Next Article in Journal
Petrography, Geochemical Features and Absolute Dating of the Mesozoic Igneous Rocks of Medvedev and Taezhniy Massifs (Southeast Russia, Aldan Shield)
Next Article in Special Issue
The Crystal Structure and Crystal Chemistry of Mineral-like Cd5(VO4)2(OH)4, a Novel Isomorph of Arsenoclasite and Gatehouseite
Previous Article in Journal
Metamorphic Ages of the Jurassic Accretionary Complexes in the Kanto Mountains, Central Japan, Determined by K–Ar Dating of Illite: Implications for the Tectonic Relationship between the Chichibu and Sanbagawa Belts
Previous Article in Special Issue
Temperature-Induced Phase Transition in a Feldspar-Related Compound BaZn2As2O8∙H2O
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mixed Valanced V3+,V2+ Phosphate Na7V4(PO4)6: A Structural Analogue of Mineral Yurmarinite

1
M.V. Lomonosov Moscow State University, Leninskye Gory 1, Moscow 119991, Russia
2
The Institute of Experimental Mineralogy RAS, Akademika Osip’yana st. 4, Chernogolovka 142432, Russia
3
National Research Centre “Kurchatov Institute”, Akademika Kurchatova pl. 1, Moscow 123182, Russia
*
Authors to whom correspondence should be addressed.
Minerals 2022, 12(12), 1517; https://doi.org/10.3390/min12121517
Submission received: 1 November 2022 / Revised: 23 November 2022 / Accepted: 24 November 2022 / Published: 27 November 2022
(This article belongs to the Special Issue Mineral-Related Oxo-Salts: Synthesis and Structural Crystallography)

Abstract

:
Two sodium vanadium phosphates, synthetic analogues of the minerals kosnarite, Na3V2(PO4)3, and yurmarinite, Na7V4(PO4)6, were obtained by hydrothermal synthesis simulating a natural hydrothermal solution. While the Na3V2(PO4)3 phase belongs to the NASICON family and is well-known for its high-ionic conductivity, the new Na7V4(PO4)6 compound is a rare case of V2+-containing oxosalts, which are hard to prepare due to their instability in air. Here we report the crystal structure of heterovalent vanadium phosphate studied by single crystal X-ray diffraction, XANES spectroscopy, and topological ion migration modelling. A discussion of divalent vanadium compounds of both natural and synthetic origin is also given, with a review of the methods for their synthesis and a comparative analysis of V–O bond lengths.

1. Introduction

Due to its high chemical activity, variable valence, and ability to form different stable complexes, vanadium is an essential element of more than 200 minerals. V3+-containing minerals are mostly igneous silicates, V5+ in the form of [VO4]3− anion occurs mainly as hydrothermal vanadates, and V4+ as the (VO)2+ cation forms minerals of mixed origin, such as metamorphic, hydrothermal or weathering [1]. The genesis of V natural phases correlates with the oxidation state of vanadium, which directly depends on the pH value of the crystallization medium and the reduction potential [2]. While V3+ is generally stable under highly anoxic conditions, V4+ is present in moderately reducing environments, particularly under acidic conditions, and V5+ is stable under oxygen conditions and in sub-oxygen alkaline environments [3]. Accordingly, only the III, IV, and V valence states of vanadium are relevant under conditions typical for the earth’s surface.
Conversely, divalent vanadium is unstable in the air and rarely forms natural phases. The unique V2+-bearing mineral dellagiustaite, Al(V2+,V3+)2O4, was recently found in rocks from Sierra de Comechingones, San Luis, Argentina, and in late pyroclastic ejecta of small Cretaceous basalt volcanoes exposed at Mt Carmel, Northern Israel [4]. It crystallizes in the inverse spinel structure type and cannot be labeled as “vanadium coulsonite” because of the predominance of Al over V3+ in the tetrahedral position. Dellagiustaite is associated with hibonite, CaAl12O19, and grossite, CaAl4O7, which constitute the modal composition of the rock. At the same time, V and Al metal alloys indicate significantly reducing conditions in the assemblages from both localities. The authors propose that mineral formation results from the interaction of deep-seated magmas and CH4 ± H2 fluids in volcanic feeder systems [4].
We obtained a new sodium phosphate, Na7V4(PO4)6, with V atoms in the 2+ and 3+ oxidation states under hydrothermal conditions. It was synthesized together with the NASICON-type Na3V2(PO4)3. The new compound is isostructural to the fumarole mineral yurmarinite, Na7(Fe3+,Mg,Cu)4(AsO4)6 [5]. Presumably, yurmarinite was formed under strongly oxidizing conditions of the Arsenatnaya fumarole in association with several alluaudite-type arsenates, namely hatertite, Na2(Ca,Na)(Fe3+,Cu)2(AsO4)3, bradaczekite, NaCu4(AsO4)3, and johillerite, NaMg3Cu(AsO4)3. The high volatility of O in fumarolic gases causes high-valency states of As5+ and Fe3+ in all phases [5]. In octahedral positions of yurmarinite crystal structure, Mg2+ and Cu2+ cations isomorphically substitute Fe3+. Moreover, the yurmarinite structure-type unites synthetic phosphates [6], arsenates [7,8,9], and even molybdates [10] with the general formula (X)6(M1)(M2)(M3)3(TO4)6 [11]. Most of them contain Na in the X and M1 positions but Fe2+ and Fe3+ inside the M2 and M3 octahedra. The preparation of heterovalent iron compounds is standard. However, the oxosalts containing V2+ are exotic because of their complicated synthesis and low stability. As part of our program of studying mineralogically probable V-bearing phases [12,13,14,15], we present here the results of hydrothermal synthesis, crystal structure, and comparative crystal chemistry of a new sodium and vanadium (II, III) orthophosphate Na7V4(PO4)6.

2. Materials and Methods

2.1. Hydrothermal Synthesis and Scanning Electron Microscopy

High-temperature hydrothermal synthesis was carried out in a phosphate system with sodium and vanadium cations. Chemically pure primary reagents V2O3 2 g (13.3 mmol) and NaCl 2 g (34.2 mmol) were ground and mixed as powders in the presence of a small amount of 0.3 g Na3C6H5O7 (1.2 mmol) taken as redox agent. All components were dissolved in 7 mL of 1M aqueous solutions of H3PO4 and placed in a copper-lined nickel–chromium alloy autoclave of 14 mL capacity. The synthesis was performed at a temperature of 723 K and a pressure of 50 MPa. The synthesis duration was 20 days, sufficient to complete the crystallization reactions. The product was cooled to room temperature for over 24 h and washed with hot distilled water. As a result, well-faceted crystals of two phases were obtained: black flattened crystals with rhombohedral habitus and a vitreous luster (Figure 1) and yellow crystals of cubic shape with a strong luster.
X-ray microprobe spectral analysis was performed at a Jeol JSM-6480LV SEM with an energy-dispersive spectrometer Oxford X-ManN (Laboratory of Analytical Techniques of High Spatial Resolution, Dept. of Petrology, Moscow State University, Russia). EDS analysis revealed that both crystalline phases include V, Na, P, and O atoms. After a preliminary X-ray experiment, yellow isometric crystals were identified as Na3V2(PO4)3 with the NASICON structure type [16]. The second black phase of the rhombohedral habitus was diagnosed as a new compound and became the subject of X-ray diffraction analysis and crystal–chemical discussion.

2.2. X-ray Diffraction and X-ray Absorption Spectroscopy

Low-temperature (T = 110 K) X-ray diffraction experimental data were collected on a Bruker D 8 Quest single-crystal X-ray diffractometer, (Bruker Corporation, Billerica, MA, USA) (MoKα radiation). The dataset was corrected for background, Lorentz and polarization effects, and absorption. All calculations were performed within the WinGX program system, Version 2021.3 [17]. The trigonal crystal structure of the new phase was solved by direct methods in the rhombohedral space group R 3 ¯ c and refined in the anisotropic approximation of thermal vibrations of atoms with the SHELX programs (SHELXS 2014/6 [18], SHELXL 2018/3 [19]) using the F2 data to residuals R = 0.023 and S = 1.04. As a result of the refinement, the crystal–chemical formula of the new compound was established to be Na7V4[PO4]6. The results of semiquantitative EDS analysis support it. Furthermore, the formula’s electroneutrality may be achieved by supposing a mixed population of two octahedral structural sites by V3+ and V2+ ions. Therefore, the compound presents a rare case of V2+/V3+ heterovalent phosphate with the crystal–chemical formula Na7(V3+0.75V2+0.25)(V3+0.75V2+0.25)3[PO4]6. The simplified formula Na7V3+3V2+[PO4]6 clearly shows three times more V3+ than V2+ ions. The range of V–O distances, bond valence calculation, and XANES spectroscopy confirm it. In Table 1, we report the crystallographic characteristics of the phase, the experimental conditions of data collection, and the results of crystal structure refinement. Table S1 lists the atomic coordinates and equivalent anisotropic displacement parameters. Figure S1 shows the calculated pattern for Na7V4[PO4]6. Characteristic distances are given in Table 2. We deposited structural data via the joint CCDC/FIZ Karlsruhe deposition service as CSD 2214946. It can be obtained free of charge from FIZ Karlsruhe via www.ccdc.cam.ac.uk/structures accessed on 24 October 2022.
To confirm the presence of V2+ in the compound, we studied X-ray absorption near-edge structure (XANES) at the K-edge of vanadium (obtained at the Structural Materials Science beamline [23] of the Kurchatov Synchrotron Radiation Source, Moscow, Russia). The parameters of the experiment were the following. The electron beam’s energy of X-ray synchrotron radiation was 2.5 GeV at an average current of 60–100 mA. The absorption spectra were collected in the fluorescent radiation detection mode, i.e., the X-ray intensity was measured up to the sample, and the fluorescent radiation was detected. An ionization chamber was used to measure the X-ray beam intensity. The fluorescent radiation was captured with the help of an Amptek SDD detector. X-ray absorption spectra were processed by standard procedures for background extraction and normalized to the magnitude of the absorption jump. This procedure was performed using the Larch software package [24].

3. Results

3.1. Crystal Structure Description in terms of Close-Packing

The crystal structure of Na7V4(PO4)6 is based on the main structural units shown in Figure 2a. They are two symmetrically independent vanadium octahedra and an orthophosphate tetrahedron. The PO4 polyhedron has minimal C1 symmetry and represents a distorted trigonal pyramid. Three oxygen atoms O1–O3 are located at distances of 1.543(1)–1.568(1) Å from the P atom (Table 2). They lie in a plane approximately parallel to (001). Additionally, they are involved in the coordination of vanadium cations (Figure 2b). The fourth oxygen vertex O4 at a shortened P–O4 distance of 1.510(1) Å points almost perpendicular to the base of the phosphate tetrahedron along the c-axis (Figure 2 and Figure 3). In addition to phosphorus, the O4 atom coordinates sodium cations. V atoms occupy two positions in the structure with site symmetry 32 and 2. In an octahedron with D3 symmetry, all V1–O3 bond lengths are equivalent and equal to 2.017(1) Å; in the octahedron with lower C2 symmetry, the V-O bond lengths vary from 1.988(1) to 2.055(1) Å (Table 2). The average V–O distances in both polyhedra are close and equal to 2.017 and 2.012 Å, respectively.
In the crystal structure of Na7V4(PO4)6, vanadium polyhedra unite in the cluster, in which every V1O6 octahedron on a three-fold axis shares common edges with three surrounding V2O6 octahedra (Figure 2b and Figure 3a). These clusters do not link with each other along the [001] direction. Instead, they form the close-packed layers parallel to the (001) plane and alternate along the c-axis in a sequence (ABCABC) (Figure 3b and Figure 4a). Furthermore, all oxygen vertices of VO6 polyhedra in the cluster are shared with orthophosphate tetrahedra. Thus, tetramers of V-centered octahedra unite via PO4 groups along the [001] direction into a close-packed framework crystal structure (Figure 3).
The negative charge of the heteropolyhedral anionic framework [V4(PO4)6]7- is compensated by Na cations, which occupy two symmetrically different sites in the structure. Regular Na1O6 octahedra with C3i symmetry have six identical Na1–O4 distances of 2.362(1) Å (Table 2). The Na2O6 octahedron with intrinsic C1 symmetry is strongly distorted: the Na-O bond lengths range in the interval 2.271(2)–2.640(2) Å with an average value of 2.534 Å. Three additional oxygen atoms are at distances of 2.808(2), 2.877(2), and 3.021(1) Å (Table 2). Sodium polyhedra share faces and edges to form clusters of seven units with a central Na1O6 octahedron and Na2O6 octahedra multiplied by six due to the 3 ¯ inversion axis. Similar to vanadium tetramers, these clusters in a cut layer parallel to the (001) plane are arranged in the densest manner (Figure 4). These layered fragments alternate in the sequence (ABCABC) corresponding to the cubic close-packing (Figure 3). Along the c-axis, sodium heptamers interconnect through common O4–O4 edges of Na2O6 polyhedra and form a three-periodic cationic framework.

3.2. BVS Calculations and XANES Spectroscopy

The requirement for electroneutrality of the title compound is achieved by including vanadium cations of two valence states, V2+ and V3+. Close values of V–O distances characterize both polyhedra centered with vanadium (Table 2). Furthermore, the average V–O bond lengths are almost equal: 2.017 and 2.012 Å, which indicates the mixed occupancy of both structural positions by V atoms in the 3+ and 2+ oxidation states. The calculation of bond–valence sums [25,26] confirms the statistical distribution of V3+ and V2+ ions (Table 3). Based on the theoretical ratio V3+/V2+ = 3:1, the total vanadium contributions 2.83 and 2.87 are quite close to 2.75, which indicates a mixed valence state of V in both structural sites: Na7V2.75V2.753[PO4]6.
In order to determine the formal oxidation state of the absorbing atom (V), the analysis of the shape and position of the XANES spectrum could be used [27,28,29,30]. We applied the procedure described in [31]. The oxidation state of vanadium atoms in the sample under study was determined by interpolating the edge position shift. We considered that the position of the edge jump is the energy at which the normalized absorption equals 0.5. Therefore, the spectrum shift at the level of half the absorption jump is assumed to be proportional to the formal oxidation state of the metal (V). The experimental XANES spectra are shown in Figure 5. The spectra of compounds VSO4·6H2O and V2(SO4)3 [32] were used as standards because of the same octahedral coordination of vanadium atoms as in the Na7V4(PO4)6 phase. Thus, it was determined that the formal oxidation state of the metal is 2.57, which qualitatively confirms the assumption that there are vanadium atoms with two oxidation states: 3+ and 2+.

3.3. The Comparative Crystal Chemistry of Na7V4(PO4)6 and the Database Survey of V2+-Containing Oxosalts

As mentioned in the Introduction, the titled vanadium phosphate is a member of the family of compounds with the general formula (X)6(M1)(M2)(M3)3(TO4)6, which includes the fumarole mineral yurmarinite, Na7(Fe3+,Mg,Cu)4(AsO4)6 [5] and synthetic iron (II, III) phosphate Na7Fe4(PO4)6 [6]. In these two phases, the valence state of the Fe atoms was confirmed by Mössbauer or Raman spectroscopy. Table 4 shows the cation–anion distances for the transition-metal octahedral sites M2 (Wyckoff site 6a, site symmetry 32) and M3 (Wyckoff site 18e, site symmetry 2) populated by V or Fe. The close average values of the M−O bond lengths in the M2O6 and M3O6 octahedra, equal to 2.02 Å and 2.01 Å, indicate the mixed occupancy of the V positions in the Na7V4(PO4)6 crystal structure. In addition, the M–O bond lengths are close enough in the Fe- and V-bearing phases. This similarity results from the close values of the cationic radii of these transition metals in the same valence states and in the same octahedral coordination (rV2+ = 0.79 Å, rFe2+ = 0.78 Å; rV3+ = 0.64 Å, rFe3+ = 0.63 Å [33]). The V ionic radii are slightly smaller than those of iron; accordingly, the unit-cell parameters and volume of Na7V4(PO4)6 are also smaller than those of Na7Fe4(PO4)6. The unit-cell parameters of yurmarinite are the largest due to the larger size of arsenate tetrahedra compared to phosphate groups.
Table 5 lists all V2+-bearing oxosalts with known crystal structures collected in the ICSD database. Five listed phases include exceptionally V2+ compounds, while seven of the other listed compounds contain both V2+ and V3+ cations. In all crystal structures, V occupies octahedral positions.
Urusov and Serezhkin [48] showed that the V2+ and V3+ cations form regular octahedra with average V–O bond lengths equal to 2.13 and 2.01 Å. Similarly, the data from Table 5 show that the average V–O distances range slightly from 2.12 Å to 2.18 Å for V2+-centered octahedra and from 2.01 Å to 2.09 Å for polyhedra isomorphically occupied by V2+ and V3+ cations. It is consistent with the smaller ionic radius of V3+ (0.64 Å) compared to V2+ (0.79 Å) [33]. For the title Na7V4(PO4)6 compound, with a mixed valence state of vanadium, the average <V–O> distance of 2.02 Å lies in the lower part of the region for compounds with V2+,3+-centered polyhedra.
Although the first inorganic oxide of divalent vanadium was synthesized more than a hundred years ago, the number of compounds containing V2+ ions in “formula” amounts is still small. The main reason for the small number of V2+ oxosalts is their high instability in the air, which significantly complicates the synthesis and preparation of precursors [49]. Three methods for synthesizing vanadium oxosalts are commonly used, and all require an inert atmosphere and redox reagents. First, Kranz suggested a synthesis by a stepwise reduction of V2O5 by electrolysis [50]. This method was modified by Shlenck’s techniques and became the basis for the synthesizing of V2+ hydrate sulfates (Table 5). One major feature of these compounds is the presence of water molecules in V2+ octahedral coordination. Consequently, all the structures possess low density equal to 1.8–1.9 Mg/m3.
The second method, the Li-intercalation, was used to obtain Li2V2+(PO4)F related to tavorite and Li2V2+2(SO4)3 with the NASICON-derived structure. Later, by the third method, heterovalent (V2+,3+) compounds were synthesized using a high-temperature solid-state reaction. As a result, α-CrPO4-type NaV3+,2+3(PO4)3, two polymorphic oxophosphates V2+,3+2OPO4, two isostructural oxides SrV3+,2+10O15 and BaV3+,2+10O15, and two spinel polymorphs Al(V2+,3+)2O4 were obtained by the third method. Interestingly, both spinels and exceptionally Sr or Ba vanadium oxides feature structures based on the hexagonal close-packing and thus have a high density of 4.6–5.2 Mg/m3. Noteworthy, the same inverse spinel, Al(V2+,V3+)2O4, called dellagiustaite, was recently found in nature and presented a unique mineral with V2+ cations as an essential element in its composition. However, it is difficult to establish the genesis of dellagiustaite due to the lack of information about the exact location of the outcrop and related rocks [4]. However, exclusively reducing conditions are necessary for mineral formation because it is associated with almost pure vanadium alloys.
We successfully obtained heterovalent V(II, III) phosphate under hydrothermal conditions. In solution under normal conditions, vanadium dissolves and exists as the V2+ cation under acidic and redox conditions exclusively, i.e., at pH = −2–2 and Eh = −1–0 V, according to the thermodynamic data of the Gibbs free energy for vanadium (see, for example, [51]). The high solubility of phosphoric acid and the use of Na3C6H5O7 as a redox agent allowed us to obtain two sodium vanadium phosphates: NASICON Na3V3+2(PO4)3 with the kosnarite structure type and Na7V3+,2+4(PO4)6 with the yurmarinite structure type. Therefore, we can assume that the "metastable" Na7V4(PO4)6 was formed as a primary phase at a higher temperature and a lower pH value. It is likely that the more stable compound Na3V2(PO4)3 was formed subsequently upon smooth cooling of the autoclave.

3.4. Topological Analysis of Ion Conductivity of Na7V4(PO4)6

Vanadium provides an excellent opportunity to exchange more than one electron per transition metal, resulting in a high theoretical energy density [52]. In addition, the presence of many Na+ ions in the crystal structure can also lead to the possible activation of plenty of redox transitions V2+/V3+/V4+/V5+ during the electrochemical extraction of sodium. Currently, four Na,V3+ phosphates are known to possess electrochemical properties. We obtained one of them, Na3V2(PO4)3, as a byproduct of the synthesis. It is a synthetic variety of mineral kosnarite, KZr2(PO4)3, whose crystal structure is close to that of the sodium superionic conductor NASICON. Na3V2(PO4)3 exhibits high energy density, thermal and structure stability, ion conductivity, and a high-voltage plateau at 3.4 V [53,54,55,56]. Likewise, other Na,V3+ phosphates, namely Na3V(PO4)2 [57,58], Na3V3(PO4)4 [59], and NaV3(PO4)3 [30,46] represent high-voltage cathode/high-performance anode materials for sodium batteries [60]. We emphasize that the Na3V(PO4)2 crystal structure is also related to mineral aphthitalite; both Na3V3(PO4)4 and NaV3(PO4)3 crystal structures are closely related to that of α-CrPO4. To evaluate the electrochemical properties without sufficient substance for experimental measurements, we applied the theoretical calculation of the ionic conductivity of Na7V4(PO4)6.
To find the migration paths of Na+ cations throughout the framework [61], we used the geometrical–topological approach from the ToposPro package [62]. The method is based on the Voronoi–Dirichlet partition model, which helps to find out the centers of vacations and channel lines. The Voronoi–Dirichlet partition model is founded on the following basement. The convex Voronoi polyhedra bend around each of the “framework” atoms. This bending results in the forming of voids located most distant from the structure atoms. Linked together, voids form the net, whose periodicity reflects the ion migration path’s topology. Remarkably, the net of voids should be infinite—like a chain, a layer, or a framework—to promote conductivity.
The primary geometric parameters for the Voronoi–Dirichlet model are the radius of the spherical domain Rsd and the channel radius Rch. In our calculation, the threshold value of Rsd is set equal to 1.31 Å, which is 10% lower than the experimental value for “real” sodium atoms in the structure (1.50 Å and 1.46 Å). It is much smaller than the theoretical average value of 1.54 Å [63] and close to 1.35 Å estimated from the migration maps of Na-ion conductors in recent studies [64,65]. The threshold value of Rch was asserted to be 2.0 Å following [64]. Theoretically, the structural channel is assumed accessible for ion flotation if the sum of the radii of the mobile ion (Na) and the framework atom (O) exceeds the channel radius Rch by 10–15%. In contrast with the theoretical preposition, the topological analysis shows the absence of migration paths for Na+ cations, even though the lowest values of the threshold parameters were taken. Small structural voids with a radius of 1.31 Å to 1.57 Å do not connect with each other in the crystal structure of Na7V4(PO4)6 and therefore do not form infinite channels large enough for Na+ passing.

4. Conclusions

Under high-temperature (450 °C) hydrothermal conditions, we successfully synthesized a new phase Na7V4(PO4)6, a V3+,2+ analogue of the fumarole mineral yurmarinite. We showed that the crystal structure of Na7V4(PO4)6 is based on densely packed layers of vanadium or sodium clusters, which alternate in the ABCABC sequence, and link by [PO4] tetrahedra. Besides the new compound Na7V4(PO4)6, the research revealed only twelve oxosalts with V2+ or isomorphic V2+/V3+ cations, including the unique divalent vanadium mineral dellagiustaite. Our study presents the first case of hydrothermally synthesized Na7V3+,2+4(PO4)6 crystals in addition to electrolysis, Li-intercalation, and the solid-state method used previously to prepare V2+ inorganic compounds. We conducted the X-ray diffraction study, BVS calculations, crystal–chemical comparative analysis of V2+-bearing compounds, and XANES measurements of Na7V4(PO4)6 crystals, which confirmed the presence of V2+ ions in it. Furthermore, the disability of Na-ion migration through Na7V4(PO4)6 was shown by performing the topological analysis of its crystal structure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min12121517/s1, Table S1: Atomic coordinates and equivalent anisotropic displacement parameters for Na7V4(PO4)6; Figure S1: Calculated pattern for Na7V4[PO4]6.

Author Contributions

Conceptualization, O.Y.; methodology, G.K., A.V., A.T. and K.L.; validation, O.Y.; investigation, G.K., V.N., O.Y., A.V., O.D., A.T. and K.L.; writing—original draft preparation, V.N. and G.K.; writing—review and editing, O.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant from the President of the Russian Federation for young scientists—candidates of science MK-1613.2021.1.5.

Data Availability Statement

Not applicable.

Acknowledgments

We thank V.O. Yapaskurt for the microprobe analysis of the crystals and E.I. Marchenko for helping with the ToposPro tools. We are grateful to O.M. Kiriukhin from the City University of Hong Kong, for checking the English.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, C.; Eleish, A.; Hystad, G.; Golden, J.J.; Downs, R.T.; Morrison, S.M.; Hummer, D.R.; Ralph, J.P.; Fox, P.; Hazen, R.M. Analysis and visualization of vanadium mineral diversity and distribution. Am. Miner. 2018, 103, 1080–1086. [Google Scholar] [CrossRef] [Green Version]
  2. Schindler, M.; Hawthorne, F.C.; Baur, W.H. A crystal-chemical approach to the composition and occurrence of vanadium minerals. Can. Miner. 2000, 38, 1443–1456. [Google Scholar] [CrossRef] [Green Version]
  3. O’Loughlin, E.J.; Boyanov, M.I.; Kemner, K.M. Reduction of Vanadium(V) by Iron(II)-Bearing Minerals. Minerals 2021, 11, 316. [Google Scholar] [CrossRef]
  4. Cámara, F.; Bindi, L.; Pagano, A.; Pagano, R.; Gain, S.E.; Griffin, W.L. Dellagiustaite: A Novel Natural Spinel Containing V2+. Minerals 2019, 9, 4. [Google Scholar] [CrossRef] [Green Version]
  5. Pekov, I.V.; Zubkova, N.V.; Yapaskurt, V.O.; Belakovskiy, D.I.; Lykova, I.S.; Vigasina, M.F.; Sidorov, E.G.; Pushcharovsky, D.Y. New arsenate minerals from the Arsenatnaya fumarole, Tolbachik volcano, Kamchatka, Russia. I. Yurmarinite, Na7(Fe3+,Mg,Cu)4(AsO4)6. Miner. Mag. 2014, 78, 905–917. [Google Scholar] [CrossRef]
  6. Lii, K.-H. Na7Fe4(PO4)6: A mixed-valence iron phosphate containing a tetramer of edge-sharing FeO6 octahedra. J. Chem. Soc. Dalton Trans. 1996, 6, 819–822. [Google Scholar] [CrossRef]
  7. Masquelier, C.; d’Yvoire, F.; Collin, G. Crystal structure of Na7Fe4(AsO4)6 and α-Na3Al2(AsO4)3, two sodium ion conductors structurally related to II-Na3Fe2(AsO4)3. J. Solid State Chem. 1995, 118, 33–42. [Google Scholar] [CrossRef]
  8. Bourguiba, N.F.; Ouerfelli, N.; Zid, M.F. Crystal Structure, Charge-Distribution and Bond-Valence-Sum Investigations of a Novel Arsenate Na3.5Cr1.83(AsO4)3. J. Adv. Chem. 2014, 10, 3160–3170. [Google Scholar]
  9. Bourguiba, N.F.; Guesmi, A.; Ouerfelli, N.; Mazza, D.; Zid, M.F.; Driss, A. Le Nouvel Arséniate Ag1.60Na1.40Al2(AsO4)3: Synthèse, Études Structurale Et Électrique, Simulation Des Chemins De Conduction Ionique. J. Soc. Chim. Tunis. 2012, 14, 201–211. [Google Scholar]
  10. Gulyaeva, O.A.; Solodovnikova, Z.A.; Solodovnikov, S.F.; Zolotova, E.S.; Mateyshina, Y.G.; Uvarov, N.F. Triple molybdates K3–xNa1 + xM4(MoO4)6 (M = Ni, Mg, Co) and K3+xLi1–xMg4(MoO4)6 isotypic with II-Na3Fe2(AsO4)3 and yurmarinite: Synthesis, potassium disorder, crystal chemistry and ionic conductivity. Acta Cryst. 2020, B76, 913–925. [Google Scholar]
  11. D‘Yvoire, F.; Bretey, E.; Collin, G. Crystal structure, non-stoichiometry and conductivity of II-Na3M2(AsO4)3 (M = Al, Ga, Cr, Fe). Solid State Ion. 1988, 28, 1259–1264. [Google Scholar] [CrossRef]
  12. Yakubovich, O.V. Microporous Vanadylphosphates—Perspective Materials for Technological Applications. In Minerals as Advanced Materials II; Krivovichev, S., Ed.; Springer: Berlin, Heidelberg, 2011; pp. 239–253. [Google Scholar]
  13. Yakubovich, O.V.; Steele, I.M.; Yakovleva, E.V.; Dimitrova, O.V. One-dimensional decavanadate chains in the crystal structure of Rb4[Na(H2O)6][HV10O28]·4H2O. Acta Cryst. 2015, C71, 465–473. [Google Scholar]
  14. Yakubovich, O.V.; Kiryukhina, G.V.; Dimitrova, O.V. Crystal chemistry of KCuMn3(VO4)3 in the context of detailed systematics of the alluaudite family. Crystallogr. Rep. 2016, 61, 566–575. [Google Scholar] [CrossRef]
  15. Shvanskaya, L.V.; Yakubovich, O.V.; Krikunova, P.V.; Kiriukhina, G.V.; Ivanova, A.G.; Volkov, A.S.; Dimitrova, O.V.; Borovikova, E.Y.; Volkova, O.S.; Vasiliev, A.N. Nonstoichiometric ellenbergerite-type phosphates: Hydrothermal synthesis, crystal chemistry, and magnetic behavior. Inorg. Chem. 2022, 61, 4879–4886. [Google Scholar] [CrossRef]
  16. Zatovsky, I.V. Nasicon-type Na3V2(PO4)3. Acta Cryst. 2010, E66, 112. [Google Scholar] [CrossRef] [Green Version]
  17. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Cryst. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  18. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  19. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  20. Rigaku, O.D. CrysAlis PRO; Rigaku Oxford Diffraction Ltd.: Yarnton, Oxfordshire, UK, 2018. [Google Scholar]
  21. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  22. Brandenburg, K. Diamond; Crystal Impact GbR: Bonn, Germany, 2006. [Google Scholar]
  23. Chernyshov, A.A.; Veligzhanin, A.A.; Zubavichus, Y.V. Structural Materials Science end-station at the Kurchatov Synchrotron Radiation Source: Recent instrumentation upgrades and experimental results. Nucl. Instrum. Methods Phys. Res. 2009, A603, 95–98. [Google Scholar] [CrossRef]
  24. Newville, M. Larch: An analysis package for XAFS and related spectroscopies. J. Phys. Conf. Ser. 2013, 430, 012007. [Google Scholar] [CrossRef]
  25. Brown, I.D.; Altermatt, D. Bond-valence parameters obtained from a systematic analysis of the inorganic crystal structure database. Acta Cryst. 1985, B41, 244–247. [Google Scholar] [CrossRef] [Green Version]
  26. Brown, I.D. Recent developments in the methods and applications of the bond valence model. Chem. Rev. 2009, 109, 6858–6919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Charles, D.S.; Guo, F.; Shan, X.; Kim, S.; Lebens-Higgins, Z.W.; Xu, W.; Su, D.; Piper, L.F.J.; Teng, X. Dual-stage K+ ion intercalation in V2O5-conductive polymer composites. J. Mater. Chem. 2021, A9, 15629–15636. [Google Scholar] [CrossRef]
  28. Dau, H.; Iuzzolino, L.; Dittmer, J. The tetra-manganese complex of photosystem II during its redox cycle—X-ray absorption results and mechanistic implications. Biochim. Biophys. Acta Bioenerg. 2001, 1503, 24–39. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, J.B.; Chae, O.B.; Chae, S.; Ryu, J.H.; Oh, S.M. Amorphous Vanadium Titanates as a Negative Electrode for Lithium-ion Batteries. J. Electrochem. Sci. Technol. 2016, 7, 306–315. [Google Scholar] [CrossRef]
  30. Wang, X.; Hu, P.; Chen, L.; Yao, Y.; Kong, Q.; Cui, G.; Shi, S.; Chen, L. An α-CrPO4-type NaV3(PO4)3 anode for sodium-ion batteries with excellent cycling stability and the exploration of sodium storage behavior. J. Mater. Chem. 2017, A5, 3839–3847. [Google Scholar] [CrossRef]
  31. Gustafsson, J.P. Vanadium geochemistry in the biogeosphere–speciation, solid-solution interactions, and ecotoxicity. Appl. Geochem. 2019, 102, 1–25. [Google Scholar] [CrossRef]
  32. Haskel, D.; Islam, Z.; Lang, J.; Kmety, C.; Srajer, G.; Pokhodnya, K.I.; Epstein, A.J.; Miller, J.S. Local structural order in the disordered vanadium tetracyanoethylene room-temperature molecule-based magnet. Phys. Rev. 2004, B70, 054422. [Google Scholar] [CrossRef] [Green Version]
  33. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, B32, 751–767. [Google Scholar] [CrossRef]
  34. Montgomery, H.; Morosin, B.; Natt, J.J.; Witkowska, A.T.; Lingafelter, E.C. The crystal structure of Tutton’s salts. VI. Vanadium (II), iron (II) and cobalt (II) ammonium sulfate hexahydrates. Acta Cryst. 1967, 22, 775–780. [Google Scholar] [CrossRef] [Green Version]
  35. Cotton, F.A.; Falvello, L.R.; Llusar, R.; Libby, E.; Murillo, C.A.; Schwotzer, W. Synthesis and characterization of four vanadium (II) compounds, including vanadium (II) sulfate hexahydrate and vanadium (II) saccharinates. Inorg. Chem. 1986, 25, 3423–3428. [Google Scholar] [CrossRef]
  36. Cotton, F.A.; Falvello, L.R.; Murillo, C.A.; Pascual, I.; Schultz, A.J.; Tomás, M. Neutron and X-ray Structural Characterization of the Hexaaquavanadium (II) Compound VSO4·7H2O. Inorg. Chem. 1994, 33, 5391–5395. [Google Scholar] [CrossRef]
  37. Ellis, B.L.; Ramesh, T.N.; Davis, L.J.M.; Goward, G.R.; Nazar, L.F. Structure and electrochemistry of two-electron redox couples in lithium metal fluorophosphates based on the tavorite structure. Chem. Mater. 2011, 23, 5138–5148. [Google Scholar] [CrossRef]
  38. Vaughan, G.B.M.; Gaubicher, J.; Le Mercier, T.; Angenault, J.; Quarton, M.; Chabre, Y. Crystal structure of the end product of electrochemical lithium intercalation in V2(SO4)3. Mater. Chem. 1999, 9, 2809–2812. [Google Scholar] [CrossRef]
  39. Glaum, R.; Gruehn, R. Beiträge zum thermischen Verhalten von wasserfreien Phosphaten. Synthese, Kristallstruktur und magnetisches Verhalten von V2PO5. Z. Kristallogr. Cryst. Mater. 1989, 186, 91–93. [Google Scholar]
  40. Pachoud, E.; Cumby, J.; Lithgow, C.T.; Attfield, J.P. Charge order and negative thermal expansion in V2OPO4. J. Am. Chem. Soc. 2018, 140, 636–641. [Google Scholar] [CrossRef] [Green Version]
  41. De Beaulieu, D.C.; Müller-Buschbaum, H. Gemischtvalente Oxovanadate. II. Synthese und Kristallstruktur von SrV10O15. Z. Anorg. Allg. Chem. 1981, 472, 33–37. [Google Scholar] [CrossRef]
  42. Matsuno, K.I.; Katsufuji, T.; Mori, S.; Moritomo, Y.; Machida, A.; Nishibori, E.; Takata, M.; Sakata, M.; Yamamoto, N.; Takagi, H. Orbital Order and Phase Transitions in Vanadium Spinels. J. Phys. Soc. Japan 2001, 70, 1456–1459. [Google Scholar] [CrossRef]
  43. Horibe, Y.; Shingu, M.; Kurushima, K.; Ishibashi, H.; Ikeda, N.; Kato, K.; Motome, Y.; Furukawa, N.; Mori, S.; Katsufuji, T. Spontaneous formation of vanadium “molecules” in a geometrically frustrated crystal: AlV2O4. Phys. Rev. Lett. 2006, 96, 086406. [Google Scholar] [CrossRef]
  44. Reuter, B.; Aust, R.; Colsmann, G.; Neuwald, C. Darstellung und Eigenschaften Vanadium (II)-haltiger und damit n-leitender Vanadium (III)-Spinelle. Z. Anorg. Allg. Chem. 1983, 500, 188–198. [Google Scholar] [CrossRef]
  45. Kalavathi, S.; Raju, S.V.; Williams, Q.; Sahu, P.C.; Sastry, V.S.; Sahu, H.K. Pressure-induced frustration in charge ordered spinel AlV2O4. J. Phys. Cond. Matter 2013, 25, 292201. [Google Scholar] [CrossRef] [PubMed]
  46. Kinomura, N.; Matsui, N.; Kumada, N.; Muto, F. Synthesis and crystal structure of NaV3P3O12: A stuffed structure of α-CrPO4. J. Solid State Chem. 1989, 79, 232–237. [Google Scholar] [CrossRef]
  47. Liu, G.; Greedan, J.E. Structure and Magnetism in AB10O15+x (A= Sr, Ba; B= V, Cr; x= 0, ∼1): Evidence for Geometric Frustration. J. Solid State Chem. 1996, 122, 416–427. [Google Scholar] [CrossRef]
  48. Urusov, V.S.; Serezhkin, V.N. Distortion of Vz+On coordination polyhedra and parameters of the bond valence model for VO bonds in inorganic crystals. Crystallogr. Rep. 2009, 54, 190–194. [Google Scholar] [CrossRef]
  49. Cotton, F.A.; Daniels, L.M.; Montero, M.L.; Murillo, C.A. Preparation of vanadium (II) compounds. Structures of a carboxylato and a 3-pyridinesulphonate compound. Polyhedron 1992, 11, 2767–2774. [Google Scholar] [CrossRef]
  50. Kranz, M.; Goldberg, D.E.; Ribner, A. Vanadium (II) Sulfate. Inorg. Synth. 1963, 7, 94–96. [Google Scholar]
  51. Li, W.; Niu, Z.; Zhu, X. Recovery of iron by jarosite crystallization and separation of vanadium by solvent extraction with extractant 7101 from titanium white waste liquid (TWWL). Water Sci. Technol. 2021, 83, 2025–2037. [Google Scholar] [CrossRef]
  52. Zeng, X.; Peng, J.; Guo, Y.; Zhu, H.; Huang, X. Research progress on Na3V2(PO4)3 cathode material of sodium ion battery. Front. Chem. 2020, 8, 635. [Google Scholar] [CrossRef]
  53. Zhang, B.; Ma, K.; Lv, X.; Shi, K.; Wang, Y.; Nian, Z.; Li, Y.; Wang, L.; Dai, L.; He, Z. Recent advances of NASICON-Na3V2(PO4)3 as cathode for sodium-ion batteries: Synthesis, modifications, and perspectives. J. Alloys Compd. 2021, 867, 159060. [Google Scholar] [CrossRef]
  54. Zheng, Q.; Yi, H.; Li, X.; Zhang, H. Progress and prospect for NASICON-type Na3V2(PO4)3 for electrochemical energy storage. J. Energy Chem. 2018, 27, 1597–1617. [Google Scholar] [CrossRef] [Green Version]
  55. Jian, Z.; Han, W.; Lu, X.; Yang, H.; Hu, Y.-S.; Zhou, J.; Zhou, Z.; Li, J.; Chen, W.; Chen, D.; et al. Superior electrochemical performance and storage mechanism of Na3V2(PO4)3 cathode for room-temperature sodium-ion batteries. Adv. Energy Mater. 2013, 3, 156–160. [Google Scholar] [CrossRef]
  56. Saravanan, K.; Mason, C.W.; Rudola, A.; Wong, K.H.; Balaya, P. The first report on excellent cycling stability and superior rate capability of Na3V2(PO4)3 for sodium ion batteries. Adv. Energy Mater. 2013, 3, 444–450. [Google Scholar] [CrossRef]
  57. Kovrugin, V.M.; David, R.; Chotard, J.N.; Recham, N.; Masquelier, C. A high voltage cathode material for sodium batteries: Na3V(PO4)2. Inorg. Chem. 2018, 57, 8760–8768. [Google Scholar] [CrossRef] [PubMed]
  58. Kim, J.; Yoon, G.; Kim, H.; Park, Y.U.; Kang, K. Na3V(PO4)2: A new layered-type cathode material with high water stability and power capability for Na-ion batteries. Chem. Mater. 2018, 30, 3683–3689. [Google Scholar] [CrossRef]
  59. Liu, R.; Liu, H.; Sheng, T.; Zheng, S.; Zhong, G.; Zheng, G.; Liang, Z.; Ortiz, G.F.; Zhao, W.; Mi, J.; et al. Novel 3.9 V Layered Na3V3(PO4)4 Cathode Material for Sodium Ion Batteries. ACS Appl. Energy Mater. 2018, 1, 3603–3606. [Google Scholar] [CrossRef]
  60. Ke, L.; Dong, J.; Lin, B.; Yu, T.; Wang, H.; Zhang, S.; Deng, C. A NaV3(PO4)3@C hierarchical nanofiber in high alignment: Exploring a novel high-performance anode for aqueous rechargeable sodium batteries. Nanoscale 2017, 9, 4183–4190. [Google Scholar] [CrossRef]
  61. Blatov, V.A.; Shevchenko, A.P. Analysis of voids in crystal structures: The methods of “dual” crystal chemistry. Acta Cryst. 2003, A59, 34–44. [Google Scholar] [CrossRef] [Green Version]
  62. Blatov, V.A.; Shevchenko, A.P.; Proserpio, D.M. Applied topological analysis of crystal structures with the program package ToposPro. Cryst. Growth Des. 2014, 14, 3576–3586. [Google Scholar] [CrossRef]
  63. Blatov, V.A. Voronoi–dirichlet polyhedra in crystal chemistry: Theory and applications. Crystallogr. Rev. 2004, 10, 249–318. [Google Scholar] [CrossRef]
  64. Fedotov, S.S.; Kabanova, N.A.; Kabanov, A.A.; Blatov, V.A.; Khasanova, N.R.; Antipov, E.V. Crystallochemical tools in the search for cathode materials of rechargeable Na-ion batteries and analysis of their transport properties. Solid State Ion. 2018, 314, 129–140. [Google Scholar] [CrossRef]
  65. Meutzner, F.; Münchgesang, W.; Leisegang, T.; Schmid, R.; Zschornak, M.; Ureña de Vivanco, M.; Shevchenko, A.P.; Blatov, V.A.; Meyer, D.C. Identification of solid oxygen-containing Na-electrolytes: An assessment based on crystallographic and economic parameters. Cryst. Res. Technol. 2017, 52, 1600223. [Google Scholar] [CrossRef]
Figure 1. SEM image of the Na7V4(PO4)6 crystal with rhombohedral and prismatic faceting.
Figure 1. SEM image of the Na7V4(PO4)6 crystal with rhombohedral and prismatic faceting.
Minerals 12 01517 g001
Figure 2. The main building units of the Na7V4(PO4)6 structure. (a) Atoms are shown in ellipsoid mode at a 90% probability level (symmetry operations: (i) −x + 2/3, −x + y + 1/3, −z + 5/6; (ii) −x + y, −x + 1, z; (iii) −y + 1, xy + 1, z; (iv) xy + 2/3, −y + 4/3, −z + 5/6; (v) y − 1/3, x+1/3, −z + 5/6; (vi) x + 1/3, xy + 2/3, z + 1/6; (vii) y + 1/3, −x + y + 2/3, −z + 2/3). (b) The cluster of four vanadium octahedra in surrounding phosphate.
Figure 2. The main building units of the Na7V4(PO4)6 structure. (a) Atoms are shown in ellipsoid mode at a 90% probability level (symmetry operations: (i) −x + 2/3, −x + y + 1/3, −z + 5/6; (ii) −x + y, −x + 1, z; (iii) −y + 1, xy + 1, z; (iv) xy + 2/3, −y + 4/3, −z + 5/6; (v) y − 1/3, x+1/3, −z + 5/6; (vi) x + 1/3, xy + 2/3, z + 1/6; (vii) y + 1/3, −x + y + 2/3, −z + 2/3). (b) The cluster of four vanadium octahedra in surrounding phosphate.
Minerals 12 01517 g002
Figure 3. Crystal structure of the Na7V4(PO4)6 projected along (a) the [0001] and (b) the [11 2 ¯ 0] directions.
Figure 3. Crystal structure of the Na7V4(PO4)6 projected along (a) the [0001] and (b) the [11 2 ¯ 0] directions.
Minerals 12 01517 g003
Figure 4. Layered “cut” of the crystal structure of Na7V4(PO4)6 on the ab plane: (a) clusters of four vanadium octahedra surrounded by PO4-tetrahedra and (b) clusters of sodium octahedra, both arranged according to the close-packing law.
Figure 4. Layered “cut” of the crystal structure of Na7V4(PO4)6 on the ab plane: (a) clusters of four vanadium octahedra surrounded by PO4-tetrahedra and (b) clusters of sodium octahedra, both arranged according to the close-packing law.
Minerals 12 01517 g004
Figure 5. XANES data for the V K-edge of Na7V4(PO4)6 (signed as V_8885) with overlayed spectra of VSO4·6H2O and V2(SO4)3.
Figure 5. XANES data for the V K-edge of Na7V4(PO4)6 (signed as V_8885) with overlayed spectra of VSO4·6H2O and V2(SO4)3.
Minerals 12 01517 g005
Table 1. Experimental details for Na7V4(PO4)6.
Table 1. Experimental details for Na7V4(PO4)6.
Crystal Data
Chemical formulaNa7V4P6O24
Mr934.51
Crystal system, space groupTrigonal, R 3 ¯ c:H
Temperature (K)110
a, c (Å)13.3463 (19), 17.809 (4)
V3)2747.2 (10)
Z6
Radiation typeMo Kα
µ (mm−1)2.81
Crystal size (mm)0.08 × 0.08 × 0.04
Data Collection 1
DiffractometerBruker D8 Quest
Absorption correctionMulti-scan
Tmin, Tmax0.453, 0.494
No. of measured, independent and
observed [I > 2σ(I)] reflections
9263, 894, 781
Rint0.045
(sin θ/λ)max−1)0.703
Refinement
R[F2 > 2σ(F2)], wR(F2), S0.023, 0.059, 1.04
No. of reflections894
No. of parameters65
Δρmax, Δρmin (e Å−3)0.46, −0.49
1 Computer programs: CrysAlis PRO [20], SADABS 2016/2 [21], Diamond [22].
Table 2. Selected distances (Å).
Table 2. Selected distances (Å).
V1 OctahedronNa1 0ctahedron
V1—O3 × 62.0173 (13)Na1—O4 × 62.3653 (14)
V2 OctahedronNa2 [6+3] Polyhedron
V2—O1 × 21.9878 (13)Na2—O42.2706 (16)
V2—O2 × 21.9942 (14)Na2—O12.5290 (16)
V2—O3 × 22.0548 (13)Na2—O42.5718 (17)
Na2—O22.5721 (16)
P TetrahedronNa2—O42.6202 (17)
P—O41.5103 (14)Na2—O22.6401 (17)
P—O21.5428 (15)Na2—O12.8084 (17)
P—O11.5513 (14)Na2—O32.8765 (16)
P—O31.5676 (14)Na2—O13.0213 (14)
Table 3. Bond–valence data 1.
Table 3. Bond–valence data 1.
V1V2PNa1Na22 Σ
O1 0.510 ×2↓1.189 0.134, 0.075, 0.0491.96
O2 0.501 ×2↓1.215 0.123, 0.1071.95
O30.471 ×6↓0.425 ×2↓1.135 0.0662.10
O4 1.3280.189 ×6↓0.230, 0.123, 0.1111.98
Σ 2.832.874.871.131.02
1 Symbols ×2↓ and ×6↓ mark a multiplication of the corresponding contribution along the column due to the symmetry; 2 Contributions from all nine neighboring O atoms were taken into account.
Table 4. Unit-cell parameters (Å), volumes (Å3), and interatomic distances (Å) for V/Fe octahedra in Na7V4(PO4)6,Na7Fe4(PO4)6, and yurmarinite, Na7(Fe3+,Mg,Cu)4(AsO4)6 crystal structures (space group R 3 ¯  c, Z = 6).
Table 4. Unit-cell parameters (Å), volumes (Å3), and interatomic distances (Å) for V/Fe octahedra in Na7V4(PO4)6,Na7Fe4(PO4)6, and yurmarinite, Na7(Fe3+,Mg,Cu)4(AsO4)6 crystal structures (space group R 3 ¯  c, Z = 6).
Na7V4(PO4)6Na7Fe4(PO4)6Yurmarinite, Na7(Fe3+, Mg, Cu)4(AsO4)6
a (b)13.3463(19)13.392(2)13.7444(2)
c17.809(4)17.858(3)18.3077(3)
V2747.2(10)2773.7(8)2995.14(8)
M2–O3 × 62.0173(13)2.063(2)1.9429(18)
<M2–O>2.022.061.94
M3–O1 × 21.9877(13)1.974(2)1.991(2)
M3–O2 × 21.9942(14)1.997(2)2.0142(19)
M3–O3 × 22.0548(13)2.085(1)2.1170(19)
<M3–O>2.012.022.04
Table 5. Synthetic and natural oxygen-containing inorganic compounds with divalent vanadium.
Table 5. Synthetic and natural oxygen-containing inorganic compounds with divalent vanadium.
Compound/
Mineral
Space Group, Z,
ρ, mg/m3
Unit-Cell Para-Meters, a, b, c, Å and Angles,°RRange of V–O Distances in Octahedra, ÅSynthesis Technique/Natural GenesisRef.
(NH4)2V(SO4)2·6H2O
NH4,V-analogue of
picromerite and
Tutton’s salt
P21/a
2, 1.80
a = 9.42(3)
b = 12.76(3)
c = 6.22(2)
β =107.2(2)
0.08
2.118–2.164
<2.15>
Modification of Kranz method to obtain VOSO4, its further cooling and electrolytic reduction in H2SO4 solution, then the addition of pyrogallol and recrystallization in N2 atmosphere.[34]
V(H2O)6SO4
V-analogue of
hexahydrite
C2/c
8, 1.91
a = 10.081(3)
b = 7.286(2)
c = 24.445(7)
β = 98.78(2)
0.039 2.123–2.136
<2.13>
2.120–2.150
<2.13>
Modification of the Kranz method.
Electrochemical reduction of VOSO4·2H2O in the solution of H2SO4 under N2 or Ag atmosphere
followed by adding ethanol on cooling.
[35]
V(H2O)6SO4·H2OP21/c
4, 1.86
a = 14.130(3)
b = 6.501(1)
c = 11.017(2)
β = 105.64(2)
0.029 2.102–2.137
<2.12>
2.119–2.150
<2.13>
Recrystallization of VSO4·6H2O precursor in the solution of H2SO4 in N2 atmosphere with the addition of ethanol and cooling.[36]
Li2VPO4F
closely related to tavorite str. type
C2/c
4, 3.17
a = 7.2255(1)
b = 7.9450(1)
c = 7.3075(1)
β = 116.771(1)
0.049 *
2.119–2.139
<2.13>
Solid-state synthesis of LiV(PO4)F at 750 °C in an Ar atmosphere, followed by Li-intercalation with LiAlH4 in tetrahydrofuran under Ar.[37]
Li2V2(SO4)3
derivative of
NASICON str. type
C2/c
4, 3.01
a =13.0582(5)
b = 8.6526(4)
c = 8.72447(4)
β = 115.443(2)
0.042 *
2.073–2.230
<2.14>
Electrochemical Li-intercalation in V2(SO4)3 with n–butyllithium C4H9Li in a vacuum.[38]
V2+,3+2OPO4
(high-T modif.)
I41/amd
4, 3.97
a = 5.362(5)
c = 12.378(9)
0.044
2.047–2.074
<2.06>
Reduction of VPO4 by metallic V at 500 °C and 900 °C via chemical vapor transport in I2.[39]
V2+,3+2OPO4
(low-temperature
modification)
C2/c
4, 3.98
a = 7.56825(7)
b = 7.60013(7)
c = 7.21794(6)
β =121.2751(5)
0.074 *
2.054–2.123
<2.09>
1.972–2.079
<2.03>
β-VOPO4 preparation at 600 °C in flowing O2;further solid-state synthesis at 1000 °C with metallic V in a vacuum.[40]
SrV3+,2+10O15 **
Ccmb
4, 5.20
a = 9.915
b =11.574
c = 9.324
0.08
1.948–2.192 <2.01>
1.950–2.148 <2.05>
1.965–2.195 <2.04>
Solid-state synthesis at 1900 °C in H2
atmosphere
[41]
Al(V2+,3+)2O4
(frustrated spinel)
R 3 ¯ m
2, 4.65
a = 5.7613(2)
c = 28.6876(10)
0.011 *
2.052–2.063 <2.06>
2.007
Solid-state synthesis with metallic Al at 1150 °C for 150 h in vacuum, phase transition at 427 °C[42,43]
Al(V2+,3+)2O4
(inverse spinel)
Fd 3 ¯ m
8, 4.66
a = 8.1546(6)No data
2.040
Solid-state synthesis with metallic V at 900 °C in a vacuum[44,45]
Dellagiustaite,
Al(V2+,V3+)2O4
(inverse spinel)
Fd 3 ¯ m
8, 4.62
a = 8.1950(1)0.014
2.045High redox conditions, probable crystallization from high-temperature melts of volcanic magma[4]
NaV3+,2+3(PO4)3
Imma
4, 3.42
a = 10.488(2)
b =13.213(3)
c = 6.455(7)
0.056 1.959–2.062 <1.99>
2.042–2.086 <2.06>
Step-wise solid-state synthesis at 900 °C in a vacuum[46]
Na7V3+,2+4(PO4)6
V, P-analogue of
yurmarinite
R 3 ¯ c
6, 3.39
a =13.3463(19)
c = 17.809(4)
0.023 2.017
1.988–2.055
<2.01>
High-temperature hydrothermal synthesis at 450 °C with Na3C6H5O7 redox agentThis
work
* Based on power data, including synchrotron and neutron radiation sources; ** Isostructural with BaV3+,2+10O15 [47].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kiriukhina, G.; Nesterova, V.; Yakubovich, O.; Volkov, A.; Dimitrova, O.; Trigub, A.; Lyssenko, K. Mixed Valanced V3+,V2+ Phosphate Na7V4(PO4)6: A Structural Analogue of Mineral Yurmarinite. Minerals 2022, 12, 1517. https://doi.org/10.3390/min12121517

AMA Style

Kiriukhina G, Nesterova V, Yakubovich O, Volkov A, Dimitrova O, Trigub A, Lyssenko K. Mixed Valanced V3+,V2+ Phosphate Na7V4(PO4)6: A Structural Analogue of Mineral Yurmarinite. Minerals. 2022; 12(12):1517. https://doi.org/10.3390/min12121517

Chicago/Turabian Style

Kiriukhina, Galina, Valentina Nesterova, Olga Yakubovich, Anatoly Volkov, Olga Dimitrova, Alexander Trigub, and Konstantin Lyssenko. 2022. "Mixed Valanced V3+,V2+ Phosphate Na7V4(PO4)6: A Structural Analogue of Mineral Yurmarinite" Minerals 12, no. 12: 1517. https://doi.org/10.3390/min12121517

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop