9H18O); 15—n-octanoic acid (C8H16O2); 16—decanal (C10H20O); 17—n-nonanoic acid (C9H18O2); 18—1,1,4,6-tetramethylindane (C13H18); 19—n-decanoic acid (C10H20O2); 20—3-fluoro-4-(methylamino)benzonitrile (C8H7FN2); 21—n-pentadecane (C15H32); 22—γ-undecalatone (C11H20O2); 23—n-dodecanoic acid (C12H24O2); 24—2-pentadecanone (C15H30O); 25—n-heptadecane (C17H36); 26—N-(2-phenylethyl)-N-ethyl-2-methylpropanamide (C14H21NO); 27—n-tetradecanoic acid (C14H28O2).
Figure 13. Results of GC-MS analysis of volatile components extracted from structural channels (cavities) and non-structural positions in cordierite (sample 02–06) from the amphibolite-facies rocks of the Yenisei series, Yenisei Ridge. (a) Chromatogram shows the total ion current (TIC); (b–e) reconstructed ion chromatograms for ion current: (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—acetaldehyde (C2H4O); 4—2-propanone (C3H6O); 5—2,3-butanedione (C4H6O2); 6—acetic acid (C2H4O2); 7—n-heptane (C7H16); 8—n-butanoic acid (C4H8O2); 9—n-octane (C8H18); 10—n-pentanoic acid (C5H10O2); 11—n-nonane (C9H20); 12—n-hexanoic acid (C6H12O2); 13—n-heptanoic acid (C7H14O2); 14—nonanal (C9H18O); 15—n-octanoic acid (C8H16O2); 16—decanal (C10H20O); 17—n-nonanoic acid (C9H18O2); 18—1,1,4,6-tetramethylindane (C13H18); 19—n-decanoic acid (C10H20O2); 20—3-fluoro-4-(methylamino)benzonitrile (C8H7FN2); 21—n-pentadecane (C15H32); 22—γ-undecalatone (C11H20O2); 23—n-dodecanoic acid (C12H24O2); 24—2-pentadecanone (C15H30O); 25—n-heptadecane (C17H36); 26—N-(2-phenylethyl)-N-ethyl-2-methylpropanamide (C14H21NO); 27—n-tetradecanoic acid (C14H28O2).
Minerals 15 00890 g013
Figure 14. Relative content of hydrocarbons (HCs), CO2, H2O, and nitrogenated and sulfonated compounds released during mechanical destruction of cordierites from rocks of (I) granulite facies of the Sutam block, Aldan Shield, (II) granulite facies of the Nimnyr block, Aldan Shield, (III) amphibolite facies of the Yenisei series, Yenisei Ridge, (IV) epidote–amphibolite facies of the Muzkol complex, Tajikistan [7].
Figure 14. Relative content of hydrocarbons (HCs), CO2, H2O, and nitrogenated and sulfonated compounds released during mechanical destruction of cordierites from rocks of (I) granulite facies of the Sutam block, Aldan Shield, (II) granulite facies of the Nimnyr block, Aldan Shield, (III) amphibolite facies of the Yenisei series, Yenisei Ridge, (IV) epidote–amphibolite facies of the Muzkol complex, Tajikistan [7].
Minerals 15 00890 g014aMinerals 15 00890 g014b
Figure 15. Volatile composition of cordierites from rocks of granulite- (I,II) and amphibolite- (III) facies metamorphism of the Aldan Shield and Yenisei Ridge, and epidote–amphibolite- (IV) facies metamorphism of the Muzkol complex, Tajikistan [7], on the COH diagram (according to gas chromatography–mass spectrometry analysis).
Figure 15. Volatile composition of cordierites from rocks of granulite- (I,II) and amphibolite- (III) facies metamorphism of the Aldan Shield and Yenisei Ridge, and epidote–amphibolite- (IV) facies metamorphism of the Muzkol complex, Tajikistan [7], on the COH diagram (according to gas chromatography–mass spectrometry analysis).
Minerals 15 00890 g015
Table 1. Chemical analysis of cordierites of the Aldan Shield (wt. %).
Table 1. Chemical analysis of cordierites of the Aldan Shield (wt. %).
ComponentSutam BlockNimnyr Block
99–77173–77174–77176–7777–7939–8042–8078–80
SiO248.6449.3249.4948.9748.8648.4448.4048.43
TiO20.010.010.010.010.010.010.010.00
Al2O332.9333.2333.3433.0233.0432.8732.7132.75
FeO5.233.713.384.166.559.936.648.09
MnO0.020.050.030.040.200.080.170.03
MgO10.6911.5111.7211.269.778.109.929.09
CaO0.010.010.000.010.010.020.010.02
Na2O0.040.030.050.020.120.080.090.05
K2O0.010.000.010.010.020.000.020.02
Cr2O30.010.000.000.000.010.000.000.01
Σ97.5897.8698.0397.5098.5899.5397.9798.48
Formula for 18 O atoms
Si4.984.995.004.994.984.974.974.98
Ti0.000.000.000.000.000.000.000.00
Al3.973.973.973.973.973.973.963.97
Fe(II)0.450.310.290.350.560.850.570.69
Mn0.000.000.000.000.020.010.010.00
Mg1.631.741.761.711.491.241.521.39
Ca0.000.000.000.000.000.000.000.00
Na0.010.010.010.000.020.020.020.01
K0.000.000.000.000.000.000.000.00
Cr0.000.000.000.000.000.000.000.00
Σ11.0411.0211.0311.0311.0411.0511.0611.05
XMg0.780.850.860.830.720.590.720.67
Table 2. Parameters of IR spectra of cordierites from rocks of the granulite facies of the Aldan Shield and assignment of bands.
Table 2. Parameters of IR spectra of cordierites from rocks of the granulite facies of the Aldan Shield and assignment of bands.
Band SystemSutam BlockNimnyr BlockAttribution
A385.7385.7385.7Lattice, Valent
Mg—O(MgO6) + O—Si—O
Deformational (SiO4)
422421.4420.5
440.7438.8439.7
488487486
B564564564Valent Al—O (AlO6) + O—Al—O Deformational (AlO4)
579.6579.6579.6
C618616617Deformational (Si,Al)—O—(Si, Al) + (Si,Al) —OH
675674674
704704704
D750 sh748 sh750 shValent Al—O (AlO6)
770.5768.6768.6
E909909909Valent Al—O (AlO6)
Si—O (SiO6)
961962.4961.4
102710271026
F114311421142Valent
Si—O (SiO6)
1176.51175.51175.5
163616361636ν2 H2O-II
234823482348CO2
Note: sh—shoulder.
Table 3. Composition of volatile components isolated during mechanical destruction of cordierites from granulites of the Sutam block of the Aldan Shield, according to GC-MS analysis (rel. %).
Table 3. Composition of volatile components isolated during mechanical destruction of cordierites from granulites of the Sutam block of the Aldan Shield, according to GC-MS analysis (rel. %).
NameMWSutam Block
99–77173–77174–77176–77
Aliphatic hydrocarbons 0.190.781.240.43
Paraffins (CH4-C17H36)16–2400.090.220.440.20
Olefins (C2H2-C17H34)26–2380.100.560.800.23
Cyclic hydrocarbons 0.100.240.300.22
Cycloalkanes (naphthenes) and cycloalkenes (C3H6-C10H16)42–1360.010.010.080.01
Arenes (C6H6-C16H26)78–2180.090.210.190.18
PAH (C10H8-C13H18)128–1740.010.010.030.03
Oxygenatedhydrocarbons 0.691.483.061.45
Alcohols (CH4O-C12H22O2)32–1980.050.080.340.11
Esters and ethers (C4H4O2-C16H19FO2)84–2620.090.130.160.10
Aldehydes (CH2O-C16H32O)30–2400.140.351.370.33
Ketones (C3H6O-C16H32O)58–2400.140.240.440.31
Carboxylic acids (C2H4O2-C14H28O2)60–2280.270.680.750.60
Heterocyclic compounds 0.010.020.030.03
Dioxanes (C4H8O2-C5H10O2)88–102<0.01<0.01<0.01<0.01
Furans (C4H4O-C15H26O)68–2220.010.020.030.03
Nitrogenated compounds (N2-C13H16F3NO)28–2590.140.350.880.26
Sulfonated compounds (H2S-C13H22S)34–2100.040.090.140.08
CO24497.5592.9788.5986.71
H2O181.284.085.7510.83
Total number of components 226237261254
CO2/(H2O + CO2) 0.990.960.940.89
H/(H + O) 0.050.140.210.24
Note: PAHs—polycyclic aromatic hydrocarbons; MW—nominal mass.
Table 4. Composition of volatile components isolated during mechanical destruction of cordierites from granulites of the Nimnyr block of the Aldan Shield, according to GC-MS analysis (rel. %).
Table 4. Composition of volatile components isolated during mechanical destruction of cordierites from granulites of the Nimnyr block of the Aldan Shield, according to GC-MS analysis (rel. %).
NameMWNimnyr Block
39–8042–8078–8077–79
Aliphatic hydrocarbons 4.392.241.051.19
Paraffins (CH4-C17H36)16–2403.330.720.180.58
Olefins (C2H2-C17H34)26–2381.061.520.870.61
Cyclic hydrocarbons 0.590.660.560.81
Cycloalkanes (naphthenes) and cycloalkenes (C3H6-C10H16)42–1360.130.010.070.06
Arenes (C6H6-C16H26)78–2180.350.550.450.70
PAH (C10H8-C13H18)128–1740.110.100.050.05
Oxygenated hydrocarbons 3.227.912.355.86
Alcohols (CH4O-C12H22O2)32–1980.712.470.300.73
Esters and ethers (C4H4O2-C16H19FO2)84–2620.880.530.680.83
Aldehydes (CH2O-C16H32O)30–2400.772.370.681.73
Ketones (C3H6O-C16H32O)58–2400.320.740.331.03
Carboxylic acids (C2H4O2-C14H28O2)60–2280.551.800.371.54
Heterocyclic compounds 0.050.110.040.08
Dioxanes (C4H8O2-C5H10O2)88–1020.010.01<0.01<0.01
Furans (C4H4O-C15H26O)68–2220.040.100.030.07
Nitrogenated compounds (N2-C13H16F3NO)28–2591.321.761.031.13
Sulfonated compounds (H2S-C13H22S)34–2100.130.320.130.22
CO24472.4069.6380.0466.86
H2O1817.8917.3714.8023.85
Total number of components 258246261245
CO2/(H2O + CO2) 0.800.800.840.74
H/(H + O) 0.410.430.320.44
Note: PAHs—polycyclic aromatic hydrocarbons; MW—nominal mass.
Table 5. Chemical analysis of cordierites from the rocks of the granulite facies of the Yenisei Ridge (wt. %).
Table 5. Chemical analysis of cordierites from the rocks of the granulite facies of the Yenisei Ridge (wt. %).
ComponentYenisei Ridge
Kan Series, Granulite Facies
16–21 a16–21 b16–21 c16–21 d16–21 e16–21 f16–21 g
SiO248.6848.4948.5748.5848.6248.3448.50
TiO20.000.000.000.000.000.000.02
Al2O332.6232.7632.9932.9632.7532.9132.77
FeO6.766.626.716.727.157.027.24
MnO0.180.180.190.190.160.180.18
MgO9.659.779.809.719.729.619.53
CaO0.010.010.010.010.010.000.01
Na2O0.040.060.100.050.070.040.04
K2O0.000.020.020.000.010.010.00
Cr2O30.010.000.010.010.010.000.00
Σ97.9597.9198.4098.2298.5198.1198.30
Formula for 18 O atoms
Si5.004.984.974.984.984.974.98
Ti0.000.000.000.000.000.000.00
Al3.953.973.983.983.953.983.96
Fe(II)0.580.570.570.580.610.600.62
Mn0.020.020.020.020.010.020.02
Mg1.481.501.491.481.481.471.46
Ca0.000.000.000.000.000.000.00
Na0.010.010.020.010.010.010.01
K0.000.000.000.000.000.000.00
Cr0.000.000.000.000.000.000.00
Σ11.0311.0411.0511.0411.0611.0511.05
XMg0.710.720.720.710.700.700.70
Note: a–g—different grains.
Table 6. Parameters of IR spectra of cordierites from rocks of granulite and amphibolite facies of the Yenisei Ridge and assignment of bands.
Table 6. Parameters of IR spectra of cordierites from rocks of granulite and amphibolite facies of the Yenisei Ridge and assignment of bands.
Band SystemYenisei Ridge, Kan Series, Granulite FaciesYenisei Ridge, Yenisei Series, Amphibolite FaciesAttribution
16–21 (a)02–06 (a)
A385.7385.7Lattice, Valent
Mg—O(MgO6) + O—Si—O
Deformational (SiO4)
422421.4
440.7438.8
488487
B564564Valent Al—O (AlO6) + O—Al—O Deformational (AlO4)
579.6579.6
C618616Deformational (Si,Al)—O—(Si, Al) + (Si,Al)—OH
675674
704704
D750 sh748 shValent Al—O (AlO6)
770.5768.6
E909909Valent Al—O (AlO6)
Si—O (SiO6)
961962.4
10271027
F11431142Valent
Si—O (SiO6)
1176.51175.5
16361636ν2 H2O-II
23482358CO2
Note: sh—shoulder; letter a denotes the specific cordierite grain analyzed.
Table 7. Composition of volatile components isolated during mechanical destruction of cordierites from rocks of the granulite and amphibolite facies of the Yenisei Ridge, according to GC-MS analysis (rel. %).
Table 7. Composition of volatile components isolated during mechanical destruction of cordierites from rocks of the granulite and amphibolite facies of the Yenisei Ridge, according to GC-MS analysis (rel. %).
NameMWGranulite FaciesAmphibolite Facies
16–21PN-102–06
Aliphatic hydrocarbons 1.213.685.04
Paraffins (CH4-C18H38)16–2540.371.111.81
Olefins (C2H2-C18H36)26–2520.842.573.23
Cyclic hydrocarbons 0.811.341.28
Cycloalkanes (naphthenes) and cycloalkenes (C10H16)820.010.030.03
Arenes (C6H6-C14H22)78–1900.741.201.08
PAH (C10H8-C13H18)128–1740.060.120.18
Oxygenated hydrocarbons 7.217.3910.35
Alcohols (CH4O-C8H10O2)32–1380.770.890.89
Esters and ethers (C4H6O2-C13H8ClFO2)84–2501.480.580.91
Aldehydes (CH2O-C15H30O)30–2261.833.303.65
Ketones (C3H6O-C15H30O)58–2261.491.151.85
Carboxylic acids (C2H4O2-C14H28O2)60–2281.641.473.05
Heterocyclic compounds 0.050.190.26
Dioxanes (C4H4O-C6H12O2)88–1160.010.040.06
Furans (C4H4O-C14H24O)68–2080.050.150.20
Nitrogenated compounds
(N2-C16H24FNO)
28–2651.060.353.72
Sulfonated compounds (H2S-C12H20S)34–1960.341.090.36
CO24484.2543.5545.07
H2O185.0742.3333.79
Total number of components 257263273
CO2/(H2O + CO2) 0.940.510.57
H/(H + O) 0.260.580.59
Note: PAHs—polycyclic aromatic hydrocarbons; MW—nominal mass.
Table 8. Chemical analysis of cordierites from the amphibolite-facies rocks of the Yenisei Ridge (wt. %).
Table 8. Chemical analysis of cordierites from the amphibolite-facies rocks of the Yenisei Ridge (wt. %).
ComponentYenisei Ridge, Yenisei Series, Amphibolite Facies
PN-1 aPN-1 bPN-1 cPN-1 dPN-1 e02–06 a02–06 b02–06 c02–06 d02–06 e
SiO248.8948.7748.7348.7548.4148.6548.6548.7849.0448.81
TiO20.010.000.000.010.010.000.000.000.000.00
Al2O332.8333.0832.8932.9132.5333.1532.9033.1733.1133.22
FeO6.666.686.726.636.636.626.566.716.706.69
MnO0.290.300.320.300.280.310.330.280.300.32
MgO9.899.909.819.759.749.819.669.689.869.84
CaO0.010.000.000.000.010.000.000.000.000.01
Na2O0.110.120.130.120.100.100.140.080.100.12
K2O0.020.000.010.010.020.020.030.020.010.01
Cr2O30.010.000.000.010.000.000.000.000.000.00
Σ98.7298.8698.6198.4897.7198.6598.2798.7199.1199.02
Formula for 18 O atoms
Si4.994.974.984.984.994.964.984.974.984.96
Ti0.000.000.000.000.000.000.000.000.000.00
Al3.943.973.963.963.953.993.973.993.963.98
Fe(II)0.570.570.570.570.570.560.560.570.570.57
Mn0.030.030.030.030.020.030.030.020.030.03
Mg1.501.501.491.481.501.491.471.471.491.49
Ca0.000.000.000.000.000.000.000.000.000.00
Na0.020.020.030.020.020.020.030.010.020.02
K0.000.000.000.000.000.000.000.000.000.00
Cr0.000.000.000.000.000.000.000.000.000.00
Σ11.0511.0611.0611.0511.0511.0511.0511.0411.0511.06
XMg0.720.720.710.710.720.720.710.710.720.71
Note: a–e—different grains.
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.; Tomilenko, A.; Bul’bak, T.; Popov, N. Fluid Components in Cordierites from Granulite- and Amphibolite-Facies Rocks of the Aldan Shield and Yenisei Ridge, Russia: Evidence from Pyrolysis-Free GC-MS, Raman, and IR Spectroscopy. Minerals 2025, 15, 890. https://doi.org/10.3390/min15090890

AMA Style

Zatolokina K, Tomilenko A, Bul’bak T, Popov N. Fluid Components in Cordierites from Granulite- and Amphibolite-Facies Rocks of the Aldan Shield and Yenisei Ridge, Russia: Evidence from Pyrolysis-Free GC-MS, Raman, and IR Spectroscopy. Minerals. 2025; 15(9):890. https://doi.org/10.3390/min15090890

Chicago/Turabian Style

Zatolokina, Ksenia, Anatoly Tomilenko, Taras Bul’bak, and Nikolay Popov. 2025. "Fluid Components in Cordierites from Granulite- and Amphibolite-Facies Rocks of the Aldan Shield and Yenisei Ridge, Russia: Evidence from Pyrolysis-Free GC-MS, Raman, and IR Spectroscopy" Minerals 15, no. 9: 890. https://doi.org/10.3390/min15090890

APA Style

Zatolokina, K., Tomilenko, A., Bul’bak, T., & Popov, N. (2025). Fluid Components in Cordierites from Granulite- and Amphibolite-Facies Rocks of the Aldan Shield and Yenisei Ridge, Russia: Evidence from Pyrolysis-Free GC-MS, Raman, and IR Spectroscopy. Minerals, 15(9), 890. https://doi.org/10.3390/min15090890

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