Next Article in Journal
Nano-Mineralogy and Mineralization of the Polymetallic Nodules from the Interbasin of Seamounts, the Western Pacific Ocean
Previous Article in Journal
Genesis of the Large-Scale Kamado Magnesite Deposit on the Tibetan Plateau
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mineral-like Synthetic Compounds Stabilized under Hydrothermal Conditions: X-ray Diffraction Study and Comparative Crystal Chemistry

1
Department of Crystallography, Geological Faculty, Lomonosov Moscow State University, Leninskie Gory 1, 119991 Moscow, Russia
2
Korzhinskii Institute of Experimental Mineralogy of Russian Academy of Sciences, 142432 Chernogolovka, Russia
3
Skolkovo Institute of Science and Technology, 121205 Moscow, Russia
4
Institute of Solid State Physics of Russian Academy of Sciences, 142432 Chernogolovka, Russia
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(1), 46; https://doi.org/10.3390/min14010046
Submission received: 29 November 2023 / Revised: 27 December 2023 / Accepted: 28 December 2023 / Published: 29 December 2023

Abstract

:
Under hydrothermal conditions emulating natural hydrothermalites, three oxo-salts with sodium and transition metal cations were obtained in the form of single crystals. Their compositions and crystal structures were studied using scanning electron microscopy, microprobe X-ray spectral analysis, and X-ray single-crystal diffraction. The sodium cobalt silicate, i.e., Na2CoSiO4, a structural analog of the mineral liberite, is well known as an ionic conductor. Its crystal structure consists of a framework derived from β-tridymite, built using the Co- and Si-centered tetrahedra sharing vertices. The sodium oxocuprate phosphate chloride Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl belongs to a group of compounds, including fumarolic minerals, characterized by the presence of oxo-centered pyroxene-like chains in their structures. The crystal structure of mineralogically probable sodium vanadium phosphate hydroxide (Na3V(OH)(HPO4)(PO4)) is based on chains built using octahedra centered by magnetically active V3+. Magnetic susceptibility measurements indicate an antiferromagnetic arrangement of V3+ ions and no transition to an ordered state up to 2 K.

Graphical Abstract

1. Introduction

Nature has long served as a source of inspiration for scientists. Minerals, renowned for their geological stability, have played a significant role in the chemistry of materials. However, it is a well-known fact, both among materials scientists and mineralogists, that natural compounds often contain a multitude of impurities. These impurities frequently interfere with the study of the physical properties of minerals, hindering their industrial applications. To solve this problem, the laboratory gradually adopted an experimental approach that mimics the natural conditions of mineral formation.
The natural pegmatite and hydrothermal process occur in superheated water under pressure beneath the earth’s surface in a temperature range of 100–700 °C. The hydrothermal technology takes its name from hydrothermal vents where extreme temperatures and pressures occur. A distinctive feature of hydrothermal vents is the presence of supercritical water (374 °C and a pressure of 218 atm), with its unique properties to dissolve solids like a liquid yet flow with next to no viscosity like a gas. By working at the upper end of the temperature range (around 500–700 °C), it is possible to achieve conditions roughly similar to those found in natural pegmatite systems. Even under these temperature (and pressure) conditions, the liquid remains sufficiently polar to partially dissolve most metal oxyanions. Synthesis at intermediate temperatures of 100 °C < T < 374 °C allows it to remain well below the critical point of water but still heat the vessel above the normal boiling point. By keeping the reaction closed, water can be maintained in a liquid state well above the point at which it would normally boil. As long as the critical point is avoided, the pressures required for this approach are much more controllable. Undoubtedly, both thermal regimes are characteristic of the geological system [1].
These temperatures and chemical environments inherent in the hydrothermal process can be simulated in the laboratory. Therefore, by exploiting the similarities between mineral forming conditions and laboratory techniques, it becomes possible to investigate the structural relationships with complex structures derived from hydrothermal fluids, using the natural materials as a motivation. This approach allows for the creation of chemically “pure” analogs of minerals from various chemical classes, including rock-forming oxides and silicates, as well as their recrystallized products. Thus, silicates of rare-earth metals, structurally related to minerals such as wadeite, fresnoite, apatite, olivine, and garnet, were obtained using high-temperature hydrothermal techniques. The synthetic analogues have found applications as phosphors, lasers, scintillators, and Faraday rotators due to their optical properties [2]. In our research into crystallization processes at medium temperatures, we successfully produced structural analogs of minerals such as mahnertite (NaCu3(AsO4)2Cl·5H2O) [3] and ellenbergerite (Mg6(Mg,Ti,Zr)2(Al,Mg)6Si8O28(OH)10) [4]. These synthetic phases exhibit unique magnetic properties, particularly low-dimensional magnetism arising from the presence of chains or layers of octahedra centered by open-shell metal atoms in their structures. Past reviews of synthetic mineral analogs [5,6] highlighted a broad range of compounds with ion-conducting properties. The synthetic equivalents of minerals such as triphylite (LiFePO4), niahite ((NH4)(Mn,Mg)(PO4)·H2O), simferite (LiMg(PO4)), tavorite (LiFe(PO4)(OH)), marićite (NaFe(PO4)), sarcopside ((Fe,Mn,Mg)3(PO4)2), and alluaudite ((Na,Ca)Mn(Fe,Mn,Fe,Mg)2(PO4)3) are actively utilized as electrode materials for sodium- or lithium-ion batteries.
Evidently, there is a substantial interest in the field of materials science not only in direct mineral analogs but also in determining their structural derivatives. By modifying the natural conditions of mineral formation in laboratory experiments by adjusting factors such as chemical reagents, temperature, pressure, and mineralizers, a wide array of mineralogically plausible compounds can be obtained. In this paper, we present the experimental results of crystallization in hydrothermal systems that mimic the expected hydrothermal conditions. We successfully synthesized three sodium-containing compounds: Na2CoSiO4 (I), a structural analog of mineral liberite (Li2BeSiO4) and the mineralogically reasonable phosphates Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl (II) and Na3V(OH)(HPO4)(PO4) (III). All investigated phases were previously obtained in the form of polycrystalline samples, and their crystal structures were studied using powder X-ray diffraction. The synthesis of these compounds in the form of single crystals allowed us to obtain more precise structural data, localize hydrogen atoms, refine their positions, and analyze hydrogen bonds. Moreover, studies have been carried out on the magnetic properties of V3+-containing phosphate.

2. Materials and Methods

2.1. Hydrothermal Synthesis

Single crystals were obtained in three different hydrothermal phosphate and silicate systems with cations of alkali and transition metals (Table 1). The initial chemically pure reagents were weighed, thoroughly ground in an agate mortar, and placed into an autoclave. In the silicate system, sodium hydroxide was added to improve the solubility of SiO2, raising the solution’s pH to 12. In the copper phosphate system, a small amount of hydrochloric acid solution was used to achieve an acidic medium with a pH of 1.5. In the phosphate system with vanadium, a small amount of lithium carbonate served as a mineralizer. Different temperature regimes were used for synthesizing these single crystals. For 270 °C middle-temperature hydrothermal conditions, the routine steel autoclave lined with fluoroplastic was utilized. However, 450 °C high-temperature hydrothermal experiments were conducted using a copper-coated high-pressure vessel of our own production made from a nickel–chromium alloy. It is critical for safety reasons to use superalloy autoclaves, as the creep rupture stress significantly enlarges with the increase in temperature. Distilled water was added according to the selected percentage of autoclave volume filling required to achieve the desired internal pressure. All experiments lasted for 10 days, after which period the autoclaves were left to cool naturally for 24 h. Crystallization products were washed multiple times with warm distilled water and air-dried. All three phases are shown in Table 1, where they are denoted as (I), (II), and (III), respectively.

2.2. X-ray Spectral Analysis

High-quality crystals selected via light microscopy were subjected to X-ray spectral analysis. Semi-quantitative ED analysis and crystal images were obtained from unpolished samples through graphite sputtering, using an accelerating voltage of 20 kV and a current rate of 10 nA, with a spectrum accumulation time of 50 s. These measurements were conducted at the Laboratory of Analytical Techniques of High Spatial Resolution, the Department of Petrology, the Faculty of Geology, Lomonosov Moscow State University, utilizing a JEOL JSM-6480LV scanning electron microscope with EDS-spectrometer. As reference standards, stoichiometric compounds and natural minerals were used, such as metallic Co and Cu, NaCl for Na and Cl, V2O3 for V, GaP for P, and natural garnet and diopside for Si and O, respectively.

2.3. Single-Crytsal X-ray Diffraction

The obtained crystals were investigated using X-ray diffraction at low temperatures for (I) and (II) and in standard conditions for (III), using MoKα radiation with a single-crystal diffractometer equipped with a CCD detector (Atlas-S2 & Sapphire3 detectors, Rigaku Oxford Diffraction, Tokyo, Japan). The datasets were corrected for background, the Lorentz and polarization effects, and absorption [7]. Most calculations for structural studies were performed using the WinGX program system [8]. The crystal structures were solved via direct methods and refined against the F2 data using the SHELX programs [9,10]. Structural data were deposited via the joint CCDC/FIZ Karlsruhe deposition service under the deposition numbers 2310942 (I), 2310943 (II), and 2310944 (III). Cif-data can be obtained free of charge from FIZ Karlsruhe via the following webpage: www.ccdc.cam.ac.uk/structures, accessed on 29 November 2023. Also, cif and checkcif files can be found in the Supplementary Materials to this paper.

3. Results and Discussion

3.1. Chemical Composition

The X-ray spectral analysis of blue lamellar crystals (I) revealed the presence of Na, Co, Si, and O atoms; the brown isometric well-faced crystals (II) contained Na, Cu, P, Cl, and O in a Na:Cu:P:Cl ratio of 2.2:3.5:2:1. The gray-green prismatic crystals in intergrowths of (III) contained Na, V, P, and O atoms with a Na:V:P ratio equal to 3:1:2.

3.2. Crystal Structure Solution

Crystal data and details of data collection and refinement are presented in Table 2. The pseudo-orthorhombic (I) Na2CoSiO4 was refined as pseudo-merohedral microtwin with a ratio of twin components equal to 0.32:0.68 for R1 = 0.025. During the solution of the crystal structure (II), three symmetrically independent positions of copper atoms were recognized. The coordination geometry for one of these positions (Cu3) indicated the presence of monovalent copper, while the other two positions (Cu1, Cu2) exhibited distorted Jahn–Teller polyhedra typical of Cu2+. The refinement revealed that the Cu3+ site is “diluted” by sodium atoms at a ratio of 0.8:0.2. The refinement resulted in the crystal–chemical formula Na2(Cu+0.8Na0.2)Cl[Cu2+3O(PO4)2] for (II). Data collection for the monoclinic phase (III) revealed non-merohedral twinning, connected by a pseudo-2-fold axis. A twin ratio between two components was refined at a ratio of 0.49:0.51. Two hydrogen atoms corresponding to two symmetrically independent OH groups were located via the electron density difference syntheses and refined based on the isotropic approximation. The final formula established through structural refinement (R1 = 0.043) was Na3V(OH)(HPO4)(PO4). The formulae refined during the X-ray diffraction structural study of the compounds were consistent with the results of the microprobe analysis.

3.3. Na2CoSiO4, a Structural Analog of the Mineral Liberite

3.3.1. The Structural Description and Analysis of Interatomic Distances

In the crystal structure of Na2CoSiO4, the cobalt and silicon tetrahedra share all vertices, forming the tetrahedral anionic framework [CoSiO4]2-, derived from the structure of β-tridymite, as illustrated in Figure 1a. The framework is negatively charged and balanced by sodium cations located within the channels. Within the bc plane, we can observe cellular layers consisting of six-membered rings composed of SiO4 and CoO4 tetrahedra alternating with each other (Figure 1b). Unlike the crystal structure of β-tridymite, these layers exhibit a uniform orientation of tetrahedra, described as UUUUUU (where ‘U’ signifies the upward orientation of the tetrahedra). These layers are interconnected along the a-axis through shared polyhedra vertices, maintaining the overall polarity of the structure.
In this monoclinic acentric crystal structure of Na2CoSiO4, all the atoms occupy general 2a Wyckoff positions. All Na, Co, and Si atoms are coordinated by oxygen ligands, forming tetrahedra (Figure 2). The Si–O bond lengths within the SiO4 tetrahedron range from 1.618(8) to 1.644(8) Å, with an average value of 1.63 Å (Table 3). The larger CoO4 tetrahedron exhibits Co–O distances in the interval 1.928(6)–1.969(5) Å (average 1.95 Å).
Sodium atoms occupy two symmetrically independent positions, both surrounded by four O ligands to form tetrahedra with Na1–O distances varying between 2.245(10) and 2.331(10) Å (average 2.28 Å). Meanwhile, the Na2–O distances are larger, ranging from 2.325(6) to 2.407(10) Å (average 2.36 Å). The next-to-nearest oxygen atoms associated with the Na1 and Na2 cations are situated at distances of 2.984(10) Å and 2.908(10) Å, respectively. Bond valence calculations show the impact of these next-to-nearest oxygens on the total valence of the sodium cations (Table 4).
Interestingly, the NaO4 polyhedra form a cationic framework that exhibits topological similarity to the anionic framework built from Si- and Co-centered tetrahedra (Figure 1b and Figure 3a). The three-periodic vertex-sharing of NaO4 tetrahedra and interconnected CoO4/PO4 tetrahedra leads to quasi-layering in (101) projection. The [CoSiO4] quasi-layers alternate in the [ 1 ¯ 01] direction with topologically equal quasi-layers of Na-centered polyhedra (Figure 3b). The electrochemical properties of Na2CoSiO4, which will be discussed below, are associated with the migration of Na+ ions through these quasi-layers parallel to the ( 1 ¯ 01) planes (Figure 3).

3.3.2. Na2CoSiO4 and the Liberite Structural Family of Ionic Conductors

The studied silicate, namely Na2CoSiO4, is a full structural analog of the mineral liberite, namely Li2BeSiO4, discovered in southern China [13]. Both liberite and disodium cobalt silicate crystallize in monoclinic or orthorhombic modifications, and both structures exhibit tetrahedral frameworks. Remarkably, the monoclinic polymorph of Na2CoSiO4 is entirely ordered, unlike the orthorhombic polymorph. In the orthorhombic structure Co and Si share the same positions, though the number of Na positions twice increased, and each site is a half populated.
Liberite and Na2CoSiO4 belong to a large group of compounds with the general formula A2MeXO4, where A+ represents the alkaline cations Li or Na; Me2+ represents Zn, Mn, Co, Mg, and Be; and X4+ represents the semimetals Si or Ge. The crystal structures of A2MeXO4 phases are based on a distorted hexagonal close-packing of oxygen atoms, with half of the tetrahedral voids occupied by the A, Me, and X atoms. The rich crystal chemistry of this family results from variations in the occupancy of the tetrahedral voids and the bonding of A+-, Me2+-, and X4+-centered tetrahedra. Accordingly, the A2MeXO4 compounds crystallize in several polymorphic modifications, mainly in the monoclinic and orthorhombic space groups P21, P21/n, Pn, Pmnb, Pmn21, and Pbn21 [5]. The good Na+/Li+ ionic conductivity of the A2MeXO4 family arises from its structural peculiarities. Channels within the anionic tetrahedral framework allow for the migration of alkali metal cations. In these compounds, the Na+/Li+ deintercalation is compensated by two electrons per Me2+ transition metal [14]. Moreover, the presence of transition metal atoms determines the magnetic properties of these phases due to the super-exchange magnetic interactions between Me2+ cations through the [XO4]4− anionic tetrahedra [15,16].
There was no clarity in the space group and unit cell parameters of the monoclinic polymorph of Na2CoSiO4, since earlier structural studies were conducted using polycrystalline Na2CoSiO4 samples synthesized through solid-state reactions [17,18]. These difficulties were apparently connected to the Na2CoSiO4 pseudo-orthorhombic symmetry of actually monoclinic crystals with β close to 90°, which exhibited a pseudo-merohedral twinning. Recently, by means of high-resolution X-ray powder diffraction using a synchrotron radiation source, a group of researchers [19] demonstrated the splitting of (10 1 ¯ ) and (101) reflections. Additionally, the absence of a small-angle (010) reflection undoubtedly confirmed the true monoclinic symmetry and the real value of the b parameter of the unit cell. In the course of our single-crystal structural study, it was supposed that the aparent orthorhombic unit cell of the sample was actually monoclinic, with the monoclinic angle being close to 90°. Accordingly, the crystal structure was refined in an anisotropic approximation of thermal vibration in the space group Pn as a two-component pseudomerohedric twin with a volume ratio of 32:68 and an obliquity of 0.01°. Our structural data, obtained via a low-temperature single-crystal X-ray experiment, correlate with the findings of powder synchrotron studies, thereby providing undoubtable confirmation of the acentric space group Pn inherent to the monoclinic modification of Na2CoSiO4.
Numerous studies have focused on the various physical properties of Na2CoSiO4 crystals, which exhibit good semiconductor properties and are promising for use in high-power Na-ion solar cells [20]. Na2CoSiO4 has also been investigated as a positive electrode material for sodium-ion capacitors. It showed excellent electrochemical performance, with a specific capacity of 42 F/g and a high specific energy of 12.4 Wh/kg (at a power density of 782.7 W/kg), and excellent cycling performance, retaining up to 84% of its capacity after 1500 charge–discharge cycles [18]. The results of [17] recommend the potential use of Na2CoSiO4 as a cathode material for sodium-ion batteries. Electrochemical measurements using cells with metallic sodium anodes revealed a reversible specific capacity of 100 mAh/g at an average discharge voltage of 3.3 V vs. Na/Na+. The atomistic modeling of Na+ diffusion demonstrated a low activation barrier and the presence of three-periodic diffusion pathways within the silicate framework, signifying favorable Na+ intercalation kinetics.

3.4. Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl with Oxo-Centered Pyroxene-like Chains

3.4.1. Analysis of Interatomic Distances and Description of the Crystal Structure

The key structural units of Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl include an orthophosphate tetrahedron, two Cu2+-centered polyhedra, and a Cu+-centered trigonal bipyramid (Figure 4a). The Cu12+O6 polyhedron with 2/m symmetry is a rare example of a compressed Jahn–Teller octahedron. There are four Cu1–O bonds with lengths equal to 2.254(2) Å and two shorter bonds of 1.851(1) Å (Table 5). The polyhedron distortion through the compressed axial oxygen atoms is associated with the requirement to avoid closer interaction between four neighboring Cu2O4Cl pyramids. The Cu22+O4Cl polyhedron is a distorted tetragonal pyramid, with the four Cu2–O distances at the base of the pyramid located in the range 1.911(2)–2.022(2) Å, and the Cl apex is located 2.579(1) Å from the Cu2.
The Cu+O2Cl3 polyhedron possesses mm2 symmetry and differs from the discussed Cu2+-centered polyhedra. The Cu3 position is 74% occupied by Cu+ cations and 26% occupied by Na+ cations. The Cu3-site atoms are coordinated by the two nearest O atoms, with the distance between the Cu3 and O2 of 2.032(2) Å and Cl anions, one at a distance of 2.394(5) Å and two at significantly larger distance of 3.1875(1) Å, resulting in trigonal bipyramidal coordination. The P–O bond lengths within the tetrahedron at the m plane range from 1.530(2) to 1.544(2) Å. The Na atom on the 2-fold axis is surrounded by eight ligands, i.e., six O anions with Na–O distances varying from 2.309 (1) to 2.886(2) Å and two Cl atoms at 3.270(2) Å (Figure 5a). This eight-fold coordination of the Na atoms is clearly confirmed via bond valence calculation (Table 6).
In the crystal structure of the compound Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl, Cu2+–centered polyhedra form ribbons, in which each CuO6 octahedron shares four edges with four CuO4Cl pyramids (Figure 4b). Likewise, each pyramid shares three edges, one edge with the neighboring CuO4Cl polyhedron and two edges with the CuO6 octahedra. The Cl vertex is common to two Cu2O4Cl polyhedra, and the oxygen ligand O4 is common to four Cu polyhedra (two CuO4Cl and two CuO6). Thus, two types of Cu2+-centered polyhedra form condensed zigzag ribbons elongated in the direction of the c-axis of the unit cell. These ribbons are arranged along the midpoints of the lateral faces of the unit cell in accordance with the base-centered lattice (oS). Orthophosphate tetrahedra connect these ribbons into a three-periodic heteropolyhedral framework (Figure 5). Similar ribbons are described in the crystal structures of the mineral aleutite ((M0.5Cl)[Cu5O2(AsO4)(VO4)]) [21] and the synthetic phase ((Na,Li)3(Cl,OH)[Cu3OAl(PO4)3]) [22].
Each PO4 group shares three oxygen vertices with Cu2+-centered polyhedra; hence, the PO4 tetrahedron connects two neighboring copper ribbons along the a-axis. The fourth oxygen vertex of the PO4 group is directed inside the framework channels and involved in the coordination of Na+ and Cu+ ions. The presence of these monovalent cations inside the large channels compensates for the negative charge of the heteropolyhedral framework.

3.4.2. Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl in the Series of Compounds with Oxo-Centered Pyroxene-like Chains

As mentioned above, the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl crystal structure is based on the chains of interconnected CuO6 octahedra and CuO4Cl pyramids, sharing edges. In these chains, neighboring Cu atoms are located at distances of 3.0264(3) Å (Cu1–Cu2) and 3.1382(6) Å (Cu2–Cu2) from each other. The oxygen ligand O4 in each of these chains is exclusively surrounded by copper atoms, forming the μ4 configuration.
According to the established nomenclature, copper oxo-salts with “additional” oxygen ligands (not shared with anionic complexes) are called oxo-cuprates. They are often described within the anion-centered approach [23,24]. In the case of the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl, oxo-centered [OCu4] tetrahedra form pyroxene-like chains [O2Cu6] (Figure 4c) that belong to the “zweier chain topological type”, according to the classification in [24]. In the structure (II), these chains, extended along the c-axis of the unit cell, are separated from each other by Na-, P-, and Cu+-centered polyhedra.
It is noteworthy that the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl is isostructural to the oxo-cuprate Na2Cu3OCu(PO4)2Cl, in the crystal structure of which there is no Cu+/Na+ isomorphous substitution [25]. Additionally, in the same family of electronically distinct copper phosphate chlorides, an oxo-cuprate arsenate Na2Cu3OCu(AsO4)2Cl was recently obtained [26]. This phase crystallizes in the lower symmetry space group Pnma, which is a subgroup of the Cmcm group inherent to phosphate.
Other structurally related phases, i.e., NaCu3OCu(PO4)2Cl [27] and Cu1,5Cu3O(PO4)2Cl [26], only contain divalent copper cations. All discussed structures are formed from the same ribbon fragments of Cu2+-centered polyhedra, interconnected through anionic tetrahedra. The differences between the structures lie in variations in the channel content within the topologically identical crystal frameworks of the composition [(Cu3O)(XO4)2]2−, where X denotes P or As. In the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl crystal structure, the channels contain salt inclusions of the |Na2(CuI0.8Na0.2)Cl|2⁺ composition. In the arsenate structure, there is a similar inclusion of|Na2CuICl|2⁺, whereas phases containing only divalent copper cations comprise |NaCuIICl|2⁺ and |CuII1.5Cl|2⁺. The channels contents differ in the amount of sodium cations, as well as the atomic positions of the interstitial ions. Notably, similar salt inclusions, often found in host–guest structures, are also observed in oxo-cuprates synthesized through halide flux methods, such as Na2Cu3OCu(PO4)2Cl [25] and NaCu3OCu(PO4)2Cl [27]. However, hydrothermal synthesis techniques are also suitable for the preparation of these compounds, as shown in [22,28].
It is remarkable that anion-centered oxo-cuprate chains ([O2Cu6]) (Figure 4c) form not only the crystal structure of (II) but also the other synthetic compounds mentioned above. These oxo-centered fragments ([O2Cu6]) are characteristic of minerals found in fumaroles, namely kamchatkite (KCu3OCl(SO4)2) [29,30], chloromenite (Cu9O2(SeO3)4Cl6) [31], vergasovaite (Cu3O[(Mo,S)O4SO4]) [32], dokuchaevite (Cu2[Cu6O2](VO4)3Cl3) [33], yaroshevskite (Cu3[Cu6O2][VO4]4Cl2) [34], and cupromolybdite ([O2Cu6][MoO4]4) [35]. Nevertheless, in the structures of minerals, tetrahedra forming oxo-centered [O2Cu6] chains have an orientation different to those in the structures of synthetic phases [26].

3.5. Mineralogically Probable Na3V(OH)(HPO4)(PO4)

3.5.1. Description of Crystal Structure and Analysis of Interatomic Distances

In the crystal structure of the Na3V(OH)(HPO4)(PO4) compound, V-centered octahedra share O2 vertices with the hydroxyl group to form chains of the 21 symmetry, parallel to the b axis of the unit cell. Each PO4 tetrahedron shares two O vertices adjacent to the chain VO4(OH)2 octahedra, resulting in the arrangement of one-periodic mixed-type anionic fragments of [V(OH)(HPO4)(PO4)]3−. These negatively charged ribbons interconnect in a framework via Na+ ions and hydrogen bonds (Figure 6). The hydrogen bonds system includes the atom O1, which serves as a donor to form an O1—H1···..O6 bond of medium strength, in accordance with the D···A distance (Table 7). The O1 atoms also acts as an acceptor in the significantly weaker hydrogen bond O2—H2···..O1 (Figure 6b).
The basic structural units of the Na3V(OH)(HPO4)(PO4) crystal structure are two symmetry-independent orthophosphate tetrahedra and a V3+O4(OH)2 octahedron, as shown in Figure 7. Both the phosphate tetrahedra with m symmetry exhibit significant distortion. The bond lengths in the P2O4 tetrahedron vary from 1.508(5) to 1.558(3) Å (Table 8). The shortened P2–O5 and P2–O7 bond lengths equal to 1.508(5) Å and 1.510(5) Å, respectively, serve to balance the valence charge on the corresponding oxygen atoms (Table 9). The P1O3(OH) tetrahedron is characterized by a notably longer P1–O1 distance of 1.582(5) Å to the oxygen atom of the hydroxyl group compared to other P1–O bond lengths ranging from 1.526(5) Å to 1.538(3) Å. Both the average P–O distances associated with P1 and P2 tetrahedra are equal to 1.534 and 1.544 Å, which are usual for the orthophosphate groups (Table 8). The V3+ ion centers the centrosymmetric V3+O4(OH)2 octahedron characterized by V–O bond distances within the range 1.966(3)–2.039(2) Å, with an average value of 2.01 Å, which is typical for V3+ cations. The structure also contains three symmetrically independent sodium atoms, forming centrosymmetric octahedra Na1O4(OH)2 and Na3O6, and a trigonal bipyramid Na2O5. In these polyhedral, the Na–O distances vary from 2.175(6) to 2.550(3) Å, while the Na1–OH values are expectedly larger and equal to 2.747(4) Å.
The crystal structure of the compound with the chemical formula Na3V(OH)(HPO4)(PO4) was first published in 2008 in [36], the authors of which used X-ray powder data in order to obtain the structure model. In our study based on a single-crystal X-ray diffraction experiment, the model was confirmed, and hydrogen atoms were also located and their positions were refined. Moreover, our structural study provides precise data regarding the geometry of PO4 tetrahedra, which allows us to exclude disorder in P-sites, as supposed in [36].
It is notable that the compound Na3V(OH)(HPO4)(PO4) has isostructural analogues with V3+ substituted for Al3+ or Ga3+ [37], which, together, form a group of compounds with the general formula Na3(M)(OH)(HPO4)(PO4), where M = V, Al, or Ga.
Similar crystal chemical properties of Fe3+ and V3+ ions determine a widespread distribution of vanadium in the magmatic process. Thus, iron is a kind of solvent of trivalent vanadium, causing its dispersed state in igneous rocks. In hydrothermal solutions in a reducing environment, vanadium migrates in the form of (V4+O)2+ or V3+ ions, which are mainly precipitated as part of hydrated silicate minerals, such as dimorphic varieties of cavansite and pentagonite, i.e., Ca(V4+O)Si4O10·4H2O. Though the (V4+O)2+ or V3+ phosphates are rare, among them are sincosite (Ca(V4+O)2(PO4)2·5H2O), and bariosincosite, (Ba(V4+O)2(PO4)2·4H2O), and a unique V3+ representative known as springcreekite (BaV3+3(PO4)(PO3OH)(OH)6). Both springcreekite and bariosincosite, found in a small vein deposit formerly mined for Cu, are considered to be formed via subaerial and near surface aqueous alteration at low temperatures [38,39]. It is remarkable that sincosite and bariosincosite synthetic analogues have been obtained in the laboratory under similar P/T conditions [40,41,42]. The geochemically reasonable system for Na3V3+(OH)(HPO4)(PO4) synthesis at 270 °C and 70 atm, as well as the discovery of natural V3+ phosphate springcreekite, indicate the mineralogical probability of this compound.

3.5.2. Magnetic Properties of Na3V(OH)(HPO4)(PO4)

V3+ has a 3d2 electronic configuration. The low value of spin S = 1 in combination with the reduced dimensionality of the magnetic subsystem can be a source of quantum behavior at low temperatures. Previously, the magnetic behavior of Na3V(OH)(HPO4)(PO4) was studied in the temperature range of 50–250 K. Magnetic susceptibility measurements evidenced sizable exchange interactions in the title compound. Therefore, it was interesting to study its behavior down to lower temperatures.
Magnetization M was measured using a Quantum Design automated PPMS-9T (Physical Property Measurement System) equipped with a Vibrating Sample Magnetometer (VSM). The temperature dependences of the magnetic susceptibility were taken in field-cooled (FC) and zero-field-cooled (ZFC) regimes at B = 0.1 T in the temperature range of 2–300 K. Through the temperature region, ZFC and FC curves almost perfectly coincided (Figure 8). At elevated temperatures, the χ(T) curves followed the Curie–Weiss law with the addition of the following temperature independent term χ0:
χ = χ 0 + C T Θ
The fitting of the FC curve in the range 200–300 K results in the Curie constant C = 0.99 emu K/mol and Weiss temperature Θ = −35 K. The value of C corresponds to the presence of V3+ ions in the structure using the g-factor g = 1.99. The negative Weiss temperature Θ suggests antiferromagnetic correlations between V3+ cations. The right inset in Figure 8 presents inverse magnetic susceptibility, showing deviation from linearity near 10 K. The extrapolation of χ−1(T) curve to negative temperatures indicates a slightly lower value of Weiss temperature, established via the approximation of the high-temperature region of the χ(T) curve.
The magnetization curve (left inset in Figure 8) reveals that at T = 2 K and in a magnetic field of B = 9 T, the magnetization is 0.78 μB. This value is significantly lower than that of the saturation magnetization, calculated using the following formula: Msat = ngSμB, which is equal to 2 μB.
According to the slightly elongated geometry of VO4(OH)2 polyhedra, two d electrons of V3+ occupy two out of three t2g orbitals. The established presence of antiferromagnetic exchange interactions between V-centered octahedra within chains is supported by the bridge angle V─O─V = 127.5° obtained from structural data. The absence of evidence for low-dimensional behavior in magnetic susceptibility measurements allows us to suggest that the values of exchange magnetic parameters of intrachain and interchain interactions are comparable in Na3V(OH)(HPO4)(PO4). Lower temperatures are needed to achieve a long-range order in the title compound.

4. Conclusions

In our search for the compounds with alkali and transition metals, three mineral-like phases were synthesized in the form of single crystals: silicate Na2CoSiO4, a structural analogue of liberite, and two phosphates, namely Na2Cu3O(Cu,Na)(PO4)2Cl and Na3V(OH)(HPO4)(PO4). Their crystal structures were refined (partially based on low temperatures) via X-ray diffraction, unlike some earlier studies performed using the powder samples. Our research, which generally confirmed previous results, allowed us to obtain more precise atomic coordinates, distances, and hydrogen bonds. It was shown that the silicate Na2CoSiO4 is characterized by a strongly pseudo-orthorhombic lattice and a monoclinic crystal structure, refined using a pseudomerohedric microtwin. It was established that in the crystal structure of Na2Cu3O(Cu,Na)(PO4)2Cl, the Cu+ position is partially substituted by 26% Na atoms, in contrast to previous structural data. We also highlighted the original crystal–chemical interpretation of the structural features and properties of Na2CoSiO4 and Na2Cu3O(Cu,Na)(PO4)2Cl and the studied magnetic behavior of Na3V(OH)(HPO4)(PO4).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min14010046/s1, cif and checkcif files.

Author Contributions

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

Funding

This work was supported by Moscow Lomonosov State University, Russian Federation (award no. AAAA-A16-116033010121-7). A. V. thanks the Russian Science Foundation, grant no. 23-73-10125, for providing financial support. The study by S.S. was carried out as part of the state assignment for the ISSP RAS.

Data Availability Statement

We deposited structural data via the joint CCDC/FIZ Karlsruhe deposition service under the deposition numbers 2310942 (I), 2310943 (II), and 2310944 (III). Cif-data, accessed on 29 November 2023, can be obtained free of charge from FIZ Karlsruhe via the following webpage: www.ccdc.cam.ac.uk/structures, accessed on 29 November 2023.

Acknowledgments

We thank N. N. Koshlyakova for carrying out the microprobe analysis of the crystals. We are grateful to N. V. Zubkova for providing assistance in collecting diffraction data, and we thank A. N. Vasiliev for consultations on magnetism.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; the collection, analyses, or interpretation of data; the writing of the manuscript; or the decision to publish the results.

References

  1. Morrison, G.; Abeysinghe, D.; Felder, J.B.; Egodawatte, S.; Ferreira, T.; zur Loye, H.-C. Where do new materials come from? Neither the stork nor the birds and the bees! In search of the next “first material”. J. South Carolina Acad. Sci. 2017, 15, 2. [Google Scholar]
  2. McMillen, C.D.; Kolis, J.W. Hydrothermal synthesis as a route to mineralogically-inspired structures. Dalton Trans. 2016, 45, 2772–2784. [Google Scholar] [CrossRef] [PubMed]
  3. Yakubovich, O.V.; Steele, I.M.; Kiriukhina, G.V.; Dimitrova, O.V. A microporous potassium vanadyl phosphate analogue of mahnertite: Hydrothermal synthesis and crystal structure. Z. Kristallog. Cryst. Mater. 2015, 230, 337–344. [Google Scholar] [CrossRef]
  4. Yakubovich, O.V.; Kiriukhina, G.V.; Dimitrova, O.V.; Shvanskaya, L.V.; Volkova, O.S.; Vasiliev, A.N. A novel cobalt sodium phosphate hydroxide with the ellenbergerite topology: Crystal structure and physical properties. Dalton Trans. 2015, 44, 11827–11834. [Google Scholar] [CrossRef]
  5. Masquelier, C.; Croguennec, L. Polyanionic (phosphates, silicates, sulfates) frameworks as electrode materials for rechargeable Li (or Na) batteries. Chem. Rev. 2013, 113, 6552–6591. [Google Scholar] [CrossRef]
  6. Yakubovich, O.V.; Khasanova, N.R.; Antipov, E.V. Mineral-Inspired Materials: Synthetic Phosphate Analogues for Battery Applications. Minerals 2020, 10, 524. [Google Scholar] [CrossRef]
  7. Agilent. CrysAlis PRO; Agilent Technologies Ltd.: Yarnton, UK, 2014. [Google Scholar]
  8. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  9. Sheldrick, G.M. SHELXT-Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  10. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar] [CrossRef]
  11. Parsons, S.; Flack, H.D.; Wagner, T. Use of intensity quotients and differences in absolute structure refinement. Acta Crystallogr. 2013, B69, 249–259. [Google Scholar] [CrossRef]
  12. Brown, I.D.; Altermatt, D. Bond-valence parameters obtained from a systematic analysis of the Inorganic Crystal Structure Database. Acta Crystallogr. 1985, 41, 244–247. [Google Scholar] [CrossRef]
  13. Han-ching, C. The crystal structure of liberite. Kexue Tongbao 1966, 17, 425–428. [Google Scholar]
  14. Armstrong, A.R.; Kuganathan, N.; Islam, M.S.; Bruce, P.G. Structure and lithium transport pathways in Li2FeSiO4 cathodes for lithium batteries. JACS 2011, 133, 13031–13035. [Google Scholar] [CrossRef] [PubMed]
  15. Avdeev, M.; Mohamed, Z.; Ling, C.D. Magnetic structures of βI-Li2CoSiO4 and γ0-Li2MnSiO4: Crystal structure type vs. magnetic topology. J. Solid State Chem. 2014, 216, 42–48. [Google Scholar] [CrossRef]
  16. Nalbandyan, V.B.; Zvereva, E.A.; Shukaev, I.L.; Gordon, E.; Politaev, V.V.; Whangbo, M.H.; Petrenko, A.A.; Denisov, R.S.; Markina, M.M.; Tzschoppe, M.; et al. A2MnXO4 Family (A = Li, Na, Ag; X = Si, Ge): Structural and Magnetic Properties. Inorg. Chem. 2017, 56, 14023–14039. [Google Scholar] [CrossRef] [PubMed]
  17. Treacher, J.C.; Wood, S.M.; Islam, M.S.; Kendrick, E. Na2CoSiO4 as a cathode material for sodium-ion batteries: Structure, electrochemistry and diffusion pathways. Phys. Chem. Chem. Phys. 2016, 18, 32744–32752. [Google Scholar] [CrossRef]
  18. Gao, S.; Zhao, J.; Zhao, Y.; Wu, Y.; Zhang, X.; Wang, L.; Liu, X.; Rui, Y.; Xu, J. Na2CoSiO4 as a novel positive electrode material for sodium-ion capacitors. Mater. Lett. 2015, 158, 300–303. [Google Scholar] [CrossRef]
  19. Wang, J.; Hoteling, G.; Shepard, R.; Wahila, M.; Wang, F.; Smeu, M.; Liu, H. Reaction Mechanism of Na-Ion Deintercalation in Na2CoSiO4. J. Phys. Chem. C 2022, 126, 16983–16992. [Google Scholar] [CrossRef]
  20. Trabelsi, K.; Karoui, K.; Jomni, F.; Rhaiem, A.B. Optical and AC conductivity behavior of sodium orthosilicate Na2CoSiO4. J. Alloys Compd. 2021, 867, 159099. [Google Scholar] [CrossRef]
  21. Siidra, O.I.; Nazarchuk, E.V.; Agakhanov, A.A.; Polekhovsky, Y.S. Aleutite [Cu5O2](AsO4)(VO4)·(Cu0.50.5)Cl, a new complex salt-inclusion mineral with Cu2+ substructure derived from Kagome-net. Mineral. Mag. 2019, 83, 847–853. [Google Scholar] [CrossRef]
  22. Yakubovich, O.V.; Kiriukhina, G.V.; Simonov, S.V.; Volkov, A.S.; Dimitrova, O.V. (Na,Li)3(Cl,OH)[Cu3OAl(PO4)3]: A first salt-inclusion aluminophosphate oxocuprate with a new type of crystal structure. Acta Crystallogr. 2023, B79, 24–31. [Google Scholar] [CrossRef] [PubMed]
  23. Krivovichev, S.V.; Filatov, S.K. Structural principles for minerals and inorganic compounds containing anion-centered tetrahedra. Am. Mineral. 1999, 84, 1099–1106. [Google Scholar] [CrossRef]
  24. Krivovichev, S.V.; Mentre, O.; Siidra, O.I.; Colmont, M.; Filatov, S.K. Anion-centered tetrahedra in inorganic compounds. Chem. Rev. 2013, 113, 6459–6535. [Google Scholar] [CrossRef] [PubMed]
  25. Etheredge, K.M.; Hwu, S.J. A novel copper (I/II) oxophosphate chloride with a quasi-one-dimensional μ4-oxo-bridged copper (II) chain. Crystal structure and magnetic properties of [Na2CuII3(PO4)2][CuIOCl]. Inorg. Chem. 1996, 35, 5278–5282. [Google Scholar] [CrossRef]
  26. Kornyakov, I.V.; Krivovichev, S.V. Na2Cu+[Cu2+3O](AsO4)2Cl and Cu3[Cu3O]2(PO4)4Cl2: Two new structure types based upon chains of oxocentered tetrahedra. Z. Kristallogr. Cryst. Mater. 2022, 237, 343–350. [Google Scholar] [CrossRef]
  27. Jin, T.T.; Liu, W.; Chen, S.; Prots, Y.; Schnelle, W.; Zhao, J.T.; Kniep, R.; Hoffmann, S. Crystal structure and magnetic properties of NaCuII[(CuII3O)(PO4)2Cl]. J. Solid State Chem. 2012, 192, 47–53. [Google Scholar] [CrossRef]
  28. Yakubovich, O.V.; Shvanskaya, L.V.; Kiriukhina, G.V.; Volkov, A.S.; Dimitrova, O.V.; Vasiliev, A.N. Hydrothermal synthesis and a composite crystal structure of Na6Cu7BiO4(PO4)4[Cl,(OH)]3 as a candidate for quantum spin liquid. Inorg. Chem. 2021, 60, 11450–11457. [Google Scholar] [CrossRef]
  29. Vergasova, L.P.; Filatov, S.K.; Seraphimova, Y.K.; Varaksina, T.V. Refractive indices of minerals and synthetic compounds. Zap. Vses. Mineral. Obshch. 1988, 117, 459–461. (In Russian) [Google Scholar]
  30. Varaksina, T.V.; Fundamensky, V.S.; Filatov, S.K.; Vergasova, L.P. The crystal structure of kamchatkite, a new naturally occurring oxychloride sulphate of potassium and copper. Mineral. Mag. 1990, 54, 613–616. [Google Scholar] [CrossRef]
  31. Vergasova, L.P.; Krivovichev, S.V.; Semenova, T.F.; Filatov, S.K.; Ananiev, V.V. Chloromenite, Cu9O2(SeO3)4Cl6, a new mineral from the Tolbachik volcano, Kamchatka, Russia. Eur. J. Mineral. 1999, 11, 119–123. [Google Scholar] [CrossRef]
  32. Berlepsch, P.; Armbruster, T.; Brugger, J.; Bykova, E.Y.; Kartashov, P.M. The crystal structure of vergasovaite Cu3O[(Mo,S)O4SO4], and its relation to synthetic Cu3O[MoO4]2. Eur. J. Mineral. 1999, 11, 101–110. [Google Scholar] [CrossRef]
  33. Siidra, O.I.; Nazarchuk, E.V.; Zaitsev, A.N.; Polekhovsky, Y.S.; Wenzel, T.; Spratt, J. Dokuchaevite, Cu8O2(VO4)3Cl3, a new mineral with remarkably diverse Cu2+ mixed-ligand coordination environments. Mineral. Mag. 2019, 83, 749–755. [Google Scholar] [CrossRef]
  34. Pekov, I.V.; Zubkova, N.V.; Zelenski, M.E.; Yapaskurt, V.O.; Polekhovsky, Y.S.; Fadeeva, O.A.; Pushcharovsky, D.Y. Yaroshevskite, Cu9O2(VO4)4Cl2, a new mineral from the Tolbachik volcano, Kamchatka, Russia. Mineral. Mag. 2013, 77, 107–116. [Google Scholar] [CrossRef]
  35. Zelenski, M.E.; Zubkova, N.V.; Pekov, I.V.; Polekhovsky, Y.S.; Pushcharovsky, D.Y. Cupromolybdite, Cu3O(MoO4)2, a new fumarolic mineral from the Tolbachik volcano, Kamchatka Peninsula, Russia. Eur. J. Mineral. 2012, 24, 749–757. [Google Scholar] [CrossRef]
  36. Ferdov, S.; Reis, M.S.; Lin, Z.; Ferreira, R.A.S. Hydrothermal synthesis, crystal structure, and magnetic properties of a new inorganic vanadium (III) phosphate with a chain structure. Inorg. Chem. 2008, 47, 10062–10066. [Google Scholar] [CrossRef] [PubMed]
  37. Lii, K.H.; Wang, S.L. Na3M(OH)(HPO4)(PO4), M = Al, Ga: Two phosphates with a chain structure. J. Solid State Chem. 1997, 128, 21–26. [Google Scholar] [CrossRef]
  38. Kolitsch, U.; Taylor, M.; Fallon, G.; Pring, A. Springcreekite, BaV3+3(PO4)2(OH,H2O)6, a new member of the crandallite group, from the Spring Creek mine, South Australia: The first natural V3+-member of the alunite family. Neues Jahrb. Miner. Monatsh. 1999, 1999, 529–544. [Google Scholar]
  39. Pring, A.; Kolitsch, U.; Birch, W.D.; Beyer, B.D.; Elliott, P.; Ayyappan, P.; Ramanan, A. Bariosincosite, a new hydrated barium vanadium phosphate, from the Spring Creek Mine, South Australia. Mineral. Mag. 1999, 63, 735–741. [Google Scholar] [CrossRef]
  40. Kang, H.Y.; Lee, W.C.; Wang, S.L.; Lii, K.H. Hydrothermal Synthesis and Structural Characterization of Four Layered Vanadyl(IV) Phosphate Hydrates A(VO)2(PO4)2·4H2O (A = Co, Ca, Sr, Pb). Inorg. Chem. 1992, 31, 4743–4748. [Google Scholar] [CrossRef]
  41. Roca, M.; Marcos, M.D.; Amoros, P.; Alamo, J.; Beltran-Porter, A.; Beltran-Porter, D. Synthesis and crystal structure of a novel lamellar barium derivative: Ba(VOPO4)2·4H2O. Synthetic pathways for layered oxovanadium phosphate hydrates M(VOPO4)2·nH2O. Inorg. Chem. 1997, 36, 3414–3421. [Google Scholar] [CrossRef]
  42. Franke, W.A.; Luger, P.; Weber, M.; Ivanova, T.I. Low hydrothermal growth of sincosite Ca(VO/PO4)2·4H2O. Zap. Vseross Miner. Obs. 1997, 126, 85–86. (In Russian) [Google Scholar]
Figure 1. The projections of the Na2CoSiO4 crystal structure on the ab (a) and bc (b) planes.
Figure 1. The projections of the Na2CoSiO4 crystal structure on the ab (a) and bc (b) planes.
Minerals 14 00046 g001
Figure 2. A symmetrically independent fragment of the crystal structure of Na2CoSiO4 (I); thermal ellipsoids are shown with 90% probability. Symmetry operations: (*) 0.5 + x, 1 − y, and −0.5 + z; (**) 0.5 + x, −y, and −0.5 + z; (***) 1 + x, y, and z.
Figure 2. A symmetrically independent fragment of the crystal structure of Na2CoSiO4 (I); thermal ellipsoids are shown with 90% probability. Symmetry operations: (*) 0.5 + x, 1 − y, and −0.5 + z; (**) 0.5 + x, −y, and −0.5 + z; (***) 1 + x, y, and z.
Minerals 14 00046 g002
Figure 3. (a) The quasi-layer of sharing O vertices’ NaO4 tetrahedra in the Na2CoSiO4 crystal structure and (b) their alteration along [ 1 ¯ 01] with topologically similar quasi-layers designed for Co- and Si-centered tetrahedra.
Figure 3. (a) The quasi-layer of sharing O vertices’ NaO4 tetrahedra in the Na2CoSiO4 crystal structure and (b) their alteration along [ 1 ¯ 01] with topologically similar quasi-layers designed for Co- and Si-centered tetrahedra.
Minerals 14 00046 g003
Figure 4. (a) The basic structural units of Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl with an atom-labeling scheme. Displacement ellipsoids are represented at the 90% probability level. Symmetry codes: (’) 2 − x, 1 − y, and 0.5 + z; (”) 2 − x, y, and z; (’’’) 2 − x, 1 − y, and 2 − z; (’’’’) x, 1 − y, and 2 − z; (*) x, y, and 2.5 − z; (**) 3 − x, y, and 2.5 − z; (***) 0.5 + x, y − 0.5, and z; (’*) 2.5 − x, 1.5 − y, and 2 − z; (”*) 2.5 − x, 1.5 − y, and 0.5 + z; (’’**) 2.5 − x, 1.5 − y, and −0.5 + z. (b) Corrugated chain of Cu2+-centered polyhedra in the classic representation and (c) the oxo-centered setting, showing [O2Cu6] chains.
Figure 4. (a) The basic structural units of Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl with an atom-labeling scheme. Displacement ellipsoids are represented at the 90% probability level. Symmetry codes: (’) 2 − x, 1 − y, and 0.5 + z; (”) 2 − x, y, and z; (’’’) 2 − x, 1 − y, and 2 − z; (’’’’) x, 1 − y, and 2 − z; (*) x, y, and 2.5 − z; (**) 3 − x, y, and 2.5 − z; (***) 0.5 + x, y − 0.5, and z; (’*) 2.5 − x, 1.5 − y, and 2 − z; (”*) 2.5 − x, 1.5 − y, and 0.5 + z; (’’**) 2.5 − x, 1.5 − y, and −0.5 + z. (b) Corrugated chain of Cu2+-centered polyhedra in the classic representation and (c) the oxo-centered setting, showing [O2Cu6] chains.
Minerals 14 00046 g004
Figure 5. The crystal structure of Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl displayed along the z (a) and x (b) axes.
Figure 5. The crystal structure of Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl displayed along the z (a) and x (b) axes.
Minerals 14 00046 g005
Figure 6. Ribbons of vanadium octahedra surrounded by phosphate tetrahedra (a) and their hydrogen bonding (b) in the Na3V(OH)(HPO4)(PO4) crystal structure projected onto the ab and ac planes.
Figure 6. Ribbons of vanadium octahedra surrounded by phosphate tetrahedra (a) and their hydrogen bonding (b) in the Na3V(OH)(HPO4)(PO4) crystal structure projected onto the ab and ac planes.
Minerals 14 00046 g006
Figure 7. An independent fragment of the crystal structure of Na3V(OH)(HPO4)(PO4) (III) with thermal ellipsoids represented at the 90% probability level. Symmetry operations: (*) x, 1 − y, z; (**) 1.5 − x, −0.5+y, −z; (***) 1.5 − x, 1.5 − y, −z.
Figure 7. An independent fragment of the crystal structure of Na3V(OH)(HPO4)(PO4) (III) with thermal ellipsoids represented at the 90% probability level. Symmetry operations: (*) x, 1 − y, z; (**) 1.5 − x, −0.5+y, −z; (***) 1.5 − x, 1.5 − y, −z.
Minerals 14 00046 g007
Figure 8. The temperature dependences of the magnetic susceptibility in Na3V(OH)(HPO4)(PO4) at B = 0.1 T taken in field-cooled (blue circles) and zero-field-cooled (white outlined circles) regimes. The dash red line illustrates the extrapolation of the Curie–Weiss law. The left inset presents the magnetization curve at T = 2 K, while the right inset shows the inverse magnetic susceptibility and its extrapolation to a negative temperature (dash red line).
Figure 8. The temperature dependences of the magnetic susceptibility in Na3V(OH)(HPO4)(PO4) at B = 0.1 T taken in field-cooled (blue circles) and zero-field-cooled (white outlined circles) regimes. The dash red line illustrates the extrapolation of the Curie–Weiss law. The left inset presents the magnetization curve at T = 2 K, while the right inset shows the inverse magnetic susceptibility and its extrapolation to a negative temperature (dash red line).
Minerals 14 00046 g008
Table 1. Secondary-electron SEM images showing the sample morphologies and experimental conditions of hydrothermal synthesis.
Table 1. Secondary-electron SEM images showing the sample morphologies and experimental conditions of hydrothermal synthesis.
(I)(II)(III)
Na2CoSiO4Na2(Cu+0.8Na0.2)Cl[Cu2+3O(PO4)2]Na3V(OH)(HPO4)(PO4)
Minerals 14 00046 i001Minerals 14 00046 i002Minerals 14 00046 i003
T 450 °C, P 500 atmT 450 °C, P 500 atmT 270 °C, P 70 atm
2 g CoCl2 (15.4 mmol)
1 g SiO2 (16.7 mmol)
NaOH solution
3 g CuCl2 (22.3 mmol)
1 g CuCl (10.1 mmol)
2 g (12.2 mmol) Na3PO4
HCl solution
1 g V2O3 (6.6 mmol)
2 g Na3PO4 (12 mmol)
0.2 g Li2CO3 (3 mmol)
H2O
pH 12pH 1.5pH 3
Copper-lined chrome-nickel autoclave, V 14 mL
Autoclave filling 50%
Copper-lined chrome-nickel autoclave, V 14 mL
Autoclave filling 50%
Steel autoclave lined with fluoroplastic, V 7 mL
Autoclave filling 72%
Table 2. The crystal data and experimental details of X-ray structural studies.
Table 2. The crystal data and experimental details of X-ray structural studies.
Structural Formula(I)(II)(III)
Na2CoSiO4Na2Cu3O(Cu0.8Na0.2)(PO4)2ClNa3V(OH)(HPO4)(PO4)
Mr197.00531.00327.87
Space group, ZPn, 2Cmcm, 4C2/m, 4
Temperature (K)170150293
Unit cell parameters a, b, c (Å)5.2271(5)13.6243(2)15.4157(10)
5.4198(4)10.3531(2)7.3107(4)
7.0466(6)6.3586(1)7.0556(4)
β (°)90.011(7)9096.702(6)
V (Å3)199.63(3)896.90(3)789.73 (8)
RadiationMo KαMo KαMo Kα
µ (mm−1)4.699.561.86
Crystal size (mm)0.15 × 0.06 × 0.020.17 × 0.11 × 0.070.19 × 0.09 × 0.06
DiffractometerOxford Diffraction GeminiXcalibur Sapphire3
Number of reflections: measured independently based on [I > 2σ(I)]1800, 909, 8948798, 732, 7271352, 1352, 1043
Rint0.0160.016-
(sin θ/λ)max−1)0.7030.7030.594
R [F2 > 2σ(F2)], wR(F2), S0.025, 0.064, 1.030.018, 0.046, 1.290.043, 0.109, 1.03
Number of refined parameters745791
Δρmax, Δρmin (e Å−3)0.46, −0.450.74, −0.910.64, −0.57
Flack parameter *0.01(3) *--
* The algorithm from [11] was used to determine the Flack parameter x.
Table 3. Characteristic distances, Å, in the Na2CoSiO4 crystal structure.
Table 3. Characteristic distances, Å, in the Na2CoSiO4 crystal structure.
Co–TetrahedronSi–Tetrahedron
Co–O21.928(6)Si1–O41.618(8)
–O31.944(5)–O31.627(4)
–O11.958(4)–O11.631(7)
–O41.969(5)–O21.644(8)
<Co–O>1.95<Si–O>1.63
Na1–TetrahedronNa2–Tetrahedron
Na1–O22.245(10)Na2–O22.325(6)
–O32.272(10)–O42.344(12)
–O42.279(5)–O12.364(11)
–O12.331(10)–O32.407(10)
<Na1–O>2.28<Na2–O>2.36
Table 4. Na2CoSiO4. Bond valence data *.
Table 4. Na2CoSiO4. Bond valence data *.
Na1Na2CoSiΣ
O10.203, 0.0520.1890.4870.9811.91
O20.2420.2050.5280.9471.92
O30.2290.173, 0.0610.5060.9921.96
O40.2260.1970.4731.0161.91
Σ0.950.831.993.94
* The algorithm and empirical parameters from [12] were used.
Table 5. The interatomic distances, Å, in the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl structure.
Table 5. The interatomic distances, Å, in the Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl structure.
Cu1 OctahedronCu2 Tetragonal PyramidCu3 * Trigonal Bipyramid
Cu1–O3 × 21.852(1)Cu2–O31.911(2)Cu3–O2 × 22.032(2)
–O4 × 42.254(2)–O11.921(2)–Cl2.394(5)
–O4 × 22.022(2)–Cl × 23.1875(1)
<Cu1–O>2.12–Cl12.5789(8)
<Cu2–O>1.97<Cu3–O>2.03
Na OctahedronP Tetrahedron
Na–O2 × 22.309(2)P–O21.532(2)
–O1 × 22.460(2)–O11.544(2)
–O4 × 2
–Cl × 2
2.884(2)
3.270(2)
–O4 × 21.544(2)
<Na–O>2.55<P–O>1.54
* Cu3-site is occupied by 74% Cu+ and 26% Na+.
Table 6. The bond valence calculations for Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl *.
Table 6. The bond valence calculations for Na2Cu3O(Cu0.8Na0.2)(PO4)2Cl *.
Cu1 Cu2 Cu3 **PNa
O1 0.520 1.2120.155 ×2↓ ×2→2.04
O2 0.182 ×2↓1.2510.212 ×2↓ ×2→1.98
O30.626 ×2↓ ×2→0.534 ×2→ 2.32
O40.212 ×4↓0.395 ×2↓ 1.213 ×2↓0.064×2↓1.88
Cl 0.209 ×2→0.144
0.02 ×2↓ ×2→
0.073 ×2↓ ×4→1.04
2.102.050.954.891.01
* Symbols ×2↓ and ×2→ indicate the multiplication of the valence contribution along the column or row according to the symmetry. ** Calculated as occupied by 74% Cu+ and 26% Na+.
Table 7. The geometric characteristics of hydrogen bonds in the Na3V(OH)(HPO4)(PO4) structure.
Table 7. The geometric characteristics of hydrogen bonds in the Na3V(OH)(HPO4)(PO4) structure.
D–H···AD–H, ÅH···A, ÅD···A, ÅD–H···A, °
O1–H1···O60.82(1)1.80(2)2.616(7)172(9)
O2–H2···O10.82(1)2.53(3)3.319(7)164(8)
Table 8. The interatomic distances (Å) in the crystal structure of Na3V(OH)(HPO4)(PO4).
Table 8. The interatomic distances (Å) in the crystal structure of Na3V(OH)(HPO4)(PO4).
V-Centered OctahedronP1–TetrahedronP2–Tetrahedron
V–O3
–O3
–O4
–O4
–O2
–O2
<V–O>
1.966(3)
1.966(3)
2.020(3)
2.020(3)
2.039(2)
2.039(2)
2.008
P1–O6
–O4
–O4
–O1
<P1–O>
1.526(5)
1.538(3)
1.538(3)
1.582(5)
1.544
P2–O5
–O7
–O3
–O3
<P2–O>
1.508(5)
1.510(5)
1.558(3)
1.558(3)
1.534
Na1–OctahedronNa2–PolyhedronNa3–Octahedron
Na1–O5
–O5
–O4
–O4
–O1
–O1
<Na1–O>
2.259(3)
2.259(3)
2.448(3)
2.448(3)
2.747(4)
2.747(4)
2.485
Na2–O7
–O5
–O2
–O4
–O4
<Na2–O>
2.175(6)
2.238(6)
2.435(5)
2.550(3)
2.550(3)
2.390
Na2–O3
–O3
–O6
–O6
–O7
–O7
<Na3–O>
2.427(3)
2.427(3)
2.460(4)
2.460(4)
2.482(4)
2.482(4)
2.456
Table 9. The bond valence data for Na3V(OH)(HPO4)(PO4).
Table 9. The bond valence data for Na3V(OH)(HPO4)(PO4).
Na1Na2Na3VP1P2H1H2Σ
O10.086 ×2↓ ×2→ 1.093 0.710.062.04
O2 0.163 0.449 ×2↓ ×2→ 0.942.00
O3 0.166 ×2↓0.547 ×2↓ 1.167 ×2↓ 1.88
O40.159 ×2↓0.129 ×2↓ 0.473 ×2↓1.231 ×2↓ 1.99
O50.235 ×2↓ ×2→0.246 1.335 2.05
O6 0.155 ×2↓ ×2→ 1.272 0.29 1.87
O7 0.280.148 ×2↓ ×2→ 1.328 1.90
Σ0.960.950.942.944.835.0011
The symbols ×2↓ and ×2→ indicate the multiplication of the valence contribution along a column or row in accordance with symmetry.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kiriukhina, G.; Yakubovich, O.; Verchenko, P.; Volkov, A.; Shvanskaya, L.; Dimitrova, O.; Simonov, S. Mineral-like Synthetic Compounds Stabilized under Hydrothermal Conditions: X-ray Diffraction Study and Comparative Crystal Chemistry. Minerals 2024, 14, 46. https://doi.org/10.3390/min14010046

AMA Style

Kiriukhina G, Yakubovich O, Verchenko P, Volkov A, Shvanskaya L, Dimitrova O, Simonov S. Mineral-like Synthetic Compounds Stabilized under Hydrothermal Conditions: X-ray Diffraction Study and Comparative Crystal Chemistry. Minerals. 2024; 14(1):46. https://doi.org/10.3390/min14010046

Chicago/Turabian Style

Kiriukhina, Galina, Olga Yakubovich, Polina Verchenko, Anatoly Volkov, Larisa Shvanskaya, Olga Dimitrova, and Sergey Simonov. 2024. "Mineral-like Synthetic Compounds Stabilized under Hydrothermal Conditions: X-ray Diffraction Study and Comparative Crystal Chemistry" Minerals 14, no. 1: 46. https://doi.org/10.3390/min14010046

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