Hostname: page-component-7bb8b95d7b-lvwk9 Total loading time: 0 Render date: 2024-10-01T07:41:48.613Z Has data issue: false hasContentIssue false

Deynekoite, Ca9□Fe3+(PO4)7 – a new mineral of the merrillite group from phosphide-bearing contact facies of paralava, Hatrurim Complex, Daba-Siwaqa, Jordan

Published online by Cambridge University Press:  11 September 2023

Evgeny V. Galuskin*
Affiliation:
Institute of Earth Sciences, Faculty of Natural Sciences, University of Silesia, Będzińska 60, 41-200 Sosnowiec, Poland
Marcin Stachowicz
Affiliation:
Institute of Geochemistry, Mineralogy and Petrology, Faculty of Geology, University of Warsaw, Żwirki i Wigury 93, 02-089 Warszawa, Poland
Irina O. Galuskina
Affiliation:
Institute of Earth Sciences, Faculty of Natural Sciences, University of Silesia, Będzińska 60, 41-200 Sosnowiec, Poland
Krzysztof Woźniak
Affiliation:
Department of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warszawa, Poland
Yevgeny Vapnik
Affiliation:
Department of Geological and Environmental Sciences, Ben-Gurion University of the Negev, P.O.B. 653, Beer-Sheva 84105, Israel
Mikhail N. Murashko
Affiliation:
Institute of Earth Sciences, Saint-Petersburg State University, Universitetskaya Nab. 7/9, 199034 St. Petersburg, Russia
Grzegorz Zieliński
Affiliation:
Polish Geological Institute – National Research Institute, Rakowiecka 4, 00-975 Warsaw, Poland
*
Corresponding author: Evgeny V. Galuskin; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Deynekoite, Ca9□Fe3+(PO4)7 (R3c, a = 10.3516(3)Å, c = 37.1599(17)Å, V = 3448.4(3)Å3 and Z = 6), a new mineral of the merrillite group was found in the contact facies of paralava of the Hatrurim Complex in the Daba-Siwaqa pyrometamorphic rock field, Jordan. The paralava, consisting of diopside, tridymite, anorthite, wollastonite and fluorapatite, is enriched in Fe-bearing phosphides and phosphates at the contact with the altered country rock. Cristobalite overgrowing tridymite has a fish-scales texture indicating that temperature of paralava could have reached 1500°C. Deynekoite with empirical formula (Ca8.90Na0.11K0.02)Σ9.03(Fe3+0.62Mg0.30Al0.05)Σ0.97P6.98V5+0.05O27.70(OH)0.30 forms transparent, light-yellow or light-brown grains up to 30–40 μm in size. Microhardness of deynekoite, VHN25 = 319(29) kg/mm2, corresponds to Mohs hardness = 4.5. Its density was calculated as 3.09 g⋅cm–3 on the basis of its empirical composition and structural data. Deynekoite is uniaxial (−), its refractive indices are ω = 1.658(3), ɛ = 1.652(3) (λ = 589 nm), and pleochroism is not observed. The formation of phosphides on the boundary of the paralava and country rock is connected with carbothermal reductive reactions and realised at temperatures above 1300°С. With decreasing temperature and increasing oxygen activity, phosphides are replaced by Fe2+-bearing phosphates. Deynekoite, which contains Fe3+ (substituting for Fe2+-phosphates) and a small amount of water, formed at temperatures of 600–800°C.

Type
Article
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Mineralogical Society of the United Kingdom and Ireland

Introduction

Deynekoite, Ca9□Fe3+(PO4)7 (R3c, a = 10.3516(3) Å, c = 37.1599(17) Å, V = 3448.4(3) Å3 and Z = 6), a new mineral isostructural with merrillite, was found in the phosphide-bearing contact facies of diopside–anorthite–tridymite paralava of the Hatrurim Complex in Jordan. According to the cerite supergroup classification, deynekoite belongs to the merrillite subgroup. Together, the merrillite subgroup and the whitlockite subgroup form the merrillite group (Table 1; Atencio and Azzi, Reference Atencio and Azzi2020). The general formula of minerals of the merrillite group can be presented as follows: (A13A23A33)Σ9XM(P1O4)3 (P2O3Ø)3(P3O3Ø), where A = Ca, Sr and Na; X = Na, Ca and vacancy (□); M = Mg, Fe2+,3+ and Mn2+; P = P; Ø = O and OH. Deynekoite is the first mineral of the merrillite group whose composition includes Fe3+; it has a synthetic analogue (Lazoryak et al., Reference Lazoryak, Morozov, Belik, Khasanov and Shekhtman1996; Benarafaa et al., Reference Benarafaa, Kacimia, Gharbagea, Millet and Ziyada2000; Deyneko et al., Reference Deyneko, Aksenov, Morozov, Stefanovich, Dimitrova, Barishnikova and Lazoryak2014).

Table 1. Minerals of the merrillite group.

* Hedegaardite was approved by the CNMNC-IMA in 2015, however a paper describing the mineral has not been published yet, so the most probable cation distribution among different sites is shown.

Minerals of the merrillite group (merrillite, Ca9NaMg(PO4)7, keplerite, Ca9(Ca0.50.5)Mg(PO4)7, and the potentially new mineral Ca8Y□Mg(PO4)7, forming a solid solution) were found recently in hematite-bearing diopside paralava hosted by the rocks of the ‘olive unit’, in the Negev Desert, Hatrurim Basin, Israel (Galuskina et al., Reference Galuskina, Galuskin and Vapnik2016; Britvin et al., Reference Britvin, Galuskina, Vlasenko, Vereshchagin, Bocharov, Krzhizhanovskaya, Shilovskikh, Galuskin, Vapnik and Obolonskaya2021b).

Merrillite is one of the most common phosphates in meteorites of different types and forms a solid solution with ferro-merrillite, Ca9NaFe2+(PO4)7, which is typical for Martian shergottites (Britvin et al., Reference Britvin, Krivovichev and Armbruster2016; Ward et al., Reference Ward, Bischoff, Roszjar, Berndt and Whitehouse2017). OH-bearing minerals of the merrillite group (whitlockite subgroup) occur only in terrestrial rocks (Frondel, Reference Frondel1943; Britvin et al., Reference Britvin, Pakhomovskii, Bogdanova and Skiba1991; Cooper et al., Reference Cooper, Hawthorne, Abdu, Ball, Ramik and Tait2013).

In this paper, we provide a description of a new mineral, deynekoite, and discuss common problems in the classification of the merrillite group. We also consider the mechanism and conditions of deynekoite formation in phosphide-bearing contact facies of paralava from Jordan. The mineral and its name (symbol Dnk) have been approved by the Commission on New Minerals, Nomenclature and Classification (CNMNC) of the International Mineralogical Association (IMA), as IMA 2020-108 (Galuskin et al., Reference Galuskin, Stachowicz, Galuskina, Woźniak, Vapnik, Murashko and Zieliński2022b). The name deynekoite is given in honour of Dr Dina V. Deyneko (born 1988) for her valuable contribution to the investigation of synthetic analogues of merrillite-group minerals (Deyneko et al., Reference Deyneko, Aksenov, Morozov, Stefanovich, Dimitrova, Barishnikova and Lazoryak2014). Type material was deposited in the mineralogical collection of the Fersman Mineralogical Museum, Leninskiy pr., 18/k. 2, 115162 Moscow, Russia, registration number: 5791/1.

Methods of investigation

The morphology and composition of deynekoite and associated minerals were studied using optical microscopy, scanning electron microscopes (Phenom XL and Quanta 250, Institute of Earth Sciences, Faculty of Natural Sciences, University of Silesia, Sosnowiec, Poland) and an electron microprobe analyser (Cameca SX100, Micro-Area Analysis Laboratory, Polish Geological Institute – National Research Institute, Warsaw, Poland). Chemical analyses were carried out in WDS-mode (wavelength-dispersive X-ray spectroscopy, settings: 15 keV, 20 nA and ~1 μm beam diameter) using the following lines and standards: NaKα – NaCl; CaKα and MgKα – diopside; AlKα and KKα – orthoclase; FeKα – Fe2O3; and VKα – V. Other chemical elements were below the detection limit.

The Raman spectra of deynekoite were recorded on a WITec alpha 300R Confocal Raman Microscope (Department of Earth Science, University of Silesia, Poland) equipped with an air-cooled solid laser (532 nm) and a CCD camera operating at –61°C. An air Zeiss LD EC Epiplan-Neofluan DIC-100/0.75NA objective was used. Raman scattered light was focused onto a multi-mode fibre and monochromator with a 600 mm–1 grating. The power of the laser at the sample position was ~30 mW, and 15–20 scans with an integration time of 3–5 s were collected and averaged. The resolution was 3 cm–1. The spectrometer monochromator was calibrated using the Raman scattering line of a silicon plate (520.7 cm–1).

Single-crystal X-ray studies were carried out with a four-circle diffractometer SuperNova with AgKα radiation (λ = 0.56087 Å) (University of Warsaw Biological and Chemical Research Centre), equipped with an Eos CCD detector (Agilent). The detector-to-crystal distance was 63.2 mm. AgKα radiation was used at 65 kV and 0.6 mA. Crystals were attached to a non-diffracting Mitegen micromount support. A frame-width of 1° in ω scans and a frame time of 410 s were used for data collection. Information relevant to the data collection is summarised in Table 2. Reflection intensities were corrected for Lorentz, polarisation and absorption effects and converted to structure factors using CrysAlisPro 1.171.40.53 (Rigaku Oxford Diffraction, Reference Rigaku Oxford Diffraction2019) software.

Table 2. Crystal data and structure refinement details for deynekoite.

*w = 1/[Σ2(F o2) + (0.0236P)2], where P is [2Fc2 + Fo2)]/3

Geological setting

Deynekoite was found in a small phosphorite prospecting quarry (31°22'01''N, 36°11'10''E) in the Daba-Siwaqa pyrometamorphic rock field, Hatrurim Complex, Jordan (Novikov et al., Reference Novikov, Vapnik and Safonova2013; Khoury et al., Reference Khoury, Sokol, Kokh, Seryotkin, Nigmatulina, Goryainov, Belogub and Clark2016). The Hatrurim Complex is distributed along the Dead Sea rift as a series of outcrops of pyrometamorphic rocks in the territories of Israel, Palestine and Jordan (Bentor, Reference Bentor1960; Gross, Reference Gross1977, Reference Gross1984; Vapnik et al., Reference Vapnik, Sharygin, Sokol and Shagam2007; Novikov et al., Reference Novikov, Vapnik and Safonova2013). The Hatrurim Complex is characterised by a wide diversity of rocks. Spurrite marble, larnite pseudoconglomerate and gehlenite hornfels dominate; however, there are also exotic rock types containing such rock-forming minerals as ye'elimite, Ca4Al6(SO4)O12; fluormayenite, Ca12Al14O32F2; nabimusaite, KCa12(SiO4)4(SO4)2O2F; ariegilatite, BaCa12(SiO4)4(PO4)2F2O; jasmundite, Ca11(SiO4)4O2S; ternesite, Ca5(SiO4)2(SO4); levantite, KCa3Al2(SiO4)(Si2O7)(PO4); silicocarnotite, Ca5(PO4)2(SiO4); oldhamite, CaS2; khesinite, Ca4(Mg2Fe3+10)O4(Fe3+20Si2)O36; and others (Galuskina et al., Reference Galuskina, Galuskin, Pakhomova, Widmer, Armbruster, Krüger, Grew, Vapnik, Dzierażanowski and Murashko2017a; Galuskin et al., Reference Galuskin, Galuskina, Gfeller, Krüger, Kusz, Vapnik, Dulski and Dzierżanowski2016, Reference Galuskin, Krüger, Galuskina, Krüger, Vapnik, Pauluhn and Olieric2019, Reference Galuskin, Galuskina, Krüger, Krüger, Vapnik, Krzątała, Środek and Zieliński2021, Reference Galuskin, Galuskina, Vapnik and Zieliński2023a). In the Hatrurim Complex, molten rocks that are diverse in composition form veins, lenses and oval bodies of paralavas and slag-like rocks from a few centimetres to tens of metres in size (Galuskin et al., Reference Galuskin, Gfeller, Galuskina, Pakhomova, Armbruster, Vapnik, Włodyka, Dzierżanowski and Murashko2015; Galuskina et al., Reference Galuskina, Galuskin, Pakhomova, Widmer, Armbruster, Krüger, Grew, Vapnik, Dzierażanowski and Murashko2017a, Reference Galuskina, Galuskin, Prusik, Vapnik, Juroszek, Jeżak and Murashko2017b, Reference Galuskina, Galuskin, Vapnik, Prusik, Stasiak, Dzierżanowski and Murashko2017c; Krzątała et al., Reference Krzątała, Krüger, Galuskina, Vapnik and Galuskin2020; Krüger et al., Reference Krüger, Galuskin, Galuskina, Krüger and Vapnik2021; Murashko et al., Reference Murashko, Britvin, Vapnik, Polekhovsky, Shilovskikh, Zaitsev and Vereshchagin2022). It is the paralavas, which feature relatively coarser-grained rocks compared to other types of pyrometamorphic rocks, that are associated with the discovery of a significant number of new minerals. One example is the diopside–anorthite–tridymite paralava which forms a body ~30 m in diameter in pyrometamorphically altered carbonate rock of the Muwaqqar Chalk–Marl Formation, Daba-Siwaqa, Jordan. Deynekoite and a series of new minerals presented by both reduced (phosphides) and oxidised (phosphates containing Fe3+) phases were discovered at the contact facies of this paralava. The phosphides discovered were transjordanite, Ni2P; zuktamrurite, FeP2; murashkoite, FeP; orishchinite, Ni2P; and nickolayite, FeMoP. The phosphates were crocobelonite, CaFe3+(PO4)O; moabite, NiFe3+(PO4)O; yakubovichite, CaNi2Fe3+(PO4)3; and nazarchukite, Ca2NiFe3+2(PO4)4 (Britvin et al., Reference Britvin, Murashko, Vapnik, Polekhovsky, Krivovichev, Vereshchagin, Vlasenko, Shilovskikh and Zaitsev2019a, Reference Britvin, Vapnik, Polekhovsky, Krivovichev, Krzhizhanovskaya, Gorelova, Vereshchagin, Shilovskikh and Zaitsev2019b, Reference Britvin, Murashko, Vapnik, Polekhovsky, Krivovichev, Krzhizhanovskaya, Vereshchagin, Shilovskikh and Vlasenko2020, Reference Britvin, Murashko, Krzhizhanovskaya, Vapnik, Vlasenko, Vereshchagin, Pankin and Vasiliev2021a, Reference Britvin, Galuskina, Vlasenko, Vereshchagin, Bocharov, Krzhizhanovskaya, Shilovskikh, Galuskin, Vapnik and Obolonskaya2021b, Reference Britvin, Murashko, Vapnik, Zaitsev, Shilovskikh, Krzhizhanovskaya, Gorelova, Vereshchagin, Vasilev and Vlasenko2022a, Reference Britvin, Murashko, Krzhizhanovskaya, Vereshchagin, Vlasenko, Vapnik and Bocharov2022b, Reference Britvin, Murashko, Krzhizhanovskaya, Vlasenko, Vereshchagin, Vapnik and Bocharov2023a, Reference Britvin, Murashko, Krzhizhanovskaya, Vapnik, Vlasenko, Vereshchagin, Pankin, Zaitsev and Zolotarev2023b; Murashko et al., Reference Murashko, Britvin, Vapnik, Polekhovsky, Shilovskikh, Zaitsev and Vereshchagin2022).

The genesis of the Hatrurim Complex rocks remains an unsolved problem. The early ‘classic hypothesis’ stated that pyrometamorphic rocks formed as a result of the spontaneous combustion of bitumen, which was contained in a sedimentary protolith (Kolodny and Gross, Reference Kolodny and Gross1974; Matthews and Gross, Reference Matthews and Gross1980; Geller et al., Reference Geller, Burg, Halicz and Kolodny2012). The relatively recent ‘mud volcano’ hypothesis suggests that deep anomalous pressure in the sedimentary sequence caused by the tectonic activity of the Dead Sea Rift affected the destruction of hydrocarbon collectors, inducing subsequent methane combustion in the vicinity of the surface (Sokol et al., Reference Sokol, Novikov, Zateeva, Vapnik, Shagam and Kozmenko2010; Novikov et al., Reference Novikov, Vapnik and Safonova2013). Our investigation shows that the reactions of combustion by-products (hot gases, fluids and melts) with the pyrometamorphic phases formed earlier sharply increased the mineral diversity of the rocks (Galuskin et al., Reference Galuskin, Galuskina, Gfeller, Krüger, Kusz, Vapnik, Dulski and Dzierżanowski2016, Reference Galuskin, Krüger, Galuskina, Krüger, Vapnik, Pauluhn and Olieric2019).

Deynekoite occurrence

The paralava studied presents a dense, fine-grained rock in which porous, pumice-like fragments are distributed in an undulating pattern (Fig. 1a). It consists of diopside with hedenbergite rims, wollastonite, anorthite, tridymite, fluorapatite and a small amount of glass (Fig. 1b). The presence of cristobalite with typical fissures showing a fish-scale texture indicates that tetragonal cristobalite is a product of phase transition and the primary crystallisation of the mineral was as a high-temperature cubic phase, growing on tridymite (Fig. 1c). In thin apophyses of paralava into the country rock, pseudomorphs of fluorapatite after fish bones were detected (Fig. 1d). Accessory minerals of the paralava include spinel of the magnesiochromite–chromite–magnetite–trevorite series, titanite and pyrrhotite.

Figure 1. (a) Heterogeneous, porous basalt-like rock from a central part of the paralava body. (b) Paralava consisting of diopside, tridymite, anorthite and wollastonite. Back-scattered electron (BSE) image. (c) Cristobalite with fish-scale texture fissures overgrowing tridymite. BSE image. (d) Pseudomorph of fluorapatite after fish bone in the paralava. BSE image. An = anorthite, Crs = cristobalite, Di = diopside, Fap = fluorapatite, Hd = hedenbergite, Gls = glass, Mgt = magnetite, Trd = tridymite, Tnt = titanite, Wol = wollastonite.

Minerals of the merrillite group forming a solid solution of merrillite–keplerite–deynekoite–whitlockite occur in a thin dark zone on the boundary of paralava and altered country rock (Fig. 2a). A contact facies of paralava has been strongly altered by low-temperature processes. Here, prismatic diopside crystals with composition close to ideal and up to 1 cm in size are embedded in a low-temperature mineral matrix containing primarily calcite and zeolites, among which gismondine-Ca prevails (Fig. 2b–d). Locally, rocks are enriched in minerals of the tobermorite, ettringite groups and unidentified opal-like Ca-hydrosilicates–phosphates.

Figure 2. (a) Contact of diopside paralava (I) and altered marl (II), black is a hematite-bearing zone. The white frame shows a fragment magnified in Fig. 2b. (b) Position of phosphate mineralisation at the contact zone of rocks: 1 – diopside paralava, 2 – zone enriched with phosphides (murashkoite, zuktamrurite), 3 – phosphate zone, 4 – hematite zone (developed after pyrrhotite), 5 – altered host rock (marl); diopside chondrules are shown by arrows. (c) Phosphate zone represented, mainly, by deynekoite. Diopside paralava is replaced intensively by calcite and gismondine-Ca; in unaltered diopside crystals, barringerite inclusions are noted. The white frame shows a fragment magnified in Fig. 2d; (d) One of the rare places where deynekoite forms almost monomineralic aggregates of elongated grains up to 50 μm in size. Bgr = barringerite, Dnk = deynekoite, Di = diopside, Fap = fluorapatite, Hem = hematite, Gsm = gismondine-Ca, Muh = murashkoite.

In the contact facies of the paralava, needle-like inclusions of 1–2 μm in thickness of barringerite (Fig. 1c) and oldhamite are observed in diopside crystals. Small cracks in the rock ~0.1 mm in thickness are filled with aggregates of phosphides of the barringerite–transjordanite series replaced by murashkoite. Rare grains of Ti-bearing pyrrhotite, daubréelite and blue Ti3+-bearing titanite were also noted.

A dark zone at the contact of the paralava with the altered country rock is characterised by zonation in the direction from paralava to country rock (Fig. 2b). The phosphide zone, where barringerite, transjordanite, murashkoite and zuktamrurite were identified, is gradually replaced by a phosphate zone. The phosphate zone consists mainly of minerals of the merrillite group and fluorapatite. Here, we identified graftonite-(Ca), CaFe2+2(PO4)2; crocobelonite, CaFe3+2(PO4)2O; Fe-analogue of stanfieldite, Ca4MgFe2+4(PO4)6; and β-Fe2P2O7 – a high-temperature analogue of nabateaite, γ-Fe2P2O7, discovered recently in phosphide-bearing paralava in Israel (Britvin et al., Reference Britvin, Murashko, Vapnik, Vlasenko, Vereshchagin, Krzhizhanovskaya and Bocharov2021c).

The phosphate zone is followed by a hematite zone, in which relics of Fe–Ni spinel and pyrrhotite are preserved, and on the boundary of which the country rock is transformed into chondrite-like rock (Fig. 2b). Individual diopside crystals are observed in all the described zones (Fig. 2b,c). Fe2+-bearing phosphates tend to occur next to phosphides, whereas Fe3+-bearing phosphates are usually in contact with hematite. Grained aggregates of deynekoite were found as intergrowths with hematite in several polished mounts. Typically, fluorapatite is intergrown with hematite (Fig. 2c,d). Small relics of Fe2+-bearing phosphates and berlinite, AlPO4, occurring in deynekoite deserve attention. The morphology of the hematite, which has a lattice-like form, indicates that it formed after the pyrrhotite crystals (Fig. 2d).

Deynekoite forms aggregates with grains up to 30–40 μm in size. The grains are transparent and light-yellow and light-brown in colour. The streak is white with a yellowish tint. The microhardness is VHN25 = 319(29), 253–331 (in kg/mm2), which corresponds to a Mohs hardness of 4.5. The mineral is brittle with a conchoidal fracture, and cleavage is not observed. Its density, calculated on the basis of the empirical composition and the structural data, is 3.09 g⋅cm–3. Deynekoite is uniaxial (−), its refractive indices are ω = 1.658(3), ε = 1.652(3) (λ = 589 nm), and pleochroism was not observed.

Deynekoite has a relatively invariable chemical composition (Tables 1, 3), and it can be described by the mean empirical formula (Ca8.90Na0.11K0.02)Σ9.03(Fe3+0.62Mg0.30Al0.05)Σ0.97P6.98V5+0.05O27.70(OH)0.30, which can be simplified to Ca9.00(Fe3+0.70Mg0.30)Σ1.0[(PO4)6.70 (PO3OH)0.3]Σ7, and the end-member formula Ca9Fe3+(PO4)7.

Table 3. Chemical composition of deynekoite.

* Water was calculated by charge balance; ** (OH) apfu; S.D. – standard deviation

Raman spectroscopy

In the Raman spectrum of deynekoite (Fig. 3) most of the bands are related to vibrations in (РО4)3– groups: bands in the range 1200–1000 cm–1 are connected with vibrations ν3(PO4)3–; 1000–900 cm–1 with ν1(PO4)3–; 550–630 cm–1 with ν4(PO4)3–; and 400–470 cm–1 with ν2(PO4)3–. Bands lower than 300 cm–1 are ascribed to vibrations of Са–О and lattice vibrations. In the deynekoite spectrum there are two bands, at 934 cm–1 and 880 cm–1, which have low intensity but appear in all its spectra. These bands are not noted in merrillite and keplerite spectra (Britvin et al., Reference Britvin, Galuskina, Vlasenko, Vereshchagin, Bocharov, Krzhizhanovskaya, Shilovskikh, Galuskin, Vapnik and Obolonskaya2021b). The band at 934 cm–1 is related to P–O vibrations in the PO3OH group. The same band is observed in the spectrum of whitlockite (Jolliff et al., Reference Jolliff, Hughes, Freeman and Zeigler2006). The band at 880 cm–1 is connected with the ν1(VO4)3– vibration. As has been shown for the pyromorphite–vanadinite series, even for a low content, when V5+ substitutes for P5+ at the tetrahedra, bands from the stretching vibration of V–O in ν1(VO4)3– are clearly visible in the Raman spectra (Solecka et al., Reference Solecka, Bajda, Topolska, Zelek-Pogudz and Manecki2018). The OH-group content in deynekoite is insignificant, so bands from O–H stretching vibrations within the characteristic range 3000–3700 cm–1 are not observed (Fig. 3).

Figure 3. Raman spectrum of deynekoite.

Deynekoite structure

The crystal structure of deynekoite was solved in the R3c space group with a dual-space algorithm implemented in Shelxt (Sheldrick, Reference Sheldrick2015a). The model was refined with least-squares minimisation using Shelxl (Sheldrick, Reference Sheldrick2015b), with Olex2 (Dolomanov et al., Reference Dolomanov, Bourhis, Gildea, Howard and Puschmann2009) as the graphical interface. Twinning by inversion was observed with two domains of 56:44 ratios. The occupancy of the M site in deynekoite was constrained to 1 and refined as Fe vs. Mg. Disorder in the P 1PO4 tetrahedron was observed, where the P1 and O10 sites split into two parts, forming P1A, O10A and P1B, O10B groups of atoms. The occupancy of P 1PO4 was constrained to 1 and the occupancies of the disordered atoms were refined as part 1 vs. part 2. The H10b atom was fixed at 0.96 Å from O10B in the direction of the O8 atom to form an H-bond interaction. The occupancy of H10b was refined together with P1B and O10B in the ‘part 2’ group of the disordered atoms. Experimental details and refinement data are summarised in Tables 2, 4–6. The crystallographic information file has been deposited with the Principal Editor of Mineralogical Magazine and is available as Supplementary material (see below). In the deynekoite structure a series of polyhedra are distinguished, parts of which are fully occupied by cations of one type. There are tetrahedrally coordinated phosphorus sites P2 and P3, eight-coordinated Ca sites (Ca1, Ca2 and Ca3) and a number of oxygen sites (Table 4). The M site in minerals of the merrillite structure is always fully occupied (Britvin et al., Reference Britvin, Galuskina, Vlasenko, Vereshchagin, Bocharov, Krzhizhanovskaya, Shilovskikh, Galuskin, Vapnik and Obolonskaya2021b). In the case of deynekoite, this site is occupied by Fe3+, Mg and Al. Refinement of the site occupation gave (Fe3+0.58Mg0.42), which yields a charge of +2.58 and 20.12 е . Data from the microprobe analyses gave the occupation of the M site (Fe3+0.62Mg0.30Al0.05), which yields a charge of +2.61 and 20.37 е . Considering that it is impossible to separate Mg and Al during structure refinement, the data regarding the occupation of the M site are relevant to the microprobe analyses data, although it cannot be ruled out that the grain used for the structure refinement has a slightly lower Fe3+ content. The tetrahedral site P1 splits into sites P1A and P1B with refined occupation in a ratio of 0.73:0.27. The occupation ratio of the О10А and О10В sites is identical; the atoms are located at the top of the tetrahedra and oriented to the opposite sides. The bases of both tetrahedra are formed by oxygen О9–О9–О9. The tetrahedron P 1APO4 has the typical interatomic distances P1A–O9 = 1.534(4) Å and P1A–O10A = 1.526(14) Å (Table 6). The first variant of the structure refinement of deynekoite, submitted in the new-mineral proposal check-list, reported an anomalously large distance P1В–O10В = 1.78(5) Å and normal distances P1В–O9 = 1.517(6) Å×3. In whitlockite the distance P1В–(O10BH) ≈ 1.61 Å (Hughes et al., Reference Hughes, Jolliff and Rakovan2008). The anomalous bond length may be due to the presence of a very small amount of Na at the X site, which overlaps with the O10B position in deynekoite. We have taken into consideration the X site in the revised deynekoite structure model and have refined the Na content at this site equal to 0.1 atoms per formula unit (apfu). It may be that the X site has the lower occupation as it can contain K (Table 2). In the revised model, the distances in the P PO4 tetrahedron are P1В–O10В = 1.64(5) Å and P1В–O9 = 1.519(6) Å ×3 (Table 5). The increase in the P1В–O10В distance in comparison with P1А–O10А is connected with the partial protonation of O10B with the formation of strong hydrogen bond O10B⋅⋅⋅O8 with the distance d O10B–O8 = 2.68(3)Å and ∠O–H⋅⋅⋅O = 175.5° (see Fig. 5e). The deynekoite structure, which is formed by the intercalation of two types of layers, is shown in Fig. 4.

Figure 4. Deynekoite structure. (a) Projection on (010). The structure can be presented as the intercalation of two types of layers (see Fig. 5) in the sequence ABA'B’A’’B”A. (b) Projection on (001). One slice of the A layer and two slices of B layers are shown. The Na site with lower occupation is not shown. Ca atoms – light-orange; Na atoms – yellow; P2PO4 tetrahedron – dark-blue; P3PO4 tetrahedron – blue; P 1APO4 tetrahedron – dark-green; P 1BPO4 tetrahedron – light-green; МО6 octahedron – brown; oxygen O10B – dark blue balls; oxygen O10A – blue balls; hydrogen – small pink balls.

Figure 5. (a,b) Layers of two types (А, В see Fig. 4) in the structure of deynekoite, projection on (00$\bar{1}$). (a) In the B layer formed by six-membered rings building from Ca-polyhedra, three types of triangular empty occupation are shown as in the ideal structures of merrillite (1), deynekoite (2) and whitlockite (3). (b) The A layer is the same in merrillite, deynekoite and whitlockite, consisting of millwheel clusters M(P2+P3PO4)6 linked with one another by triplets of Ca-polyhedra. (c–e) Atom arrangement near the P1 site in holotype deynekoite: c – merrillite type (10%); d – deynekoite type (63%); e – whitlockite type (27%). Ca-polyhedra – light-orange; P2PO4 tetrahedron – dark-blue; P3PO4 tetrahedron – blue; P 1APO4 tetrahedron – dark-green; P 1BPO4 tetrahedron – light-green; truncated prism XO6 – yellow; octahedron МО6 – brown (Fe3+)/dark-orange (Mg); oxygen – blue balls; hydrogen – pink balls (hydrogen bonds are shown); empty circle – vacancy.

Table 4. Fractional atomic coordinates and isotropic* or equivalent isotropic displacement parameters (Å2) for deynekoite.

Table 5. Atomic displacement parameters (Å2) for deynekoite.

Table 6. Selected bond lengths (Å) and bond-valence sum (BVS) calculation for deynekoite.

As deynekoite occurs only in tiny amounts and its crystals contain large amounts of inclusions of other phases, powder X-ray diffraction data were not collected, as it could be calculated more reliably from the results of single-crystal structure refinements. The calculated data are listed in Supplementary Table S1.

Discussion

Nomenclature problems

The cerite supergroup combines two groups, cerite and merrillite, of which the latter is divided into two more subgroups: merrillite and whitlockite (Atencio and Azzi, Reference Atencio and Azzi2020). An archetype of the phosphate structure of the merrillite group is β-Ca3(PO4)2 (Z = 21, R3с (#161), a ≈ 10.35Å, c ≈ 37.16 Å; Sugiyama and Tokonami, Reference Sugiyama and Tokonami1987), which in turn is a stable, low-pressure polymorph of the mineral tuite, γ-Ca3(PO4)2 (Z = 3, R $\bar{3}$m, a ≈ 5.26 Å, c ≈ 18.73 Å; Xie et al., Reference Xie, Minitti, Chen, Mao, Wang, Shu and Fei2003).

The general formula of minerals of the cerite supergroup is as follows: A 9XM[TO3Ø]7W 3, where A = Ce, La, Ca, Sr, (Na), (other REE); X = □ [vacancy], Ca and Na; M = Mg, Fe2+, Fe3+, Al and Mn; T = Si and P; Ø = O and OH; W = □, OH and F (Atencio and Azzi, Reference Atencio and Azzi2020). The known minerals of the merrillite group have simpler formulae as the W site is absent: A 9XM(TO4)3(TO3Ø)4. β-Ca3(PO4)2 is not known in Nature, and its formula as a potential member of the merrillite group can be written as Ca9(Ca0.50.5)Ca(PO4)7 (Z = 6). Deynekoite, Ca9□Fe3+(PO4)7, is the first mineral of the merrillite group with Fe3+ at the M octahedral site and a vacant X site (Table 1). Deynekoite has unit cell parameters close to the parameters of other minerals of the merrillite subgroup (Supplementary Table S2). The refractive indexes and birefringence for deynekoite are higher than for other known minerals of the merrillite subgroup (Supplementary Table S2).

The recently approved classification of the cerite supergroup was intended to solve a problem with the unification of mineral formulas of the merrillite group (see the IMA formula in Table 1). The authors of the new classification (Atencio and Azzi, Reference Atencio and Azzi2020) propose the use of generalised formulas employing the sign “#” to replace minor chemical elements that balance the charge. For example, merrillite would have the formula (Ca,#)9(Na,#)(Mg,#)(PO4)7, keplerite (Ca,#)9(Ca,#)(Mg,#)(PO4)7 and hedegaardite (Ca,#)9(Ca,#)(Mg,#)[PO3(OH)](PO4)6. Using the approach of Atencio and Azzi (2022), deynekoite would be described by the formula (Ca,#)9(□,#)(Fe3+,#)(PO4)7. Apparently, this approach does not include end-member formulas with a double site occupation at the one site (keplerite and matyhite, Table 1). Thus, the principles of mineral systematics (Bosi et al., Reference Bosi, Biagioni and Oberti2019a, Reference Bosi, Hatert, Hålenius, Pasero, Miyawaki and Mills2019b) based on the definition of the end-member formula may not be fulfilled.

Furthermore, with the nomenclature proposed by Atencio and Azzi (Reference Atencio and Azzi2020), keplerite, (Ca,#)9(Ca,#)(Mg,#)(PO4)7, and hedegaardite, (Ca,#)9(Ca,#)(Mg,#)[PO3(OH)] (PO4)6, are distinguished by only one hydrogen atom (~0.85% wt. H2O). As a result, these formulas will not help to identify these mineral species on the basis of routine microprobe analyses. However, using the specific formulas of the end-members hedegaardite and keplerite (Table 1) makes the identification of minerals of the cerite supergroup relatively simply. We believe that the classification of the cerite supergroup may be improved by taking these issues into account.

Structural aspects

The crystal structure of the merrillite-group minerals is formed by a framework of cation-centred polyhedra linked by common corners and edges. There are PO4 tetrahedra, MO6 octahedra, AO8 polyhedra and XO6 distorted trigonal prisms. The merrillite type structure is usually considered a combination of columns of different types along the Z axis (Britvin et al., Reference Britvin, Galuskina, Vlasenko, Vereshchagin, Bocharov, Krzhizhanovskaya, Shilovskikh, Galuskin, Vapnik and Obolonskaya2021b). It can also be described as an interlayering of corrugated layers (modules) of two types, nested within each other (Fig. 4, 5). This interpretation was applied, for example, in the description of the structure of Ca9Y(VO4)7 (Lazoryak et al., Reference Lazoryak, Deyneko, Aksenov, Stefanovich, Fortalnova, Petrova, Baryshnikova, Kosmyna and Shekhovtsov2018).

The first type of layer (type B, Fig. 4) is presented by corrugated six-fold rings formed by the edge sharing ACaO8 polyhedra. In the inside ring in the merrillite layer there is a truncated XNaO6 prism, sharing common edges with Ca polyhedra (Fig. 5a). The prism also shares a top face with P 1APO4 tetrahedra and bottom corners with three other phosphate groups, which are linked to the Ca ring apically (Fig. 5c). In deynekoite, an Na(X) site has low occupation (0.1 apfu) and inside the ring there is a virtual trigonal bipyramid (Fig. 5a) which is formed by a disordered PO4 tetrahedron oriented either down (Р 1A4) or up (Р 1B4) in the ratio 0.73:0.27 (Fig. 5d,e). As a rule, in minerals of the merrillite subgroup the Р 1A4 tetrahedron is occupied (at the top of the truncate Na-prism), whereas in the whitlockite-subgroup minerals, the Р1B4 tetrahedron is practically fully occupied (at the place of the missing Na-site) (Figs 5c, 6). In a synthetic phase, which is very close in composition to natural deynekoite (Deyneko et al., Reference Deyneko, Aksenov, Morozov, Stefanovich, Dimitrova, Barishnikova and Lazoryak2014), the tetrahedron-up Р 1B4 is fully occupied as in whitlockite (Figs. 5e, 6). However, in whitlockite the tetrahedron-down Р 1A4 can have a low phosphorus content (Fig. 6d). The second type of structural layer (А type, Fig. 4) repeats without changes in all minerals of the merrillite group and features M(PO4)6 clusters. The clusters are linked to each other by Ca-polyhedra (CaO8) triplets (Fig. 5b).

Figure 6. Comparison of deynekoite [(a), Na site lower occupied is not shown] and its synthetic analogue [(b), crystal structure data from Deyneko et al., Reference Deyneko, Aksenov, Morozov, Stefanovich, Dimitrova, Barishnikova and Lazoryak2014], merrillite [(c), crystal structure data from Xie et al., Reference Xie, Yang, Gu and Downs2015] and whitlockite [(d), crystal structure data from Hughes et al., 2008], projection on (010). Са – light-orange balls; P2PO4 tetrahedron – dark blue; P3PO4 tetrahedron – blue; P 1APO4 tetrahedron – dark-green; P 1BPO4 tetrahedron – light-green; truncate prism XO6 – yellow; МО6 octahedron – brown (Fe)/light-green (Mg); hydrogen – pink balls. The numerals show tetrahedral site occupation in percent.

Deynekoite is the first mineral of the merrillite group with a formally vacant X (Na) site (Na pfu ≤ 0.1). Natural deynekoite, ~Ca9(□0.9Na0.1)[Fe3+0.58(Mg,Al)0.42)(P O4)6.73(PO3OH)0.27, has a synthetic analogue with close composition – Ca9(Fe3+0.63Mg0.37)(PO4)6.63(PO3OH)0.37 (Deyneko et al., Reference Deyneko, Aksenov, Morozov, Stefanovich, Dimitrova, Barishnikova and Lazoryak2014). In the synthetic analogue the tetrahedron-up Р 1B4 with apical oxygen, which is partially (0.37) protonated, is fully occupied (Fig. 6b). In deynekoite the tetrahedron-down Р 4 has 0.73 occupation, vs. 0.27 for the tetrahedron-up Р 1B4 (Fig. 6a). The apical oxygen O10BO of Р 1B4 forms an OH group, where H10BH is disordered between three symmetrically equivalent positions. The distance P 1BP–O10BO = 1.64 Å is increased compared to the distance P 1BP–O10BO = 1.53 Å (Table 6) and close to the distance P 1BP–O10BO = 1.614 Å in whitlockite (Hughes et al., Reference Hughes, Jolliff and Rakovan2008).

The iron valence in deynekoite was not determined by the direct method because of the limited amount of material and very small grain size. The yellow-to-brown colour of deynekoite, totals from microprobe analyses close to 100% and the significant decrease of the mean M–O = 2.041 Å in comparison with other minerals of the merrillite group (which have the distance M–O = 2.08 Å) all indicate that iron in deynekoite is trivalent. This is confirmed by the bond-valence sum (BVS) calculation (Table 5), which shows that the trivalent cation predominates at the M site.

The differences in the occupation of the Р 1(PО4)3– sites in deynekoite and synthetic analogues can be explained by the peculiar features of deynekoite's origin. Initially, in high-temperature and reduced conditions, ferromerrillite, NaCa9(Fe2+,Mg)(PO4)7, forming a solid solution with keplerite, Ca0.5Ca9(Fe2+,Mg)(PO4)7, crystallises. As the temperature decreases and oxygen activity increases, ferromerrillite is replaced by deynekoite with the removal of Na from the structure. Deynekoite inherits the characteristics of the ferromerrillite structure – the P PO4 tetrahedron preserves significant occupation, though a small part of the phosphorus changes its position and occupies the P1BPO4 site, accompanied by the protonation of the O10BO apical oxygen.

Genesis

Diopside–anorthite–tridymite paralava, on the contact of which deynekoite was found, was generated at the near-surface combustion focus. This was a continuous and relatively long-lasting combustion phenomenon that determined the anomalous high temperature of the generated melt (~1500°С). One indicator of the super-high combustion temperature is cristobalite with a fish-scale or ballen structure (Fig. 1c), which reflects volume shrinkage upon the inversion of β-cristobalite to α-cristobalite (Schmieder et al., Reference Schmieder, Buchner and Kröchert2009).

The genesis of phosphides and phosphates on the boundary of the diopside–anorthite–tridymite paralava was connected with the interaction of the paralava and country rocks, which were a source of reductant (altered bitumen) and also phosphorus and iron. These probably came from the phosphorite-bearing rocks of the Muwaqqar formation, the so-called Р–Fe–C (± Si) series analogous to the rocks of the lower part of the Ghareb formation, Israel (Shahar et al., Reference Shahar, Yaacov and Yair1989). A find of fluorapatite pseudomorphs after fish bones in the paralava confirms the presence of phosphorite inclusions in the carbonate protolith (Fig. 1d). Accessory fluorapatite in the paralava was stable on the contact of paralava with the country rocks and cannot be a source of phosphorus. The genesis of phosphides is related to carbothermal reductive reactions on the boundary of paralava and country rocks. These reactions were considered by us in the frame of the phosphide genesis model defined for the contact of gehlenite paralava and fragments of country rocks in breccia of the Hatrurim Basin in the Negev Desert, Israel: Fe2O3 + 3C → 2Fe(lq) + 3CO(g); 8Ca5(PO4)3F [fluorapatite from phosphorite with (OH) and (CO3)2– impurities] + 22SiO2 + 20Al2O3 + 60C = 12P2(g) + 60CO(g) + 20Ca2Al2O7 + 2SiF4(g); nFe(lq) + ½P2(g) = FenP, n = 1,2,3 (Galuskin et al., Reference Galuskin, Galuskina, Kamenetsky, Vapnik, Kusz and Zieliński2022a, Reference Galuskin, Kusz, Galuskina, Książek, Vapnik and Zieliński2023b). The crystallisation of phosphides took place in the context of the increasing activity of P that is reflected in the crystallisation sequence of Fe2P (barringerite) – FeP (murashkoite) – FeP2 (zuktamrurite) (Britvin et al., Reference Britvin, Murashko, Vapnik, Polekhovsky, Krivovichev, Vereshchagin, Vlasenko, Shilovskikh and Zaitsev2019a, Reference Britvin, Vapnik, Polekhovsky, Krivovichev, Krzhizhanovskaya, Gorelova, Vereshchagin, Shilovskikh and Zaitsev2019b). The temperature of phosphide formation from a melt was higher than 1300°C (the temperature of barringerite crystallisation is 1350°C), which is similar to the temperature at which troilite (pyrrhotite) crystallises from sulfide melt. With the further decrease in temperature and increasing influence of atmospheric oxygen (processes occurring at near-surface conditions), phosphides were replaced by Fe2+-bearing phosphates, which were changed by association of Fe3+-bearing phosphates. Our investigations of phosphide-bearing rocks from Jordan and Israel show that other minerals of the merrillite group (merrillite, keplerite, ferromerrillite and possibly matyhite), also appear in them. These minerals contain Fe2+ and crystallise from the locally generated phosphate melt at high temperatures (~1200°C). Deynekoite, which contains Fe3+ (substituting Fe2+-phosphates) and a small amount of water (see below), was formed at lower temperatures (600–800°C). Most likely, when the early associations were replaced by later ones, metasomatic mechanisms of the replacement of minerals in the quasi-solid state and with the participation of intergranular melt and active gases were realised.

Acknowledgements

The authors thank Peter Leverett and two anonymous reviewers for their useful and constructive comments. Investigations were supported by the National Science Center of Poland Grant (grant number 2021/41/B/ST10/00130).

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1180/mgm.2023.71.

Competing interests

The authors declare none.

Footnotes

Associate Editor: Daniel Atencio

References

Atencio, D. and Azzi, A. (2020) Cerite: A new supergroup of minerals and cerite-(La) renamed ferricerite-(La). Mineralogical Magazine, 84, 928931.10.1180/mgm.2020.86CrossRefGoogle Scholar
Benarafaa, A., Kacimia, M., Gharbagea, S., Millet, J.-M. and Ziyada, M. (2000) Structural and spectroscopic properties of calcium-iron Ca9Fe(PO4)7 phosphate. Materials Research Bulletin, 35, 20472055.10.1016/S0025-5408(00)00403-7CrossRefGoogle Scholar
Bentor, Y.K. (editor) (1960) Israel. In: Lexique Stratigraphique International, Asie, Vol. III, (10.2). Centre national de la recherche scientifique, Paris.Google Scholar
Bosi, F., Biagioni, C. and Oberti, R. (2019a) On the chemical identification and classification of minerals. Minerals, 9, 591.10.3390/min9100591CrossRefGoogle Scholar
Bosi, F., Hatert, F., Hålenius, U., Pasero, M., Miyawaki, R. and Mills, S. (2019b) On the application of the IMA−CNMNC dominant-valency rule to complex mineral compositions. Mineralogical Magazine, 83, 627632.10.1180/mgm.2019.55CrossRefGoogle Scholar
Britvin, S.N., Pakhomovskii, Y.A., Bogdanova, A.N. and Skiba, V.I. (1991) Strontiowhitlockite, Sr9Mg(PO3OH)(PO4)6, a new mineral species from the Kovdor deposit, Kola Peninsula, U.S.S.R. The Canadian Mineralogist, 29, 8793.Google Scholar
Britvin, S.N., Krivovichev, S.V. and Armbruster, T. (2016) Ferromerrillite, Ca9NaFe2+(PO4)7, a new mineral from the Martian meteorites, and some insights into merrillite-tuite transformation in shergottites. European Journal of Mineralogy, 28, 125136.10.1127/ejm/2015/0027-2508CrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Vapnik, Ye., Polekhovsky, Yu.S., Krivovichev, S.V., Vereshchagin, O.S., Vlasenko, N.S., Shilovskikh, V.V. and Zaitsev, A.N. (2019a) Zuktamrurite, FeP2, a new mineral, the phosphide analogue of löllingite, FeAs2. Physics and Chemistry of Minerals, 46, 361369.10.1007/s00269-018-1008-4CrossRefGoogle Scholar
Britvin, S.N., Vapnik, Ye., Polekhovsky, Yu.S., Krivovichev, S.V., Krzhizhanovskaya, M.G., Gorelova, L.A., Vereshchagin, O.S., Shilovskikh, V.V. and Zaitsev, AN (2019b) Murashkoite, FeP, a new terrestrial phosphide from pyrometamorphic rocks of the Hatrurim Formation, Southern Levant. Mineralogy and Petrology, 113, 237248.10.1007/s00710-018-0647-yCrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Vapnik, Ye., Polekhovsky, Yu.S., Krivovichev, S.V., Krzhizhanovskaya, M.G., Vereshchagin, O.S., Shilovskikh, V.V. and Vlasenko, N.S. (2020) Transjordanite, Ni2P, a new terrestrial and meteoritic phosphide, and natural solid solutions barringerite–transjordanite (hexagonal Fe2P–Ni2P). American Mineralogist, 105, 428436.10.2138/am-2020-7275CrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Krzhizhanovskaya, M.G., Vapnik, Ye., Vlasenko, N.S., Vereshchagin, O.S., Pankin, D.V. and Vasiliev, E.A. (2021a) Moabite, IMA 2020–092. CNMNC Newsletter 60; Mineralogical Magazine, 85, 454458.Google Scholar
Britvin, S.N., Galuskina, I.O., Vlasenko, N.S., Vereshchagin, O.S., Bocharov, V.N., Krzhizhanovskaya, M.G., Shilovskikh, V.V., Galuskin, E.V., Vapnik, Ye. and Obolonskaya, E.V. (2021b) Keplerite, Ca9(Ca0.50.5)Mg(PO4)7, a new meteoritic and terrestrial phosphate isomorphous with merrillite, Ca9NaMg(PO4)7. American Mineralogist, 106, 19171927.10.2138/am-2021-7834CrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Vapnik, Ye., Vlasenko, N.S., Vereshchagin, O.S., Krzhizhanovskaya, M.G. and Bocharov, V.N. (2021c) Nabateaite, IMA 2021–026. CNMNC Newsletter 62; Mineralogical Magazine, 85, 634638.Google Scholar
Britvin, S.N., Murashko, M.N., Vapnik, Ye., Zaitsev, A.N., Shilovskikh, V.V., Krzhizhanovskaya, M.G., Gorelova, L.A., Vereshchagin, O.S., Vasilev, E.A. and Vlasenko, N.S. (2022a) Orishchinite, a new terrestrial phosphide, the Ni-dominant analogue of allabogdanite. Mineralogy and Petrology, 116, 369378.10.1007/s00710-022-00787-xCrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Krzhizhanovskaya, M.G., Vereshchagin, O.S., Vlasenko, N.S., Vapnik, Ye. and Bocharov, V.N. (2022b) Nazarchukite, IMA 2022-005, in: CNMNC Newsletter 67. Mineralogical Magazine, 86, 849853.Google Scholar
Britvin, S.N., Murashko, M.N., Krzhizhanovskaya, M.G., Vlasenko, N.S., Vereshchagin, O.S., Vapnik, Ye. and Bocharov, V.N. (2023a) Crocobelonite, CaFe3+2(PO4)2O, a new oxyphosphate mineral, the product of pyrolytic oxidation of natural phosphides. American Mineralogist, 108, 19731983.10.2138/am-2022-8757CrossRefGoogle Scholar
Britvin, S.N., Murashko, M.N., Krzhizhanovskaya, M.G., Vapnik, Ye., Vlasenko, N.S., Vereshchagin, O.S., Pankin, D.V., Zaitsev, A.N. and Zolotarev, A.A. (2023b) Yakubovichite, CaNi2Fe3+(PO4)3, a new nickel phosphate mineral of non-meteoritic origin. American Mineralogist, 108, 21422150.10.2138/am-2022-8800CrossRefGoogle Scholar
Cooper, M.A., Hawthorne, F.C., Abdu, Y., Ball, N.A., Ramik, R.A. and Tait, K.T. (2013) Wopmayite, ideally Ca6Na3Mn(PO4)3(PO3OH)4, a new phosphate mineral from the Tanco Mine, Bernic Lake, Manitoba. The Canadian Mineralogist, 51, 93106.10.3749/canmin.51.1.93CrossRefGoogle Scholar
Deyneko, D.V., Aksenov, S.M., Morozov, V.A., Stefanovich, S.Yu., Dimitrova, O.V., Barishnikova, O.V. and Lazoryak, B.I. (2014) A new hydrogen-containing whitlockite-type phosphate Ca9(Fe0.63Mg0.37)H0.37(PO4)7: hydrothermal synthesis and structure. Zeitschrift für Kristallographie, 229, 823830.10.1515/zkri-2014-1774CrossRefGoogle Scholar
Dolomanov, O.V., Bourhis, L., Gildea, R.J., Howard, J.A.K. and Puschmann, H. (2009) OLEX2: a complete structure solution, refinement and analysis program. Journal of Applied Crystallography, 42, 339341.10.1107/S0021889808042726CrossRefGoogle Scholar
Frondel, C. (1943) Mineralogy of the calcium phosphates in insular phosphate rock. American Mineralogist, 28, 215232.Google Scholar
Galuskin, E.V., Gfeller, F., Galuskina, I.O., Pakhomova, A., Armbruster, T., Vapnik, Y., Włodyka, R., Dzierżanowski, P. and Murashko, M. (2015) New minerals with a modular structure derived from hatrurite from the pyrometamorphic Hatrurim Complex. Part II. Zadovite, BaCa6[(SiO4)(PO4)](PO4)2F and aradite, BaCa6[(SiO4)(VO4)](VO4)2F, from paralavas of the Hatrurim Basin, Negev Desert, Israel. Mineralogical Magazine, 9, 10731087.10.1180/minmag.2015.079.5.04CrossRefGoogle Scholar
Galuskin, E.V., Galuskina, I.O., Gfeller, F., Krüger, B., Kusz, J., Vapnik, Ye., Dulski, M. and Dzierżanowski, P. (2016) Silicocarnotite, Ca5[(SiO4)(PO4)](PO4), a new ‘old’ mineral from the Negev Desert, Israel, and the ternesite–silicocarnotite solid solution: indicators of high-temperature alteration of pyrometamorphic rocks of the Hatrurim Complex, Southern Levant. European Journal of Mineralogy, 28, 105123.10.1127/ejm/2015/0027-2494CrossRefGoogle Scholar
Galuskin, E.V., Krüger, B., Galuskina, I.O., Krüger, H., Vapnik, Ye., Pauluhn, A. and Olieric, V. (2019) Levantite, KCa3(Al2Si3)O11(PO4), a new latiumite-group mineral from the pyrometamorphic rocks of the Hatrurim Basin, Negev Desert, Israel. Mineralogical Magazine, 83, 713721.10.1180/mgm.2019.37CrossRefGoogle Scholar
Galuskin, E., Galuskina, I., Krüger, B., Krüger, H., Vapnik, Ye., Krzątała, A., Środek, D. and Zieliński, G. (2021) Nomenclature and classification of the arctite supergroup. Aravaite, Ba2Ca18(SiO4)6[(PO4)3(CO3)]F3O, a new arctite supergroup mineral from Negev Desert, Israel. The Canadian Mineralogist, 59, 191209.10.3749/canmin.2000035CrossRefGoogle Scholar
Galuskin, E.V., Galuskina, I.O., Kamenetsky, V., Vapnik, Ye., Kusz, J. and Zieliński, G. (2022a) First in-situ terrestrial osbornite (TiN) in the pyrometamorphic Hatrurim Complex, Israel. Lithosphere, 81277447.10.2113/2022/8127747CrossRefGoogle Scholar
Galuskin, E., Stachowicz, M., Galuskina, I.O., Woźniak, K., Vapnik, Y., Murashko, N.N. and Zieliński, G. (2022b) Deynekoite, IMA 2021-108. CNMNC Newsletter 66. Mineralogical Magazine, 86, https://doi.org/10.1180/mgm.2022.33Google Scholar
Galuskin, E.V., Galuskina, I.O., Vapnik, Y. and Zieliński, G. (2023a) Discovery of “meteoritic” layered disulphides ACrS2 (A = Na, Cu, Ag) in terrestrial rock. Minerals, 13, 381.10.3390/min13030381CrossRefGoogle Scholar
Galuskin, E.V., Kusz, J., Galuskina, I.O., Książek, M., Vapnik, Ye. and Zieliński, G. (2023b) Discovery of terrestrial andreyivanovite, FeCrP, and the effect of Cr and V substitution in barringerite-allabogdanite low-pressure transition. American Mineralogist, 108, 15061515.10.2138/am-2022-8647CrossRefGoogle Scholar
Galuskina, I.O., Galuskin, E.V. and Vapnik, Y.A. (2016) Terrestrial merrillite. Plinius, 42, 563 [2nd European Mineralogical Conference. Abstract].Google Scholar
Galuskina, I.O., Galuskin, E.V., Pakhomova, A.S., Widmer, R., Armbruster, T., Krüger, B., Grew, E.S., Vapnik, Y., Dzierażanowski, P. and Murashko, M. (2017a) Khesinite, Ca4Mg2Fe3+10O4 [(Fe3+10Si2)O36], a new rhönite-group (sapphirine supergroup) mineral from the Negev Desert, Israel—Natural analogue of the SFCA phase. European Journal of Mineralogy, 29, 101116.10.1127/ejm/2017/0029-2589CrossRefGoogle Scholar
Galuskina, I.O., Galuskin, E.V., Prusik, K., Vapnik, Y., Juroszek, R., Jeżak, L. and Murashko, M. (2017b) Dzierżanowskite, CaCu2S2 – a new natural thiocuprate from Jabel Harmun, Judean Desert, Palestine Autonomy, Israel. Mineralogical Magazine, 81, 10731085.10.1180/minmag.2016.080.153CrossRefGoogle Scholar
Galuskina, I.O., Galuskin, E.V., Vapnik, Y., Prusik, K., Stasiak, M., Dzierżanowski, P. and Murashko, M. (2017c) Gurimite, Ba3(VO4)2 and hexacelsian, BaAl2Si2O8 - Two new minerals from schorlomite-rich paralava of the Hatrurim Complex, Negev Desert, Israel. Mineralogical Magazine, 81, 10091019.10.1180/minmag.2016.080.147CrossRefGoogle Scholar
Geller, Y.I., Burg, A., Halicz, L. and Kolodny, Y. (2012) System closure during the combustion metamorphic Mottled Zone event, Israel. Chemical Geology, 334, 2536.10.1016/j.chemgeo.2012.09.029CrossRefGoogle Scholar
Gross, S. (1977) The mineralogy of the Hatrurim Formation, Israel. Geological Survey of Israel Bulletin, 70, 180.Google Scholar
Gross, S. (1984) Occurrence of ye'elimite and ellestadite in an unusual cobble from the „pseudo-conglomerate” of the Hatrurim Basin. Geological Survey of Israel Bulletin, 84, 14.Google Scholar
Hughes, J.M., Jolliff, B.L. and Rakovan, J. (2008) The crystal chemistry of whitlockite and merrillite and the dehydrogenation of whitlockite to merrillite. American Mineralogist, 93, 13001305.10.2138/am.2008.2683CrossRefGoogle Scholar
Hwang, S.L., Shen, P., Chu, H.T., Yui, T.F., Varela, M.E. and Iizuka, Y. (2019) New minerals tsangpoite Ca5(PO4)2(SiO4) and matyhite Ca9(Ca0.50.5)Fe(PO4)7 from the D'Orbigny angrite. Mineralogical Magazine, 83, 293313.10.1180/mgm.2018.125CrossRefGoogle Scholar
Jolliff, B.L., Hughes, J.M., Freeman, J.J. and Zeigler, R.A. (2006) Crystal chemistry of lunar merrillite and comparison to other meteoritic and planetary suites of whitlockite and merrillite. American Mineralogist, 91, 15831595.10.2138/am.2006.2185CrossRefGoogle Scholar
Khoury, H.N., Sokol, E.V., Kokh, S.N., Seryotkin, Y.V., Nigmatulina, E.N., Goryainov, S.V., Belogub, E.V. and Clark, I.D. (2016) Tululite, Ca14(Fe3+,Al)(Al,Zn,Fe3+,Si,P,Mn, Mg)15O36: a new Ca zincate-aluminate from combustion metamorphic marbles, Central Jordan. Mineralogy and Petrology, 110, 125140.10.1007/s00710-015-0413-3CrossRefGoogle Scholar
Kolodny, Y. and Gross, S. (1974) Thermal metamorphism by combustion of organic matter: isotopic and petrological evidence. Journal of Geology, 82, 489506.10.1086/627995CrossRefGoogle Scholar
Krüger, B., Galuskin, E.V., Galuskina, I.O., Krüger, H. and Vapnik, Ye. (2021) Kahlenbergite KAl11O17, a new – alumina mineral and Fe-rich hibonite from the Hatrurim Basin, the Negev desert, Israel. European Journal of Mineralogy, 33, 341355.10.5194/ejm-33-341-2021CrossRefGoogle Scholar
Krzątała, A., Krüger, B., Galuskina, I., Vapnik, Y. and Galuskin, E. (2020) Walstromite, BaCa2(Si3O9), from Rankinite Paralava within Gehlenite Hornfels of the Hatrurim Basin, Negev Desert, Israel. Minerals, 10, 407.10.3390/min10050407CrossRefGoogle Scholar
Lazoryak, B.I., Morozov, V.A., Belik, A.A., Khasanov, S.S. and Shekhtman, V.Ch. (1996) Crystal structures and characterization of Ca9Fe(PO4)7 and Ca9FeH0.9(PO4)7. Journal of Solid State Chemistry, 122, 1521.10.1006/jssc.1996.0074CrossRefGoogle Scholar
Lazoryak, B.I., Deyneko, D.V., Aksenov, S.M., Stefanovich, S.Yu., Fortalnova, E.A., Petrova, D.A., Baryshnikova, O.V., Kosmyna, M.B. and Shekhovtsov, A.N. (2018) Pure, lithium- or magnesium-doped ferroelectric single crystals of Ca9Y(VO4)7: cation arrangements and phase transitions. Zeitschrift für Kristallographie, 233, 111.Google Scholar
Li, T., Li, Z., Huang, Z., Zhong, J., Fan, G., Guo, D., Qin, M., Zhang, J., Li, J., Liu, H., Qiu, L., Wang, F., He, S., Yu, A., Liu, R., Wu, Y., Deng, L., Tai, Z., He, Y. and Lin, Y. (2022) Changesite-(Y), IMA 2022–023. CNMNC Newsletter 69; Mineralogical Magazine, 86, 988992.Google Scholar
Matthews, A. and Gross, S. (1980) Petrologic evolution of the Mottled Zone (Hatrurim) metamorphic complex of Israel. Israel Journal of Earth Sciences, 29, 93106.Google Scholar
Murashko, M.N., Britvin, S.N., Vapnik, Ye., Polekhovsky, Y.S., Shilovskikh, V.V., Zaitsev, A.N. and Vereshchagin, O.S. (2022) Nickolayite, FeMoP, a new natural molybdenum phosphide. Mineralogical Magazine, 86, 749757.10.1180/mgm.2022.52CrossRefGoogle Scholar
Novikov, I., Vapnik, Ye. and Safonova, I (2013) Mud volcano origin of the Mottled Zone, South Levant. Geoscience Frontiers, 4, 597619.10.1016/j.gsf.2013.02.005CrossRefGoogle Scholar
Rigaku Oxford Diffraction, (2019) CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.Google Scholar
Schmieder, M., Buchner, E. and Kröchert, J. (2009) “Ballen silica” in impactites and magmatic rocks. Abstracts 40th Lunar and Planetary Science Conference. Texas, USA.Google Scholar
Shahar, Y., Yaacov, N. and Yair, S (1989) Two close associations: Si-P-Fe and Si-P-C in the Upper Campanian and Lower Maastrichtian sediments of the Negev, Israel. Sciences Géologiques, bulletins et mémoires, 42, 155171.10.3406/sgeol.1989.1820CrossRefGoogle Scholar
Sheldrick, G.M. (2015a) SHELXT – Integrated space-group and crystal-structure determination. Acta Crystallographica, A71, 38.Google Scholar
Sheldrick, G.M. (2015b) Crystal structure refinement with SHELXL. Acta Crystallographica, C71, 38.Google Scholar
Sokol, E., Novikov, I., Zateeva, S., Vapnik, Ye., Shagam, R. and Kozmenko, O. (2010) Combustion metamorphism in the Nabi Musa dome: New implications for a mud volcanic origin of the Mottled Zone, Dead Sea area. Basin Research, 22, 414438.10.1111/j.1365-2117.2010.00462.xCrossRefGoogle Scholar
Solecka, U., Bajda, T., Topolska, J., Zelek-Pogudz, S. and Manecki, M. (2018) Raman and Fourier transform infrared spectroscopic study of pyromorphite-vanadinite solid solutions. Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy, 190, 96103.10.1016/j.saa.2017.08.061CrossRefGoogle ScholarPubMed
Sugiyama, K. and Tokonami, M. (1987) Structure and crystal chemistry of a dense polymorph of tricalcium phosphate Ca3(PO4)2: a host to accommodate large lithophile elements in the Earth's mantle. Physics and Chemistry of Minerals, 15, 12513010.1007/BF00308774CrossRefGoogle Scholar
Vapnik, Y., Sharygin, V.V., Sokol, E.V. and Shagam, R. (2007) Paralavas in a combustion metamorphic complex: Hatrurim Basin, Israel. Reviews in Engineering Geology, 18, 121.Google Scholar
Ward, D., Bischoff, A., Roszjar, J., Berndt, J. and Whitehouse, M.J. (2017) Trace element inventory of meteoritic Ca-phosphates. American Mineralogist, 102, 18561880.10.2138/am-2017-6056CrossRefGoogle Scholar
Wherry, E.T. (1917) Merrillite, meteoritic calcium phosphate. American Mineralogist, 2, 119119.Google Scholar
Witzke, T., Phillips, B.L., Woerner, W., Countinho, J.M.V., Färber, G. and Contreira Filho, R.R. (2015) Hedegaardite, IMA 2014-069. CNMNC Newsletter No. 23, February 2015, page 54. Mineralogical Magazine, 79, 5158.Google Scholar
Xie, X., Minitti, M.E., Chen, M., Mao, H.-K., Wang, D., Shu, J. and Fei, Y. (2003) Tuite, γ-Ca3(PO4)2 – A new mineral from the Suizhou L6 chondrite. European Journal of Mineralogy, 15, 10011005.10.1127/0935-1221/2003/0015-1001CrossRefGoogle Scholar
Xie, X., Yang, H., Gu, X., Downs, R.T. (2015) Chemical composition and crystal structure of merrillite from the Suizhou meteorite. American Mineralogist, 100, 27532756.10.2138/am-2015-5488CrossRefGoogle Scholar
Figure 0

Table 1. Minerals of the merrillite group.

Figure 1

Table 2. Crystal data and structure refinement details for deynekoite.

Figure 2

Figure 1. (a) Heterogeneous, porous basalt-like rock from a central part of the paralava body. (b) Paralava consisting of diopside, tridymite, anorthite and wollastonite. Back-scattered electron (BSE) image. (c) Cristobalite with fish-scale texture fissures overgrowing tridymite. BSE image. (d) Pseudomorph of fluorapatite after fish bone in the paralava. BSE image. An = anorthite, Crs = cristobalite, Di = diopside, Fap = fluorapatite, Hd = hedenbergite, Gls = glass, Mgt = magnetite, Trd = tridymite, Tnt = titanite, Wol = wollastonite.

Figure 3

Figure 2. (a) Contact of diopside paralava (I) and altered marl (II), black is a hematite-bearing zone. The white frame shows a fragment magnified in Fig. 2b. (b) Position of phosphate mineralisation at the contact zone of rocks: 1 – diopside paralava, 2 – zone enriched with phosphides (murashkoite, zuktamrurite), 3 – phosphate zone, 4 – hematite zone (developed after pyrrhotite), 5 – altered host rock (marl); diopside chondrules are shown by arrows. (c) Phosphate zone represented, mainly, by deynekoite. Diopside paralava is replaced intensively by calcite and gismondine-Ca; in unaltered diopside crystals, barringerite inclusions are noted. The white frame shows a fragment magnified in Fig. 2d; (d) One of the rare places where deynekoite forms almost monomineralic aggregates of elongated grains up to 50 μm in size. Bgr = barringerite, Dnk = deynekoite, Di = diopside, Fap = fluorapatite, Hem = hematite, Gsm = gismondine-Ca, Muh = murashkoite.

Figure 4

Table 3. Chemical composition of deynekoite.

Figure 5

Figure 3. Raman spectrum of deynekoite.

Figure 6

Figure 4. Deynekoite structure. (a) Projection on (010). The structure can be presented as the intercalation of two types of layers (see Fig. 5) in the sequence ABA'B’A’’B”A. (b) Projection on (001). One slice of the A layer and two slices of B layers are shown. The Na site with lower occupation is not shown. Ca atoms – light-orange; Na atoms – yellow; P2PO4 tetrahedron – dark-blue; P3PO4 tetrahedron – blue; P1APO4 tetrahedron – dark-green; P1BPO4 tetrahedron – light-green; МО6 octahedron – brown; oxygen O10B – dark blue balls; oxygen O10A – blue balls; hydrogen – small pink balls.

Figure 7

Figure 5. (a,b) Layers of two types (А, В see Fig. 4) in the structure of deynekoite, projection on (00$\bar{1}$). (a) In the B layer formed by six-membered rings building from Ca-polyhedra, three types of triangular empty occupation are shown as in the ideal structures of merrillite (1), deynekoite (2) and whitlockite (3). (b) The A layer is the same in merrillite, deynekoite and whitlockite, consisting of millwheel clusters M(P2+P3PO4)6 linked with one another by triplets of Ca-polyhedra. (c–e) Atom arrangement near the P1 site in holotype deynekoite: c – merrillite type (10%); d – deynekoite type (63%); e – whitlockite type (27%). Ca-polyhedra – light-orange; P2PO4 tetrahedron – dark-blue; P3PO4 tetrahedron – blue; P1APO4 tetrahedron – dark-green; P1BPO4 tetrahedron – light-green; truncated prism XO6 – yellow; octahedron МО6 – brown (Fe3+)/dark-orange (Mg); oxygen – blue balls; hydrogen – pink balls (hydrogen bonds are shown); empty circle – vacancy.

Figure 8

Table 4. Fractional atomic coordinates and isotropic* or equivalent isotropic displacement parameters (Å2) for deynekoite.

Figure 9

Table 5. Atomic displacement parameters (Å2) for deynekoite.

Figure 10

Table 6. Selected bond lengths (Å) and bond-valence sum (BVS) calculation for deynekoite.

Figure 11

Figure 6. Comparison of deynekoite [(a), Na site lower occupied is not shown] and its synthetic analogue [(b), crystal structure data from Deyneko et al., 2014], merrillite [(c), crystal structure data from Xie et al., 2015] and whitlockite [(d), crystal structure data from Hughes et al., 2008], projection on (010). Са – light-orange balls; P2PO4 tetrahedron – dark blue; P3PO4 tetrahedron – blue; P1APO4 tetrahedron – dark-green; P1BPO4 tetrahedron – light-green; truncate prism XO6 – yellow; МО6 octahedron – brown (Fe)/light-green (Mg); hydrogen – pink balls. The numerals show tetrahedral site occupation in percent.

Supplementary material: File

Galuskin et al. supplementary material 1

Galuskin et al. supplementary material
Download Galuskin et al. supplementary material 1(File)
File 289.4 KB
Supplementary material: File

Galuskin et al. supplementary material 2

Galuskin et al. supplementary material
Download Galuskin et al. supplementary material 2(File)
File 89.8 KB
Supplementary material: File

Galuskin et al. supplementary material 3

Galuskin et al. supplementary material
Download Galuskin et al. supplementary material 3(File)
File 25.8 KB