Hostname: page-component-7c8c6479df-5xszh Total loading time: 0 Render date: 2024-03-17T10:32:33.729Z Has data issue: false hasContentIssue false

Petrovite, Na10CaCu2(SO4)8, a new fumarolic sulfate from the Great Tolbachik fissure eruption, Kamchatka Peninsula, Russia

Published online by Cambridge University Press:  30 June 2020

Stanislav K. Filatov*
Affiliation:
Institute of Earth Sciences, Saint Petersburg State University, University Emb. 7/9., 199034, Saint Petersburg, Russia
Andrey P. Shablinskii
Affiliation:
Institute of Silicate Chemistry of the Russian Academy of Sciences, Makarova Emb. 2., 199034, Saint Petersburg, Russia
Sergey V. Krivovichev
Affiliation:
Institute of Earth Sciences, Saint Petersburg State University, University Emb. 7/9., 199034, Saint Petersburg, Russia Institute of Silicate Chemistry of the Russian Academy of Sciences, Makarova Emb. 2., 199034, Saint Petersburg, Russia Nanomaterials Research Centre, Kola Science Centre of the Russian Academy of Sciences, Fersmana str. 14., 184209, Apatity, Russia
Lidiya P. Vergasova
Affiliation:
Institute of Volcanology and Seismology, Far Eastern Branch of the Russian Academy of Sciences, Piip Boulevard 9, 683006, Petropavlovsk-Kamchatsky, Russia
Svetlana V. Moskaleva
Affiliation:
Institute of Volcanology and Seismology, Far Eastern Branch of the Russian Academy of Sciences, Piip Boulevard 9, 683006, Petropavlovsk-Kamchatsky, Russia
*
*Author for correspondence: Stanislav K. Filatov, Email: filatov.stanislav@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

Petrovite, Na10CaCu2(SO4)8, is a new sulfate mineral discovered on the Second scoria cone of the Great Tolbachik fissure eruption. The mineral occurs as globular aggregates of tabular crystals up to 0.2 mm in maximal dimension, generally with gaseous inclusions. The empirical formula calculated on the basis of O = 32 is Na6(Na1.80K0.20)Σ2Na(Ca0.82Na0.06Mg0.02)Σ0.90(Cu1.84Mg0.16)Σ2(Na0.520.48)Σ1S8.12O32. The crystal-chemical formula is CuNa6−2xCax(SO4)4, which, for x ≈ 0.5, results in the idealised formula Na10CaCu2(SO4)8. The crystal structure of petrovite was determined using single-crystal X-ray diffraction data; the space group is P21/c, a = 12.6346(8), b = 9.0760(6), c = 12.7560(8) Å, β = 108.75(9)°, V = 1385.1(3) Å3, Z = 2 and R1 = 0.051. There are one Cu and six Na sites, one of which is also occupied by the essential amount of Ca. The Cu atom forms five Cu–O bonds in the range 1.980–2.180 Å and two long bonds ≈ 2.9 Å resulting in the formation of the CuO7 polyhedra, which share corners with SO4 tetrahedra to form isolated [Cu2(SO4)8]12− clusters. The clusters are surrounded by Na sites, which provide their linkage into a three-dimensional framework. The Mohs’ hardness is 4. The mineral is biaxial (+), with α = 1.498(3), βcalc = 1.500, γ = 1.516(3) and 2V = 20(10) (λ = 589 nm). The seven strongest lines of the powder X-ray diffraction pattern [d, Å (I, %) (hkl)] are: 7.21(27)(110); 6.25(38)(102); 4.47(31)(212); 3.95(21)(30$\bar{2}$); 3.85(17)(121); 3.70(36)(202); and 3.65(34)(22$\bar{1}$). The mineral is named in honour of Prof Dr Tomas Georgievich Petrov (b. 1931) for his contributions to mineralogy and crystallography and, in particular, for the development of technology for the industrial fabrication of jewellery malachite.

Type
Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2020

Introduction

The fumarole activity of the scoria cones of the Great Tolbachik fissure eruption of 1975–1976 (Fedotov and Markhinin, Reference Fedotov and Markhinin1983) and the Tolbachik fissure eruption 2012–2013 (Karpov et al., Reference Karpov, Krivovichev, Vergasova, Chernyat'eva, Anikin, Moskaleva and Filatov2013) is accompanied by intensive exhalation mineralisation, in which sulfate mineralisation plays one of the leading roles. Nowadays, more than 25 new sulfates are described on the Tolbachik volcano, Kamchatka peninsula, Russia, most of which are found on the Second scoria cone of the Great fissure Tolbachik eruption (Pekov et al., Reference Pekov, Agakhanov, Zubkova, Koshlyakova, Shchipalkina, Sandalov, Yapaskurt, Turchkova and Sidorov2020). Recently, many new sulfates have been discovered, including Na anhydrous sulfates without additional anions: ivsite, Na3H(SO4)2 (Filatov et al., Reference Filatov, Karpov, Shablinskii, Krivovichev, Vergasova and Antonov2013), bubnovaite, K2Na8Ca(SO4)6 (Gorelova et al., Reference Gorelova, Vergasova, Krivovichev, Avdontseva, Moskaleva, Karpov and Filatov2016), puninite, Na2Cu3O(SO4)3 (Siidra et al., Reference Siidra, Nazarchuk, Zaitsev, Lukina, Avdontseva, Vergasova, Vlasenko, Filatov and Karpov2017), saranchinaite, Na2Cu(SO4)2 (Siidra et al., Reference Siidra, Lukina, Nazarchuk, Depmeier, Bubnova, Agakhanov, Avdontseva, Filatov and Kovrugin2018), itelmenite, Na2CuMg2(SO4)4 (Nazarchuk et al., Reference Nazarchuk, Siidra, Agakhanov, Lukina, Avdontseva and Karpov2018), belomarinaite, KNaSO4 (Filatov et al., Reference Filatov, Shablinskii, Vergasova, Saprikina, Bubnova, Moskaleva and Belousov2019), koryakite, NaKMg2Al2(SO4)6 (Siidra et al., Reference Siidra, Nazarchuk, Zaitsev and Vlasenko2020), natroapthitalite, Na3K(SO4)2 (Shchipalkina et al., Reference Shchipalkina, Pekov, Chukanov, Zubkova, Belakovskiy, Koshlyakova, Britvin, Sidorov and Vozchikova2020), metathénardite, Na2SO4 (Pekov et al., Reference Pekov, Shchipalkina, Zubkova, Gurzhiy, Agakhanov, Belakovskiy, Chukanov, Lykova, Vigasina, Koshlyakova, Sidorov and Giester2019) and dobrovolskiite, Na4Ca(SO4)3 (Shablinskii et al., Reference Shablinskii, Filatov, Vergasova, Moskaleva, Avdontseva and Bubnova2020). Reviews on the fumarolic mineralisation of the Tolbachik scoria cones have been provided by Vergasova and Filatov (Reference Vergasova and Filatov2012, Reference Vergasova and Filatov2016).

Petrovite has been approved by the Commission on New Minerals, Nomenclature and Classification of the International Mineralogical Association (IMA2018-149b, Filatov et al., Reference Filatov, Shablinskii, Vergasova, Krivovichev and Moskaleva2020). The mineral is named in honour of Prof. Dr. Tomas Georgievich Petrov (b. 1931), a former chairman of the Crystallogenesis Laboratory of the Department of Crystallography, St. Petersburg State University. The scientific activities of Tomas Petrov were devoted to the modelling of crystal growth of minerals (Petrov et al., Reference Petrov, Treywus and Kasatkin1969). He was the first to develop the industrial technology of fabrication of jewellery malachite back in 1977 (Petrov et al., Reference Petrov, Protopopov and Shuyskiy2013). Tomas Petrov is the author of the two-parameter Alphabet for the Coding of Structural–Chemical Information and RHAT-catalogue of modal mineral compositions of magmatic rocks (Petrov et al., Reference Petrov, Andriyanets-Buyko and Moshkin2012; Petrov, Reference Petrov2014). Type material is stored in the Saint-Petersburg State University Mineralogical Museum, University Emb. 7/9, St. Petersburg 199034, Russia, under catalogue number 1/19696.

Occurrence and association

The holotype material was collected in the fumarole located on the west side of the micrograben of the Second scoria cone of the Northern Breakthrough of the Great Tolbachik fissure eruption that occurred in 1975–1976 on the Tolbachik volcano, Kamchatka, Far-Eastern Region, Russia (55°41'N, 160°14'E, 1200 m asl). The specimen was found in 2000. The temperature of the surface of the fumarole was ~200°C. The mineral is a product of exhalative fumarolic activity. The host rock for the mineral is a basalt scoria. The sample was placed into a glass tube, which was closed and waxed and kept under ambient conditions. When the glass tube was opened, no changes of sample were noted.

The graben at the place of the sample collecting was filled with the pyroclastic material (see Fig. 1). The presence of the gas stream was manifested by relatively high temperatures (up to 200°C), newly formed mineral phases and the development of an oxidation process.

Fig. 1. A view of the fissure of the micrograben on the west side of the Second Cinder cone of Great Tolbachik fissure eruption (photo taken in 1981).

The new mineral in the form of blue cryptocrystalline crusts enveloped fine pyroclastic material. Petrovite occurs in association with tiny black scales of tenorite in close intergrowths with transparent green particles of euchlorine, NaKCu3O(SO4)3 and white particles of dobrovolskiite, Na4Ca(SO4)3. The change of eruptive pyroclastic material was characterised by the wide development of oxidation processes accompanied by the formation of a fine hematite phase. The mineral is relatively stable, which allowed its detailed investigation. The crystallisation of the mineral most likely happened by direct precipitation from volcanic gases.

General appearance and physical properties

The mineral occurs as blue and green globular aggregates of tabular crystals up to 0.2 mm in maximal dimension, generally with gaseous inclusions (Figs 2 and 3). The streak is white and the lustre is vitreous. Fracture is conchoidal. No cleavage or parting was observed. The Mohs’ hardness is 4. The calculated density based on the empirical formula and single-crystal unit-cell parameters is 2.80 g/cm3.

Fig. 2. Scanning electron microscopy image of an individual grain of petrovite. Part of the petrovite sample stored in the Mineralogical Museum of St. Petersburg State University (1/19696).

Fig. 3. Blue cryptocrystalline crusts of petrovite enveloped by fine pyroclastic material. Parts of the petrovite sample stored in the Mineralogical Museum of St. Petersburg State University (1/19696).

The mineral is optically biaxial (+), with α = 1.498(3), βcalc = 1.500, γ = 1.516(3) and 2V = 20(10) (λ = 589 nm). No dispersion or pleochroism was observed. The Gladstone–Dale compatibility index (Mandarino, Reference Mandarino1981) based on empirical formula and unit-cell parameters from powder X-ray diffraction data is calculated as (1 – K P/K C) = –0.0007 (superior).

Chemical composition

The chemical composition of petrovite was studied using a TESCAN “Vega3” electron microprobe equipped with an Oxford Instruments X-max 50 silicon drift energy-dispersive spectroscopy system, operated at 20 kV and 700 pA, with a beam size of 220 nm. Analytical results are given in Table 1. The data processing was done using Aztec software and an X-MAX-80 mm2 detector (Institute of Volcanology and Seismology, Petropavlovsk-Kamchatsky, Russia). The empirical formula calculated on the basis of 32 O atoms per formula unit is Na9.38Ca0.82K0.20Cu1.84Mg0.18S8.12O32. The formula that takes into account the site assignment is Na3(Na,K)(Cu,Mg)(Na,Ca,Mg)(Na,□)(SO4)4. The idealised formula derived from the combination of chemical and structural studies (see below) is Na10Cu2Ca(SO4)8.

Table 1. Chemical composition of petrovite (wt.%).

S.D. – standard deviation

Powder X-ray diffraction

The powder X-ray diffraction data were collected using a Rigaku R-AXIS RAPID II (Gandolfi mode with CoKα) and handled using the domestic software (Britvin et al., Reference Britvin, Dolivo-Dobrovolsky and Krzhizhanovskaya2017). Petrovite is monoclinic, P21/c, a = 12.6346(8), b = 9.0760(6), c = 12.7560(8) Å, β = 110.75(9)°, V = 1385.1(3) Å3 and Z = 2.

The measured and calculated powder-diffraction data are given in Table 2. The seven strongest lines are [d, Å (I, %) (hkl)]: 7.21(27)(110); 6.25(38)(102); 4.47(31)(212); 3.95(21)(30$\bar{2}$); 3.85(17)(121); 3.70(36)(202); and 3.65(34)(22$\bar{1}$).

Table 2. Powder X-ray diffraction data (d in Å) for petrovite.

The seven strongest lines are given in bold

Single-crystal X-ray diffraction

The single-crystal X-ray diffraction data were collected using a Bruker Smart APEX II diffractometer equipped with a CCD detector using MoKα radiation. A hemisphere of three-dimensional data was collected using a frame width of 0.5° in ω, with 60 s used to acquire each frame. The data were corrected for Lorentz, polarisation and background effects using the Bruker program APEX. A semi-empirical absorption-correction based on the intensities of equivalent reflections was applied in the SADABS program.

The crystal structure of petrovite was solved by charge flipping and refined on the basis of 2470 unique observed reflections using the JANA2006 program suite (Petříček et al., Reference Petříček, Dusek and Palatinus2006). The observed isotropic displacement parameter of the Na6 site was too high (0.327 Å2), so the site was split into two mutually exclusive sites with reasonable displacement parameters. Crystallographic data and refinement parameters, atomic coordinates and isotropic displacement parameters, atomic anisotropic displacement parameters and selected interatomic distances are summarised in Tables 3–5. The crystallographic information files have been deposited with the Principal Editor of Mineralogical Magazine and are available as Supplementary material (see below).

Table 3. Crystal data, data collection information and structure refinement details for petrovite.

Table 4. Atomic coordinates and displacement parameters (Å2) for petrovite.

Table 5. Selected interatomic distances (Å) for petrovite.

Crystal structure

The crystal structure of petrovite belongs to a new structure type. It contains four symmetrically independent SO4 tetrahedra with the average <S–O> bond lengths in the range 1.45–1.47 Å, in general agreement with the average value of 1.475 Å given for sulfate minerals (Hawthorne et al., Reference Hawthorne, Krivovichev, Burns, Alpers, Jambor and Nordstrom2000). The short S–O bonds in several tetrahedra can be explained by libration effects, typical for the crystal structures with statistical or dynamical disorder.

There are six Na and one Cu sites with different occupancies. The Cu atom forms five Cu–O bonds in the range 1.980–2.180 Å and two long bonds ≈ 2.9 Å resulting in the formation of the CuO7 polyhedra with [5+2] coordination of Cu (Fig. 4a). The Cu site is predominantly occupied by Cu (89%) with the admixture of Mg (11%). This type of coordination geometry of Cu2+ cations is rather unusual, but was described previously in the crystal structures of saranchinaite, Na2Cu(SO4)2 (Siidra et al., Reference Siidra, Lukina, Nazarchuk, Depmeier, Bubnova, Agakhanov, Avdontseva, Filatov and Kovrugin2018; Kovrugin et al., Reference Kovrugin, Nekrasova, Siidra, Mentre, Masquelier, Stefanovich and Colmont2019), the high pressure phase (II) CuGeO3 (Yoshiasa et al., Reference Yoshiasa, Yagyu, Ito, Yamanaka and Nagai2000) and Cu3(Hbtc)(btc)(bpy)2 (Nadeem et al., Reference Nadeem, Bhadbhade, Bircher and Stride2010). After saranchinaite, petrovite is the second mineral with heptacoordinated Cu2+.

Fig. 4. The crystal structure of petrovite: (a) CuO7 polyhedron; (b) the [Cu2(SO4)8]12– cluster; (c) arrangement of the Cu2(SO4)8 clusters in the structure; and (d) three-dimensional framework of the crystal structure.

The CuO7 polyhedra share corners with SO4 tetrahedra to form isolated [Cu2(SO4)8]12− anionic clusters shown in Fig. 4b, which are the fundamental building blocks for the structure of petrovite. The arrangement of the clusters in shown in Fig. 4c. The four Na sites, Na2, Na3, Na4 and Na5, are fully occupied by Na. In the Na1 site, the essential amount of Ca is present (ca. 47%), which explains its high bond-valence sum (see below). The Na6 site is only partially occupied (53%). The Na1, Na2, Na3, Na5, Na6 and Na6’ atoms are surrounded by six O atoms each, forming distorted octahedra with the average bond lengths equal to 2.29, 2.37, 2.57, 2.35, 2.49 and 2.48 Å, respectively. The Na4 site is coordinated by seven O atoms with the Na4−O bond lengths in the range 2.347–2.993 Å.

The petrovite structure can also be described as a three-dimensional framework, taking into account the essential occupancy of the Na1 site by Ca2+ cations. In this description, the [Cu2(SO4)8]12− clusters are linked by Na1O6 octahedra, forming a porous three-dimensional framework (Fig. 4d) with cavities occupied by Na+ cations.

The bond-valence calculations were performed using empirical parameters taken from Brown and Altermatt (Reference Brown and Altermatt1985). The results are presented in Table 6. The high bond-valence sums of the S sites and the Na1 site are significantly overbonded, whereas the bond valence sum for Na3 is 0.61 valence units (vu). The deviation of the bond-valence sums from the expected values may be explained by the structural and positional disorder as well as the high ionic mobility of the Na atoms (see below).

Table 6. Bond-valence analysis (vu = valence units) for petrovite.

* The site is occupied by two cations

** The site is partially occupied

Discussion

Taking into account the essential admixture of Ca in the Na1 site, and the low occupancy of the Na6 site, the crystal chemical formula of petrovite can be written as Na6−2xCaxCu(SO4)4. For petrovite, x ≈ 0.5, which results in the idealised formula Na10CaCu2(SO4)8. However, at least theoretically, the x parameter may vary, resulting in different mineral species. For instance, for x = 0, 1, 1.5 and 2, the idealised formulae are Na6Cu(SO4)4, Na4CaCu(SO4)4, Na6Ca3Cu2(SO4)8 and Na2Ca2Cu(SO4)4, respectively. The x = 2 member of the series, Na2Ca2Cu(SO4)4, is chemically similar to itelmenite, Na2Mg2Cu(SO4)4, assuming the Mg-for-Ca replacement. However, the structure types of petrovite and itelmenite are different.

From the chemical point of view, petrovite demonstrates chemical and structural similarities to saranchinaite, Na2Cu(SO4)2. The three-dimensional [Cu(SO4)2]2− anionic framework in saranchinaite is formed by corner-sharing SO4 tetrahedra and CuOn polyhedra with the Na+ cations located in the cavities. In the crystal structure of petrovite, there are [Cu2(SO4)8]12− clusters of corner-sharing Cu and S polyhedra.

The chemical formula of petrovite can be obtained from that of saranchinaite through the hypothetical reaction:

$$2{\rm N}{\rm a}_2{\rm Cu}\lpar {{\rm S}{\rm O}_4} \rpar _2\lpar {{\rm saranchinaite}} \rpar + {\rm CaS}{\rm O}_4 + 3{\rm N}{\rm a}_2\lpar {{\rm S}{\rm O}_4} \rpar \to {\rm N}{\rm a}_{10}{\rm CaC}{\rm u}_2\lpar {{\rm S}{\rm O}_4} \rpar_8\lpar {{\rm petrovite}} \rpar.$$

The reaction implies the incorporation of the ionic CaSO4 and Na2(SO4) components into the largely covalent [Cu(SO4)2]2− anionic framework with the reduction of the dimensionality of the Cu sulfate polymeric network from three (in saranchinaite) to one (in petrovite). The crystal−chemical behaviour of this kind is known as the dimensional reduction, which has been described previously in oxides and oxysalts (Alekseev et al., Reference Alekseev, Krivovichev, Armbruster, Depmeier, Suleimanov, Chuprunov and Golubev2007; Krivovichev, Reference Krivovichev2009; Kovrugin et al., Reference Kovrugin, Gurzhiy and Krivovichev2012).

It is of interest that both petrovite and saranchinaite contain unusual Cu2+O7 polyhedra. According to Siidra et al. (Reference Siidra, Lukina, Nazarchuk, Depmeier, Bubnova, Agakhanov, Avdontseva, Filatov and Kovrugin2018), the presence of the Cu2+O7 polyhedra is one of the potential reasons for the crystallisation of saranchinaite in the non-centrosymmetric space group P21. Obviously, this kind of reasoning cannot be applied to petrovite, which both contains Cu2+O7 polyhedra and crystallises in the centrosymmetric space group P21/c.

Many chemical compounds and materials have been synthesised based on known mineral species (see, e.g. the recent work on synthetic saranchinaite by Siidra et al., Reference Siidra, Lukina, Nazarchuk, Depmeier, Bubnova, Agakhanov, Avdontseva, Filatov and Kovrugin2018). Therefore, the analysis of the functional properties of new minerals is important from the viewpoint of their material science applications. As the Cu2+/Cu3+ redox potential may provide a high operating voltage (Sun et al., Reference Sun, Rousse, Abakumov, Saubanère, Doublet, Rodríguez- Carvajal, Van Tendeloo and Tarascon2015), we have calculated the bond-valence energy landscape (BVEL) of petrovite using the BondStr software (Rodríguez-Carvajal, Reference Rodríguez-Carvajal2004). As a result, we have found that there are interconnected pathways for the Na+ migration in petrovite (Fig. 5), considering a percolation energy of 1.6 eV reported for the mobility of Na+ in polyanionic compounds (Boivin et al., Reference Boivin, Chotard, Bamine, Carlier, Serras, Palomares, Rojo, Iadecola, Dupont and Bourgeois2017; Kovrugin et al., Reference Kovrugin, Nekrasova, Siidra, Mentre, Masquelier, Stefanovich and Colmont2019). Thus, the petrovite structure type is promising as a cathode material. The theoretical capacity of petrovite based on the oversimplified formula Na6Cu(SO4)4 (x = 0) is 274.5 mAh/g, but the low Cu content decreases the theoretical capacity to ~46 mAh/g.

Fig. 5. The crystal structure of petrovite with the scheme of Na migration pathways. Migration pathways are calculated in BondStr software (Rodríguez-Carvajal, Reference Rodríguez-Carvajal2004).

Acknowledgements

This work was supported financially by the Russian Found of Basic Research, grant no. 18-29-12106. Technical support by the SPbSU X-ray Diffraction Centre is gratefully acknowledged.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1180/mgm.2020.53

Footnotes

Associate Editor: Irina O Galuskina

References

Alekseev, E.V., Krivovichev, S.V., Armbruster, T., Depmeier, W., Suleimanov, E.V., Chuprunov, E.V. and Golubev, A.V. (2007) Dimensional reduction in alkali metal uranyl molybdates: Synthesis and structure of Cs2[(UO2)O(MoO4)]. Zeitschrift für Anorganische und Allgemeine Chemie, 633, 19791984.CrossRefGoogle Scholar
Boivin, E., Chotard, J.-N., Bamine, T., Carlier, D., Serras, P., Palomares, V., Rojo, T., Iadecola, A., Dupont, L. and Bourgeois, L. (2017) Vanadyl-type defects in tavorite-like NaVPO4F: from the average long range structure to local environments. Journal of Material Chemistry A, 5, 2504425055.CrossRefGoogle Scholar
Britvin, S.N., Dolivo-Dobrovolsky, D.V. and Krzhizhanovskaya, M.G. (2017) Software for processing the X-ray powder diffraction data obtained from the curved image plate detector of Rigaku RAXIS Rapid II diffractometer. Proceedings of the Russian Mineralogical Society, 146, 104107 [in Russian].Google Scholar
Brown, I.D. and Altermatt, D. (1985) Bond-valence parameters obtained from a systematic analysis of the Inorganic Crystal Structure Database. Acta Crystallographica, B41, 244247.CrossRefGoogle Scholar
Fedotov, S.A. and Markhinin, Y.K. (1983) The Great Tolbachik Fissure Eruption. Cambridge University Press, New York.Google Scholar
Filatov, S.K., Karpov, G.A., Shablinskii, A.P., Krivovichev, S.V., Vergasova, L.P. and Antonov, A.V. (2013) Ivsite, Na3H(SO4)2, a new mineral from volcanic exhalations of fumaroles of the fissure Tolbachik eruption of the 50th anniversary of the Institute of Volcanology and Seismology, Far East Branch, Russian Academy of Sciences. Doklady Earth Sciences, 468, 632635.CrossRefGoogle Scholar
Filatov, S.K., Shablinskii, A.P., Vergasova, L.P., Saprikina, O.U, Bubnova, R.S, Moskaleva, S.V. and Belousov, A.B. (2019) Belomarinaite KNa(SO4): a new sulfate from 2012–2013 Tolbachik Fissure eruption, Kamchatka Peninsula, Russia. Mineralogical Magazine, 83, 569575.CrossRefGoogle Scholar
Filatov, S.K., Shablinskii, A.P., Vergasova, L.P., Krivovichev, S.V. and Moskaleva, S.V. (2020) Petrovite, IMA 2018-149b. CNMNC Newsletter No. 53; Mineralogical Magazine, 84, https://doi.org/10.1180/mgm.2020.5Google Scholar
Gorelova, L.A., Vergasova, L.P., Krivovichev, S.V., Avdontseva, E.Y., Moskaleva, S.V., Karpov, G.A. and Filatov, S.K. (2016) Bubnovaite, K2Na8Ca(SO4)6, a new mineral species with modular structure from the Tolbachik volcano, Kamchatka peninsula, Russia. European Journal of Mineralogy, 28, 677686.CrossRefGoogle Scholar
Hawthorne, F.C., Krivovichev, S.V. and Burns, P.C. (2000) The crystal chemistry of sulfate minerals. Pp. 1112 in: Sulfate minerals – Crystallography, Geochemistry and Environmental Significance (Alpers, C.N., Jambor, J.L. and Nordstrom, D.K., editors). Reviews in Mineralogy & Geochemistry, 40. Mineralogical Society of America and the Geochemical Society, Washington, DC.Google Scholar
Karpov, G.A., Krivovichev, S.V., Vergasova, L.P., Chernyat'eva, A.P., Anikin, L.P., Moskaleva, S.V. and Filatov, S.K. (2013) Oxysulfates of copper, sodium, and potassium in the lava flows of the 2012–2013 Tolbachik Fissure Eruption. Journal of Volcanology and Seismology, 7, 362370.CrossRefGoogle Scholar
Kovrugin, V.M., Gurzhiy, V.V. and Krivovichev, S.V. (2012) Structural topology and dimensional reduction in uranyl oxysalts: eight novel phases in the methylamine–(UO2)(NO3)2–H2SeO4–H2O system. Structural Chemistry, 23, 20032017.CrossRefGoogle Scholar
Kovrugin, V.M., Nekrasova, D.O., Siidra, O.I., Mentre, O., Masquelier, C., Stefanovich, S.Y. and Colmont, M. (2019) Mineral-inspired crystal growth and physical properties Na2Cu(SO4)2 and review of Na2M(SO4)2(H2O)x (x = 0–6) Compounds. Crystal Growth and Design, 19, 12331244.CrossRefGoogle Scholar
Krivovichev, S.V. (2009) Structural Crystallography of Inorganic Oxysalts. Oxford University press, Oxford, UK.CrossRefGoogle Scholar
Mandarino, J.A. (1981) The Gladstone-Dale relationship: Part IV. The compatibility concept and its application. The Canadian Mineralogist, 19, 441450.Google Scholar
Nadeem, M.A., Bhadbhade, M., Bircher, R. and Stride, J.A. (2010) Three isolated structural motifs in one crystal: penetration of two 1D chains through large cavities within 2D polymeric sheets. CrystEngComm, 12, 13911393.CrossRefGoogle Scholar
Nazarchuk, E.V., Siidra, O. I., Agakhanov, A.A., Lukina, E.A., Avdontseva, E. Y. and Karpov, G. A. (2018) Itelmenite, Na2CuMg2(SO4)4, a new anhydrous sulphate mineral from the Tolbachik volcano. Mineralogical Magazine, 82, 12331241.CrossRefGoogle Scholar
Pekov, I.V., Shchipalkina, N.V., Zubkova, N.V., Gurzhiy, V.V., Agakhanov, A.A., Belakovskiy, D.I., Chukanov, N.V., Lykova, I.S., Vigasina, M.F., Koshlyakova, N.N., Sidorov, E.G. and Giester, G. (2019) Alkali sulfates with aphthitalite-like structures from fumaroles of the Tolbachik volcano, Kamchatka, Russia. I. Metathénardite, a natural high-temperature modification of Na2SO4. The Canadian Mineralogist, 57, 885901.CrossRefGoogle Scholar
Pekov, I.V., Agakhanov, A.A., Zubkova, N.V., Koshlyakova, N.N., Shchipalkina, N.V., Sandalov, F.D., Yapaskurt, V.O., Turchkova, A.G. and Sidorov, E.G. (2020) Oxidizing-type fumarolic systems of the Tolbachik volcano – a mineralogical and geochemical unique. Russian Geology and Geophysics, https://doi.org/10.2205/2013ES000529Google Scholar
Petříček, V., Dusek, M. and Palatinus, L. (2006) Jana2006. The Crystallographic Computing System. Institute of Physics, Academy of Sciences of the Czech Republic, Prague.Google Scholar
Petrov, T.G. (2014) Separation-mixing as a model of composition evolution of any nature. Journal on Systemics, Cybernetics and Informatics, 12, 7681.Google Scholar
Petrov, T.G., Treywus, E.B. and Kasatkin, A.P. (1969) Crystal Growth from Solution. Plenum Publishing Corporation., New York 106 p.Google Scholar
Petrov, T.G., Andriyanets-Buyko, A. A., Moshkin, S. V. (2012) A two-parameter alphabet for coding structural–chemical information and its systematization (using the example of tourmaline). Automatic Documentation and Mathematical Linguistics, 46, 4049.CrossRefGoogle Scholar
Petrov, T.G., Protopopov, E.N. and Shuyskiy, A.V. (2013) Decorative grown malachite. Nature and technology. Russian Journal of Earth Sciences, 13, https://doi.org/10.2205/2013ES000529CrossRefGoogle Scholar
Rodríguez-Carvajal, J. (2004) The program BondStr and its GUI GBondStr. http://www.ccp14.ac.uk/ccp/web-mirrors/plotr/BondStr/Bond_Str.htmGoogle Scholar
Shablinskii, A.P., Filatov, S.K., Vergasova, L.P., Moskaleva, S.V., Avdontseva, E.Y. and Bubnova, R.S. (2020) Dobrovolskyite, IMA 2019-106, in: CNMNC Newsletter 54, Mineralogical Magazine, 84, 359365.Google Scholar
Shchipalkina, N.V., Pekov, I.V., Chukanov, N.V., Zubkova, N.V., Belakovskiy, D.I., Koshlyakova, N.N., Britvin, S.N., Sidorov, E.G. and Vozchikova, S.A. (2020) Alkali sulfates with aphthitalite-like structures from fumaroles of the Tolbachik volcano, Kamchatka, Russia. II. A new mineral, natroaphthitalite, and new data on belomarinaite. The Canadian Mineralogist, 58, 167181.CrossRefGoogle Scholar
Siidra, O.I., Nazarchuk, E.V., Zaitsev, A.N., Lukina, E.A., Avdontseva, E.Y., Vergasova, L.P., Vlasenko, N.S., Filatov, S.K. and Karpov, G.A. (2017) Copper oxosulphates from fumaroles of Tolbachik volcano: puninite, Na2Cu3O(SO4)3 – a new mineral species and structure refinements of kamchatkite and alumoklyuchevskite. European Journal of Mineralogy, 29, 499510.CrossRefGoogle Scholar
Siidra, O.I., Lukina, E.A., Nazarchuk, E.V., Depmeier, W., Bubnova, R.S., Agakhanov, A.A., Avdontseva, E.Y., Filatov, S.K. and Kovrugin, V.M. (2018) Saranchinaite, Na2Cu(SO4)2, a new exhalative mineral from Tolbachik volcano, Kamchatka, Russia, and a product of the reversible dehydration of kröhnkite, Na2Cu(SO4)2(H2O)2. Mineralogical Magazine, 82, 257274.CrossRefGoogle Scholar
Siidra, O.I., Nazarchuk, E.V., Zaitsev, A.N. and Vlasenko, N.S. (2020) Koryakite, NaKMg2Al2(SO4)6, a new NASICON-related anhydrous sulfate mineral from Tolbachik volcano, Kamchatka, Russia. Mineralogical Magazine, 84, 283287.CrossRefGoogle Scholar
Sun, M., Rousse, G., Abakumov, A.M., Saubanère, M., Doublet, M.L., Rodríguez- Carvajal, J., Van Tendeloo, G. and Tarascon, J.M. (2015) Li2Cu2O(SO4)2: A Possible Electrode for Sustainable Li-Based Batteries Showing a 4.7 V Redox Activity vs Li+/Li0. Chemistry of Materials, 27, 30773087.CrossRefGoogle Scholar
Vergasova, L.P. and Filatov, S.K. (2012) New mineral species in products of fumarole activity of Great Tolbachik fissure eruption. Journal of Volcanology and Seismology, 6, 281289.CrossRefGoogle Scholar
Vergasova, L.P. and Filatov, S.K. (2016) A study of volcanogenic exhalation mineralization. Journal of Volcanology and Seismology, 10, 7185.CrossRefGoogle Scholar
Yoshiasa, A., Yagyu, G., Ito, T., Yamanaka, T. and Nagai, T. (2000) Crystal structure of the high pressure phase(II) in CuGeO3. Zeitschrift für Anorganische und Allgemeine Chemie, 626, 3641.3.0.CO;2-2>CrossRefGoogle Scholar
Figure 0

Fig. 1. A view of the fissure of the micrograben on the west side of the Second Cinder cone of Great Tolbachik fissure eruption (photo taken in 1981).

Figure 1

Fig. 2. Scanning electron microscopy image of an individual grain of petrovite. Part of the petrovite sample stored in the Mineralogical Museum of St. Petersburg State University (1/19696).

Figure 2

Fig. 3. Blue cryptocrystalline crusts of petrovite enveloped by fine pyroclastic material. Parts of the petrovite sample stored in the Mineralogical Museum of St. Petersburg State University (1/19696).

Figure 3

Table 1. Chemical composition of petrovite (wt.%).

Figure 4

Table 2. Powder X-ray diffraction data (d in Å) for petrovite.

Figure 5

Table 3. Crystal data, data collection information and structure refinement details for petrovite.

Figure 6

Table 4. Atomic coordinates and displacement parameters (Å2) for petrovite.

Figure 7

Table 5. Selected interatomic distances (Å) for petrovite.

Figure 8

Fig. 4. The crystal structure of petrovite: (a) CuO7 polyhedron; (b) the [Cu2(SO4)8]12– cluster; (c) arrangement of the Cu2(SO4)8 clusters in the structure; and (d) three-dimensional framework of the crystal structure.

Figure 9

Table 6. Bond-valence analysis (vu = valence units) for petrovite.

Figure 10

Fig. 5. The crystal structure of petrovite with the scheme of Na migration pathways. Migration pathways are calculated in BondStr software (Rodríguez-Carvajal, 2004).

Supplementary material: File

Filatov et al. supplementary material

Filatov et al. supplementary material

Download Filatov et al. supplementary material(File)
File 174.8 KB