Next Article in Journal
Experimental Study on the Effect of an Organic Matrix on Improving the Strength of Tailings Strengthened by MICP
Previous Article in Journal
A Double Cross-Linked Injectable Hydrogel Derived from Muscular Decellularized Matrix Promotes Myoblast Proliferation and Myogenic Differentiation
Previous Article in Special Issue
Synthesis, Characteristics, and Effect of Zinc Oxide and Silver Nanoparticles on the In Vitro Regeneration and Biochemical Profile of Chrysanthemum Adventitious Shoots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical Pressure Effect on the Stabilization of Rock-Salt ZnO—Lin−2MeOn−1 Solid Solutions Synthesized at High Pressure

by
Petr S. Sokolov
1,
Andrey N. Baranov
2 and
Vladimir L. Solozhenko
1,*
1
LSPM–CNRS, Université Sorbonne Paris Nord, 93430 Villetaneuse, France
2
Chemistry Department, Moscow State University, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(15), 5336; https://doi.org/10.3390/ma16155336
Submission received: 30 June 2023 / Revised: 24 July 2023 / Accepted: 27 July 2023 / Published: 29 July 2023

Abstract

:
Metastable ZnO—Lin−2MeOn−1 (Me = Sc3+, Ti4+, Ta5+) solid solutions with a rock-salt structure were synthesized through the solid-state reaction of ZnO with Lin−2Men+On−1 (n = 3, 4, 5) complex oxides at 7.7 GPa and 1300–1500 K. In all investigated systems, single-phase rock-salt solid solutions can be quenched down to ambient conditions in a wide (up to 80 mol% ZnO) concentration range. The phase composition, thermal stability, and thermal expansion of the recovered rock-salt solid solutions were studied by synchrotron powder X-ray diffraction. At ambient pressure, these solid solutions exhibit high thermal stability (up to 1000 K), with the decomposition temperature and decomposition products depending on the nature of the multiple charge cations.

1. Introduction

There are a number of crystal structures that have the potential to combine a wide variety of cations in a single compound over a wide range of concentrations. This is necessary in the design of catalytic materials or to control the properties of semiconductors and other functional materials. Examples of such universal structures are perovskites, garnets, spinels, etc. [1,2]. Also, as a universal platform for creating inorganic materials, we can consider the rock-salt structure, which is typical for alkali metal halides or binary ionic compounds such as Me2+O oxides. Rock-salt modification is also known for a semiconducting material such as zinc oxide. In this regard, zinc oxide and ZnO-based solid solutions are very promising objects [3,4]. Zinc oxide belongs to the family of wide band gap semiconductors with strong ionic character of chemical bonds [3,4]. Under ambient conditions, ZnO has a hexagonal wurtzite structure (P63mc, S.G. 186, w-ZnO), which changes to a cubic rock-salt structure (Fm-3m, S.G. 225, rs-ZnO) at pressures above 5 GPa [5,6,7,8,9].
The rock-salt structure, in contrast to the wurtzite one, allows the synthesis of ZnO solid solutions with a wide variety of cations. Previously, for Zn2+ ← Me2+ substitution, isovalent cations (Me2+ = Mg2+, Ni2+, Fe2+, Co2+, and Mn2+) with a tendency to octahedral environments were selected, which significantly affected the thermal stability of the formed solid solutions [8,9,10,11,12,13,14]. The most stable solid solutions with high ZnO content (up to 80 mol%) were those with magnesium and nickel oxides. For example, the cubic Ni0.2Zn0.8O is stable up to 970 K, whereas the cubic Co0.3Zn0.7O is stable only up to 670 K. The decomposition products are usually mixtures of two solid solutions, one with a cubic structure and another with a wurtzite structure [8,9,10].
In addition to binary solid solutions, ternary ZnO solid solutions were also synthesized by replacing two zinc cations with a combination of a single-charged lithium cation and a triple-charged metal cation, i.e., 2Zn2+ ← Li+ + Me3+. Extensive solid solutions have been reported for LiFeO2 [9,15], LiCrO2 [16] and LiTiO2 [9,15,17]. Limited cubic solid solutions in the ZnO–Li2TiO3 system have been described at ambient pressure [18]. As in the case of isovalent substitution, the concentration range and thermal stability of the synthesized solid solutions with the combination of Li+ + Me3+ essentially depend on the nature of the trivalent cation. Rock-salt (LiTiO2)0.2(ZnO)0.8 is stable up to 800 K, while (LiFeO2)0.2(ZnO)0.8 is stable only up to 670 K [15]. In fact, other examples of ternary cubic ZnO-based solid solutions are not found in the literature.
In the present work, in addition to the system with the trivalent cation scandium, a similar principle has been extended to the four- (3 Zn2+ ← 2 Li+ + Me4+) and five-valent (4 Zn2+ ← 3 Li+ + Me5+) cations, using titanium and tantalum as examples. All selected multivalent cations in the maximum oxidation state (Men+) form individual phases with lithium (Lin−2Men+On−1) although not in a rock-salt crystal lattice.

2. Experiment

The precursors listed in Table 1 were prepared by solid-state reactions of lithium carbonate (Li2CO3, Merck, min. 99%) with the corresponding metal oxide (reactive grade; a full list of the starting reagents is given in the Supplementary Materials). Due to lithium evaporation at high temperatures, Li2CO3 was taken in 5% excess. The fine powder mixtures were pressed into pellets and annealed in an alumina crucible in air at 970 K for 2 h. The product was reground and annealed again at 1070 K for 2 h. According to X-ray diffraction measurements (MZIII Seifert CuKα radiation), all of the as-synthesized precursors were single phase (see Figures S1 and S2 in the Supplementary Materials).
Mixtures of complex oxide and w-ZnO (Alfa Aesar, 99.99%) with a ZnO molar fraction (x) from 0.5 to 0.8 (with a 0.1 step) were thoroughly ground in an agate mortar with acetone, pressed into pellets and placed in gold capsules. High-pressure synthesis was performed in a toroid-type apparatus at the LSPM–CNRS. The experimental details and pressure–temperature calibration have been described elsewhere [8]. In general, the samples were gradually compressed up to 7.7 GPa, heated for 5 min at the desired temperature (1300–1500 K), then quenched by turning off the power supply and slowly decompressed. After disassembling the high-pressure cell, samples with high ZnO content (x > 0.7) usually crumbled into a white microcrystalline powder, while the recovered samples with lower ZnO content retained the tablet shape and looked like dense and well-sintered ceramics. The morphology of the recovered samples was studied by scanning electron microscopy Leo Supra 50VP (Carl Zeiss, Germany).
The high-pressure (up to 5 GPa)—high-temperature (up to 1400 K) behavior of the initial LiScO2 complex oxide was studied in situ by synchrotron energy-dispersive X-ray diffraction using the MAX80 multi-anvil system at F2.1 beamline of the DORIS III storage ring (HASYLAB-DESY); the details have been described elsewhere [13,14].
Thermal stability in air at temperatures up to 1200 K was studied by simultaneous thermogravimetric/differential thermal analysis using a Perkin Elmer Diamond Thermal Analyzer (TG/DTA) at a heating rate of 10 K/min.
In situ high-temperature (up to 1100 K) powder X-ray diffraction measurements at ambient pressure were performed at the B2 beamline of the DORIS III storage ring (HASYLAB-DESY) [33]. Debye–Scherrer geometry with a rotating fused silica capillary was used. Diffraction patterns were collected in the 2–70° 2θ-range for 2 min in real time using an OBI image plate detector [34]. A NIST powder LaB6 (Pm-3m, a = 4.15695 Å) was used as the standard sample for detector calibration. The furnace temperature was kept constant within 1 K using a Eurotherm temperature controller and Pt10%Rh–Pt thermocouple. The temperature step was 50 or 100 K. Before each data acquisition step, the sample temperature was stabilized for 2–3 min; the acquisition time was 2 min. The collected powder diffraction patterns were analyzed by the Le Bail method [35] using Powder Cell 2.4 [36] and DatLab 1.34 [37] software.

3. Results and Discussion

3.1. ZnO—LiScO2 System

The as-synthesized tetragonal (I41/amd) LiScO2 contained no impurities and had lattice parameters of a = 4.187(5) Å and c = 9.315(3) Å, which were in good agreement with the literature data [19,20]. It should be noted that the I41/amd structure is the only crystal form of LiScO2 that has been reported. This is a cation-ordered NaCl-like structure. The partial cation-ordering results in the primitive tetragonal body-centered unit cell of LiScO2, which can be described as a NaCl superstructure doubled along the c-axis. The coordination polyhedron for both lithium and scandium cations is a distorted octahedron, slightly elongated in the c-direction, making this structure similar to anatase [19,20].
In a special series of experiments at 5 GPa, the crystal structure of LiScO2 was studied in situ up to 1400 K using a MAX80 multi-anvil press and energy-dispersive X-ray diffraction. No melting or phase transition(s) of the pristine I41/amd tetragonal structure was observed in the investigated temperature range. The LiScO2 quenched from 7.7 GPa after a 5-min hold at 1370 K had the initial I41/amd structure with a small amount of cubic (Ia-3) Sc2O3 impurity; no other cubic phase(s) was observed (see Figure S1a).
Four compositions with x varying from 0.8 to 0.5 (in steps of 0.1) were synthesized in the ZnO—LiScO2 system. The results are summarized in Figure 1, Figure 2 and Figure 3 and Table 2, where the observed ranges of solid solutions are given in terms of x, the fraction of Zn2+ ions substituted by other cations. In all cases, the recovered samples were white insulating bulks. According to the SEM data, the synthesized solid solutions were dense, practically poreless ceramics with a grain size of about 10 microns.
Figure S3a–c shows the representative X-ray diffraction patterns of the synthesized rock-salt solid solution containing a small amount of Sc2O3 as an impurity. The sharp symmetrical narrow lines indicate that the solid solution is well crystallized. The dependence of the lattice parameters on the composition of the ZnO—LiScO2 system is shown in Figure 1. This dependence is nearly linear and follows Vegard’s law, as in the case of the related ZnO—LiTiO2 and ZnO—LiFeO2 systems [15].
A sequence of powder X-ray diffraction patterns for the rs-(LiScO2)1−x(ZnO)x solid solution composition with high ZnO content (x = 0.7) collected during stepwise heating up to 1100 K is shown in Figure 2. The first signs of change become visible at 873 K, when weak and broad reflections of the wurtzite phase appear, indicating the beginning of the decomposition of the initial single-phase cubic solid solution into a mixture of wurtzite solid solutions and Sc2O3 according to the following scheme:
rs-(LiScO2)1−x(ZnO)x → w-ZnO:Li + Sc2O3
where w-ZnO:Li is ZnO-Li2O solid solutions of different compositions.
The intensities of the lines corresponding to the cubic phase gradually decrease, and these lines disappear completely only at the highest temperatures reached in the experiment (~1100 K). At the same time, according to the simultaneous TG/DTA analysis, the weight loss and any thermal effects are not observed over the whole temperature range studied.
The temperature dependencies of the unit cell volumes of rs-(LiScO2)1−x(ZnO)x solid solutions are shown in Figure 3a. The experimental thermal expansion data were fitted to the following equation:
V(T) = V0[1 + α1(T − 98) + α2(T − 298)2]
the coefficients of which are given in Table 2. The unit for V(T) and V0 is Å3; the units for α1 and α2 are K−1 and K−2, respectively. The quadratic dependence of V(T) has been previously observed for ZnO-rich cubic solid solutions in other Li- containing systems [9,10,15], and α1 and α2 are close to the previously reported values [15].
The temperature dependencies of full width at half maximum (FWHM) for the 200 line (100% intensity) of rs-(LiScO2)1−x(ZnO)x solid solutions are shown in Figure 3b. In comparison, the most intense line 110 of the NIST standard LaB6 also has a similar (~0.05°) FWHM, indicating a high degree of crystallinity of the synthesized solid solutions. As the temperature increases, the FWHM of the 200 line of the cubic phase remains almost constant up to 700 K (Figure 3b). Then a sharp (2 to 3 times) increase in FWHM is observed, which begins 100–200 K below the temperature when the decomposition of the cubic phase becomes visible (i.e., the appearance of w-ZnO lines). Such FWHM behavior is associated with the segregation of microcrystalline nonequilibrium cubic solid solutions at the nanoscale with the formation of ZnO-enriched and -depleted solid solutions (pre-phase solid solution separation). Then the wurtzite zinc oxide and cubic scandium oxide phases crystallize from these intermediate solid solutions.
Thus, (LiScO2)1−x(ZnO)x cubic solid solutions of the composition 0.5 ≤ x ≤ 0.8 with high thermal stability at ambient pressure were synthesized for the first time at high pressures and high temperatures.

3.2. ZnO—Li2TiO3 System

The as-synthesized monoclinic (C2/c) Li2TiO3 contained no impurities (Figure S2) and had lattice parameters of a = 5.049(2) Å; b = 8.807(1) Å; and c = 9.715(6) Å, which were in good agreement with the literature data [22,23,24,25]. This compound is the only precursor known to have a high-temperature phase with a rock-salt structure (Fm-3m), in which Li+ and Ti4+ cations are completely disordered in their sublattice. At ambient pressure, the cubic phase has its stability field in the phase diagram at temperatures of 1430–1820 K, i.e., up to the melting point [21,22,23], and can be quenched under normal conditions. Another feature of Li2TiO3 is that the Ti4+ ion has the minimum size (0.60 Å) for the whole studied series of Men+ cations [8,9,12,15].
Eight single-phase cubic solid solutions with x varying from 0.8 to 0.1 (in steps of 0.1) were synthesized in the ZnO—Li2TiO3 system. The powder X-ray diffraction pattern of the solid solution with the maximum ZnO content, (Li2TiO3)0.2(ZnO)0.8, is shown in Figure S3d. The diffraction lines of all synthesized solid solutions are perfectly symmetric, the lines themselves are of high intensity; the FWHM of the 200 line depends on the composition as shown in Figure 4. The most ZnO-rich compositions (x = 0.8–0.7) have rather broad lines, whereas for compositions with x ≤ 0.6, the 200 line width remains almost constant and is equal to ~0.06° (which is close to the line width of the LaB6 standard), indicating their high crystallinity.
At room temperature, the dependence of the cell parameter of cubic solid solutions on the composition is nonlinear (Figure 5), with a negative deviation from Vegard’s law. This is indicative of the manifestation of the chemical pressure effect, when the multi-charged cation with a smaller radius produces an electrostatic lattice compression effect [38]. Since in this case, we know the endpoints of the concentration dependence of the a parameter, it is possible to quantify this effect from the difference in the lattice parameters and compressibility of rs-ZnO [39]. The chemical pressure calculated in this way was 3–5 GPa, depending on the composition of the solid solution. It should be noted that for the ZnO—LiScO2 system, such an effect is not observed due to the larger radius of the Sc3+ cation compared to Zn2+, and a smaller charge compared to Ti4+.
It was previously reported that Li2TiO3 and LiFeO2 form continuous NaCl-type solid solutions with MgO at ambient pressure above 1300 K, and a deviation from Vegard’s law was observed [18,40,41,42]. Due to the similarity of the ionic radii and chemical nature of the Zn2+ and Mg2+ cations, the related effects of incipient immiscibility and/or short-range ordering of cations [41] can occur in ZnO-based cubic solid solutions.
The mechanism of decomposition can be described by the following scheme:
rs-(Li2TiO3)1−x(ZnO)x → w-ZnO + rs-(Li2TiO3)1−y(ZnO)y
As the ZnO fraction in the cubic solid solution decreases, the onset temperature of decomposition (Td) increases from 870 K for x = 0.8 (Figure 6) to 1040 K for x = 0.5. Compositions with lower zinc oxide content are kinetically stable up to 1100 K, the maximum temperature available in our experiments. In some cases, small amounts of the C2/c monoclinic phase, characteristic of Li2TiO3 in this temperature range, have been observed in the X-ray diffraction patterns of decomposition products [25]. The FWHM values of the diffraction lines of all cubic solid solutions studied remain almost constant with increasing temperature.
The temperature dependencies of the unit cell volumes of rs-(Li2TiO3)1−x(ZnO)x solid solutions are shown in Figure 7.
The experimental thermal expansion data were fitted to the following equation
V(T) = V0[1 + α(T − 298)]
where the coefficients of which are given in Table 3. The unit of V(T) and V0 is Å3, and the unit of α is K−1. For these solid solutions, V(T) is linear or quasi-linear. For x < 0.6, the dependence of the thermal expansion coefficient on the composition is practically absent (Table 3, Figure S4), whereas for x ≥ 0.6, a tendency for its decrease with increasing zinc oxide content is observed. Note that extrapolation to 100% ZnO content (rs-ZnO) gives a value close to that previously reported by us [10].
Thus, for the first time, we have synthesized cubic ZnO—Li2TiO3 solid solutions in a wide concentration range (0.1 ≤ x ≤ 0.8), which are characterized by high thermal stability.

3.3. ZnO—Li3TaO4 System

The initial Li3TaO4 was metastable and had a cubic structure (Fm-3m) with lattice parameter a = 4.215(1) Å, which is in agreement with the literature data (a = 4.2207(4) Å [31] and a = 4.214(5) Å [32]). Quenched from 7.7 GPa and 1370 K, Li3TaO4 lost its cubic structure; the X-ray diffraction patterns before and after the high-pressure—high-temperature treatment are shown in Figure S1b.
Seven compositions with x varying from 0.8 to 0.3 (in steps of 0.1) were synthesized in the ZnO—Li3TaO4 system. The results are summarized in Figure 8, Figure 9 and Figure 10 and Figure S5 and Table 4, where the observed ranges of solid solutions are given in terms of x, the fraction of Zn2+ ions substituted by other cations. In all cases, the recovered samples were white insulating bulks.
The powder X-ray diffraction pattern of the (Li3TaO4)0.5(ZnO)0.5 solid solution is shown in Figure S3f, which indicates that it is cubic and single phase. The profiles of the diffraction lines are perfectly symmetric, and the lines themselves are of high intensity. Note that in the related system, MgO-Li3TaO4 solid solutions with a cubic structure do not form [43].
The diffraction pattern of the solid solution with nominal composition (Li3TaO4)0.2(ZnO)0.8 shows a more complex picture (Figure S3e): two groups of narrow and intense lines are seen, both of which can be described as structures belonging to the Fm-3m space group. The lattice parameter of the first (18 vol% content) is 4.278(2) Å, and the lattice parameter of the second is 4.2404(1) Å (82 vol% content) while other lines are absent. The presence of two cubic NaCl-like structures with slightly different lattice parameters may indicate the presence of micron-scale regions with different compositions. With the increasing temperature, the first group of lines gradually decreases in intensity and disappears completely at 620 K. The intensity of the second group of lines also decreases smoothly up to 1070 K. Previously, only rs-Ni1−xZnxO cubic solid solutions [10] have shown such high thermal stability.
For the initial composition (Li3TaO4)0.7(ZnO)0.3, the formation of a mixture of phases is observed, among which we can distinguish rock-salt and wurtzite phases. Also observed is a group of lines that cannot be described by any known structure but that are similar to the lines of the rs-Li3TaO4 decomposition products.
A negative deviation from Vegard’s law is observed for the dependence of the cell parameter of cubic solid solution on the composition (Figure 8). The calculated values of chemical pressure vary in the range of 3–5 GPa, depending on the composition of the solid solution.
For the compositions (Li3TaO4)0.3(ZnO)0.7, (Li3TaO4)0.4 (ZnO)0.6, and (Li3TaO4)0.5(ZnO)0.5, the cubic structure is stable up to about 1000 K (Figure 9). At ~1100 K, numerous new lines of low intensity appear, which we could not interpret.
The mechanism of decomposition of cubic solid solutions in this case is as follows:
rs-(Li3TaO4)1−x(ZnO)xw-ZnO + rs-(Li3TaO4)1−y(ZnO)y + phase(s) X
The temperature dependencies of the unit cell volumes of rs-(Li3TaO4)1−x(ZnO)x solid solutions are shown in Figure 10. The experimental thermal expansion data were fitted to Equation (4), the coefficients of which are given in Table 4.
Thus, we have demonstrated for the first time that synthesis at high pressure and high temperature allows for single-phase cubic solid solutions in the ZnO—Li3TaO4 system to be obtained in a relatively narrow concentration range (0.5 ≤ x ≤ 0.8). The synthesized solid solutions exhibit high thermal and phase stability.

3.4. General Remarks

The concentration dependence of the decomposition temperature of metastable ZnO—Lin−2MeOn−1 solid solutions (Men+ = Sc3+, Ti4+, and Ta5+) at ambient pressure for all investigated systems is shown in Figure 11. It is easy to see that the thermal stability increases with the cation charge (n). Such behavior is apparently related to the difficulty of diffusion of multivalent ions in the cubic lattice.
It is known that there are several factors that determine the kinetic stability of metastable phases quenched from high pressure and/or high temperature. In the case of the pure cubic ZnO phase, it is the kinetic stabilization of the small grains of the high-pressure phase that prevents the appearance of wurtzite phase nuclei in a small volume [7]. High-entropy multicomponent solid solutions whose stabilization is determined by high values of configurational entropy have been described in the literature [44,45]. For example, an equimolar mixture of MgO, CoO, NiO, CuO, and ZnO oxides can be stabilized in the rock-salt structure, although two of these oxides (CuO and ZnO) do not form cubic polymorphs as individual oxides. The necessary condition for this type of stabilization is the same cation charge. In the case of the cubic solid solutions considered in the present work, these factors either do not work or their contribution can be estimated to be negligible [46].
In multiphase systems, one of the components can play the role of a “shell”, preventing the reverse phase transition of the metastable phase when the pressure is released. Such examples are described for the MgO-ZnO [47] and NaCl-ZnO [48] systems. Stability in such systems is provided either by the similar structure (MgO) or mechanical properties (NaCl) of the second component.
The octahedral environment preference factor [49] previously reported in [9,10,15] cannot be a determining factor in the case of ZnO solid solutions with Li2TiO3 and Li3TaO4. Although lithium tends to an octahedral environment, its radius (0.76 Å) is slightly larger than that of Zn2+ (0.74 Å) and the charge is half as low, so the influence of Li+ on the stability of the cubic solid solution cannot be significant. However, it should be noted that the strong ionicity of the Li-O bond (lithium has the lowest electronegativity value in the series of elements considered) is a favorable factor.
In the case of ternary systems, the stabilization of the metastable cubic ZnO could be explained by the presence of a multi-charged cation in the solid solution lattice, i.e., by the fact that a cation with a smaller radius but a higher charge (see Table 1) “compresses” the crystal lattice (the so-called “chemical pressure” effect), which is accompanied by a kinetic stabilization of the cubic structure in the form of solid solution. According to the deviation of the lattice parameters of solid solutions from Vegard’s law, the chemical pressure for titanium Ti4+ and tantalum Ta5+ in the ZnO-Li2TiO3 and ZnO-Li3TaO4 systems was estimated to be 3–5 GPa.
Thus, the metastable ZnO-Lin−1MeOn−2 solid solutions in Figure 11 are completely consistent with the charge of the Men+ cation. Solid solutions with Sc3+, where the effect of chemical compression on the concentration dependence of the lattice parameter is not observed, exhibit the minimum thermal stability, and solid solutions with Ta5+ show the maximum.
We believe that some solid solutions with a rock-salt crystal structure, synthesized for the first time in the present work, may be of interest as promising Li-containing materials [50,51] or as materials with high dielectric characteristics [52,53].
The described approach of stabilizing metastable high-pressure phases under normal conditions by forming extended substitution solid solutions may also be promising for a number of other oxide and non-oxide (ZnS-LiGaS2, ZnS-Li2TiS3 and ZnS-Li3NbS4 [54,55,56]) systems whose high-pressure phases may have interesting functional properties. However, this hypothesis certainly requires experimental verification.

4. Conclusions

For the synthesized rock-salt ZnO—Lin−2MeOn−1 (Me = Sc3+, Ti4+, Ta5+) solid solutions, the limits of concentration (x up to 0.8) and thermal (up to 1000 K) stabilities were established. Coefficients of thermal expansion were determined for all the systems studied, showing either linear or quadratic dependence on temperature. The decomposition temperature of metastable solid solutions depends on the composition and nature of the multivalent cation, which allows us to define it as a chemical pressure effect.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16155336/s1.

Author Contributions

Conceptualization, V.L.S. and A.N.B.; methodology, A.N.B. and V.L.S.; investigation, P.S.S., A.N.B. and V.L.S.; formal analysis, P.S.S. and A.N.B.; data curation, P.S.S., A.N.B. and V.L.S.; validation, A.N.B. and V.L.S.; resources, V.L.S.; writing—original draft preparation, A.N.B. and P.S.S.; writing—review and editing, V.L.S.; visualization, P.S.S.; supervision, V.L.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request.

Acknowledgments

The synchrotron X-ray diffraction experiments at HASYLAB-DESY were carried out during beamtime allocated to I-20070033 EC and I-20100234 EC projects.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. West, A.R. Solid State Chemistry and Its Applications; John Wiley & Sons: Hoboken, NJ, USA, 1991. [Google Scholar]
  2. Rao, C.N.R. Chemical Approaches to Synthesis of Inorganic Materials; John Wiley & Sons: Hoboken, NJ, USA, 1995. [Google Scholar]
  3. Kołodziejczak-Radzimska, A.; Jesionowski, T. Zinc oxide—From synthesis to application: A review. Materials 2014, 7, 2833–2881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Janotti, A.; Van de Walle, C.G. Fundamentals of zinc oxide as a semiconductor. Rep. Prog. Phys. 2009, 72, 126501. [Google Scholar] [CrossRef] [Green Version]
  5. Solozhenko, V.L.; Kurakevych, O.O.; Sokolov, P.S.; Baranov, A.N. Kinetics of the wurtzite to rock-salt phase transformation in ZnO at high pressure. J. Phys. Chem. A 2011, 115, 4354–4358. [Google Scholar] [CrossRef] [PubMed]
  6. Kusaba, K.; Syono, Y.; Kikegawa, T. Phase transition of ZnO under high pressure and temperature. Proc. Jpn. Acad. Ser. B 1999, 75, 1–6. [Google Scholar] [CrossRef] [Green Version]
  7. Baranov, A.N.; Sokolov, P.S.; Tafeenko, V.A.; Lathe, C.; Zubavichus, Y.V.; Veligzhanin, A.A.; Chukichev, M.V.; Solozhenko, V.L. Nanocrystallinity as a route to metastable phases: Rock salt ZnO. Chem. Mater. 2013, 25, 1775–1782. [Google Scholar] [CrossRef] [Green Version]
  8. Baranov, A.N.; Sokolov, P.S.; Kurakevich, O.O.; Tafeenko, V.A.; Trots, D.; Solozhenko, V.L. Synthesis of rock-salt MeO–ZnO solid solutions (Me=Ni2+, Co2+, Fe2+, Mn2+) at high pressure and high temperature. High Press. Res. 2008, 28, 515–519. [Google Scholar] [CrossRef] [Green Version]
  9. Baranov, A.N.; Sokolov, P.S.; Solozhenko, V.L. ZnO under pressure: From nanoparticles to single crystals. Crystals 2022, 12, 744. [Google Scholar] [CrossRef]
  10. Sokolov, P.S.; Baranov, A.N.; Solozhenko, V.L. Phase stability and thermal expansion of ZnO solid solutions with 3d transition metal oxides synthesized at high pressure. J. Phys. Chem. Solids 2023, 180, 111437. [Google Scholar] [CrossRef]
  11. Baranov, A.N.; Solozhenko, V.L.; Chateau, C.; Bocquillon, G.; Petitet, J.P.; Panin, G.N.; Kang, T.W.; Shpanchenko, R.V.; Antipov, E.V.; Oh, Y.J. Cubic MgxZn1–xO wide band gap solid solutions synthesized at high pressures. J. Phys. Cond. Matter. 2005, 7, 3377–3384. [Google Scholar] [CrossRef]
  12. Solozhenko, V.L.; Baranov, A.N.; Turkevich, V.Z. High-pressure formation of MgxZn1−xO solid solutions with rock salt structure. Solid State Comm. 2006, 138, 534–537. [Google Scholar] [CrossRef] [Green Version]
  13. Sokolov, P.S.; Baranov, A.N.; Lathe, C.; Solozhenko, V.L. High-pressure synthesis of FeO–ZnO solid solutions with rock salt structure: In situ X-ray diffraction studies. High Press. Res. 2010, 30, 39–43. [Google Scholar] [CrossRef] [Green Version]
  14. Sokolov, P.S.; Baranov, A.N.; Lathe, C.; Turkevich, V.Z.; Solozhenko, V.L. High-pressure synthesis of MnO–ZnO solid solutions with rock salt structure: In situ X-ray diffraction studies. High Press. Res. 2011, 31, 43–47. [Google Scholar] [CrossRef] [Green Version]
  15. Sokolov, P.S.; Baranov, A.N.; Tafeenko, V.A.; Solozhenko, V.L. High pressure synthesis of LiMeO2–ZnO (Me=Fe3+, Ti3+) solid solutions with a rock salt structure. High Press. Res. 2011, 31, 304–309. [Google Scholar] [CrossRef] [Green Version]
  16. Hanck, K.W.; Laitimen, H.A. Structural and thermal stability studies of LiZn2CrO4 and Co2CrO4. J. Inorg. Nucl. Chem. 1970, 33, 63–73. [Google Scholar] [CrossRef]
  17. Brixner, L.H. Preparation, structure and electrical properties of some substituted lithium-oxo-metallates. J. Inorg. Nucl. Chem. 1960, 16, 162–163. [Google Scholar] [CrossRef]
  18. Hernandez, V.S.; Martinez, L.M.T.; Mather, G.C.; West, A.R. Stoichiometry, structures and polymorphism of spinel-like phases, Li1.33xZn2−2xTi1+0.67xO4. J. Mater. Chem. 1996, 6, 1533–1536. [Google Scholar] [CrossRef]
  19. Hewston, T.A.; Chamberland, B.L. A survey of first-row ternary oxides LiMO2 (M = Sc-Cu). J. Phys. Chem. Solids 1987, 48, 97–108. [Google Scholar] [CrossRef]
  20. Mather, G.C.; Dussarrat, C.; Etourneau, J.; West, A.R. A review of cation-ordered rock salt superstructure oxides. J. Mater. Chem. 2000, 10, 2219–2230. [Google Scholar] [CrossRef]
  21. Kleykamp, H. Enthalpy, heat capacity and enthalpy of transformation of Li2TiO3. J. Inorg. Nucl. Mat. 2001, 295, 244–248. [Google Scholar] [CrossRef]
  22. Izquierdo, G.; West, A.R. Phase equilibria in the system Li2O-TiO2. Mater. Res. Bull. 1980, 14, 1655–1660. [Google Scholar] [CrossRef]
  23. Campos, A.; Quintana, P.; West, A.R. Order-disorder in rock salt-like phases and solid solutions, Li2(Ti1−xZrx)O3. J. Solid State Chem. 1990, 86, 129–130. [Google Scholar] [CrossRef]
  24. Laumann, A.; Fehr, K.T.; Wachsmann, M.; Holzapfel, M.; Iversen, B.B. Metastable formation of low temperature cubic Li2TiO3 under hydrothermal conditions—Its stability and structural properties. Solid State Ionics. 2010, 181, 1525–1529. [Google Scholar] [CrossRef]
  25. Laumann, A.; Fehr, K.T.; Boysen, H.; Hozel, M.; Holzapfel, M. Temperature-dependent structural transformations of hydrothermally synthesized cubic Li2TiO3 studied by in-situ neutron diffraction. Z. Kristallogr. 2011, 226, 53–61. [Google Scholar] [CrossRef]
  26. Du Boulay, D.; Sakaguchi, A.; Suda, K.; Ishizawa, N. Reinvestigation of β-Li3TaO4. Acta Cryst. E 2003, 59, i80–i82. [Google Scholar] [CrossRef] [Green Version]
  27. Zocchi, M.; Gatti, M.; Santoro, A.; Roth, R.S. Neutron and X-ray diffraction study on polymorphism in lithium orthotantalate, Li3TaO4. J. Solid State Chem. 1983, 48, 420–430. [Google Scholar] [CrossRef]
  28. Nyman, M.; Anderson, T.M.; Provencio, P.P. Comparison of aqueous and non-aqueous soft-chemical syntheses of lithium niobate and lithium tantalate powders. Cryst. Growth Des. 2009, 9, 1036–1040. [Google Scholar] [CrossRef]
  29. Blasse, G. On the structure of some compounds Li3Me5+O4, and some other mixed metal oxides containing lithium. Z. Anorg. Allg. Chem. 1964, 331, 44–50. [Google Scholar] [CrossRef]
  30. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  31. Kim, C.; Pham, T.L.; Lee, J.-S.; Kim, Y.-I. Synthesis, thermal analysis, and band gap of ordered and disordered complex rock salt Li3TaO4. J. Solid State Chem. 2022, 315, 123450. [Google Scholar] [CrossRef]
  32. Grenier, J.-C.; Martin, C.; Durif, A. Étude cristallographique des orthoniobates et orthotantalates de lithium. Bull. Soc. Franç. Minér. Crist. 1964, 87, 316–320. [Google Scholar] [CrossRef]
  33. Knapp, M.; Baehtz, C.; Ehrenberg, H.; Fuess, H. The synchrotron powder diffractometer at beamline B2 at HASYLAB/DESY: Status and capabilities. J. Synchrotron. Rad. 2004, 11, 328–334. [Google Scholar] [CrossRef]
  34. Knapp, M.; Joco, V.; Baehtz, C.; Brecht, H.H.; Berghäuser, A.; Ehrenberg, H.; von Seggern, H.; Fuess, H. Position-sensitive detector system OBI for high resolution X-ray powder diffraction using on-site readable image plates. Nucl. Instr. Meth. Phys. Res. A 2004, 521, 565–570. [Google Scholar] [CrossRef]
  35. Le Bail, A.; Duroy, H.; Fourquet, J.L. Ab-initio structure determination of LiSbWO6 by X-ray powder diffraction. Mat. Res. Bull. 1988, 23, 447–452. [Google Scholar] [CrossRef]
  36. Nolze, G.; Kraus, W. PowderCell 2.0 for Windows. Powder Diffr. 1998, 13, 256–259. [Google Scholar]
  37. Syassen, K. Computer Code DATLAB 1.34; Max Planck Institute: Stuttgart, Germany, 2012. [Google Scholar]
  38. Lin, K.; Li, Q.; Yu, R.; Chen, J.; Attfield, J.P.; Xing, X. Chemical pressure in functional materials. Chem. Soc. Rev. 2022, 51, 5351–5364. [Google Scholar] [CrossRef]
  39. Decremps, F.; Datchi, F.; Saitta, A.M.; Polian, A.; Pascarelli, S.; Di Cicco, A.; Itié, J.P.; Baudelet, F. Local structure of condensed zinc oxide. Phys. Rev. B 2003, 68, 104101. [Google Scholar] [CrossRef] [Green Version]
  40. Castellanos, M.; West, A.R. Order-disorder phenomena in oxides with rock salt structures: The system Li2TiO3-MgO. J. Mater. Sci. 1979, 14, 450. [Google Scholar] [CrossRef]
  41. Castellanos, M.; West, A.R. Deviations from Vegard’s law in oxide solid solutions. The systems Li2TiO3–MgO and Li2TiO3–Na2TiO3. J. Chem. Soc. Faraday Trans. 1980, 76, 2159–2169. [Google Scholar] [CrossRef]
  42. Adam, D.B.; Brindley, G.W. Cation solutions in crystalline MgO of the Type Mg1−x(R+, Rn+)xO. Mat. Sci. Eng. 1969, 3, 252–254. [Google Scholar] [CrossRef]
  43. Castellanos, M.; Gard, J.A.; West, A.R. Crystal data for a new family of phases, Li3Mg2XO6:X=Nb, Ta, Sb. J. Appl. Cryst. 1982, 15, 116–119. [Google Scholar] [CrossRef]
  44. Rost, C.M.; Sachet, E.; Borman, T.; Moballegh, A.; Dickey, E.C.; Hou, D.; Jones, J.L.; Curtarolo, S.; Maria, J.-P. Entropy-stabilized oxides. Nat. Commun. 2015, 6, 8485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sarkar, A.; Djenadic, R.; Usgarani, N.J.; Sanghvi, K.P.; Chakravadhanula, V.S.K.; Gandhi, A.S.; Hahn, H.; Bhattacharya, S.S. Nanocrystalline multicomponent entropy stabilised transition metal oxides. J. Eur. Ceram. Soc. 2017, 37, 747–754. [Google Scholar] [CrossRef]
  46. Fracchia, M.; Coduri, M.; Manzoli, M.; Ghigna, P.; Tamburini, U.A. Is configurational entropy the main stabilizing term in rock-salt Mg0.2Co0.2Ni0.2Cu0.2Zn0.2O high entropy oxide? Nat. Commun. 2022, 13, 2977. [Google Scholar] [CrossRef] [PubMed]
  47. Baranov, A.N.; Kurakevych, O.O.; Tafeenko, V.A.; Sokolov, P.S.; Panin, G.N.; Solozhenko, V.L. High-pressure synthesis and luminescent properties of cubic ZnO/MgO nanocomposites. J. Appl. Phys. 2010, 107, 073519. [Google Scholar] [CrossRef] [Green Version]
  48. Sokolov, P.S.; Baranov, A.N.; Dobrokhotova, Z.V.; Solozhenko, V.L. Synthesis and thermal stability of cubic ZnO in the salt nanocomposites. Russ. Chem. Bull. 2010, 59, 325–328. [Google Scholar] [CrossRef] [Green Version]
  49. Urusov, S.V. Interaction of cations on octahedral and tetrahedral sites in simple spinels: A reply. Phys. Chem. Miner. 1984, 10, 194–195. [Google Scholar] [CrossRef]
  50. Li, B.; Sougrati, S.T.; Rouse, G.; Morozov, A.V.; Dedryvere, R.; Ladecola, A.; Senyshyn, A.; Zhang, L.; Abakumov, A.M.; Doublet, M.L.; et al. Correlating ligand-to-metal charge transfer with voltage hysteresis in a Li-rich rock-salt compound exhibiting anionic redox. Nat. Chem. 2021, 13, 1070–1108. [Google Scholar] [CrossRef] [PubMed]
  51. Lun, Z.; Ouyang, B.; Kwon, D.H.; Ha, Y.; Foley, E.E.; Huang, T.-Y.; Cai, H.; Kim, H.; Balasubramanian, M.; Sun, Y.; et al. Cation-disordered rocksalt-type high-entropy cathodes for Li-ion batteries. Nat. Mater. 2021, 20, 214–221. [Google Scholar] [CrossRef]
  52. Yuan, L.L.; Bian, J.J. Microwave Dielectric Properties of the Lithium Containing Compounds with Rock Salt Structure. Ferroelectrics 2009, 387, 123–129. [Google Scholar] [CrossRef]
  53. Uyama, T.; Mukai, K.; Yamada, I. High-Pressure Synthesis of Cation-Disordered Rock-Salt Oxyfluorides with High Crystallinity. Electrochemistry 2021, 89, 94–99. [Google Scholar] [CrossRef]
  54. Chen, M.-M.; Xue, H.-G.; Guo, S.-P. Multinary metal chalcogenides with tetrahedral structures for second-order nonlinear optical, photocatalytic, and photovoltaic applications. Coordin. Chem. Rev. 2018, 368, 115–133. [Google Scholar] [CrossRef]
  55. Sakuda, A.; Takeuchi, T.; Okamura, K.; Kobayashi, H.; Sakaebe, H.; Tatsumi, K.; Ogumi, Z. Rock-salt-type lithium metal sulphides as novel positive-electrode materials. Sci. Rep. 2014, 4, 4883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Marchandier, T.; Mariayppan, S.; Kirsanova, M.A.; Abakunov, A.M.; Rousse, G.; Foix, D.; Sougratim, M.-T.; Doublet, M.L.; Tarascon, J.-M. Triggering Anionic Redox Activity in Li3NbS4 Through Cationic Disordering or Substitution. Adv. Energy Mater. 2022, 12, 2201417. [Google Scholar] [CrossRef]
Figure 1. Lattice parameters of rs-(LiScO2)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. The error bars are smaller than the symbols. The dashed line is a guide for the eye. The open circles are our data, and the solid circle is the literature value [7,8,9]. The solid square is our estimate of the lattice parameter of the hypothetical rock-salt LiScO2, derived from the simple relationship a3 = Vtet/2, where Vtet is the unit cell volume of the tetragonal (I41/amd) phase.
Figure 1. Lattice parameters of rs-(LiScO2)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. The error bars are smaller than the symbols. The dashed line is a guide for the eye. The open circles are our data, and the solid circle is the literature value [7,8,9]. The solid square is our estimate of the lattice parameter of the hypothetical rock-salt LiScO2, derived from the simple relationship a3 = Vtet/2, where Vtet is the unit cell volume of the tetragonal (I41/amd) phase.
Materials 16 05336 g001
Figure 2. Sequence of synchrotron powder X-ray diffraction patterns of (LiScO2)0.3(ZnO)0.7 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.65148 Å). The initial rock-salt phase is stable up to 870–900 K; the decomposition products are a mixture of two phases—wurtzite-like (P63mc) and cubic (Ia-3) Sc2O3.
Figure 2. Sequence of synchrotron powder X-ray diffraction patterns of (LiScO2)0.3(ZnO)0.7 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.65148 Å). The initial rock-salt phase is stable up to 870–900 K; the decomposition products are a mixture of two phases—wurtzite-like (P63mc) and cubic (Ia-3) Sc2O3.
Materials 16 05336 g002
Figure 3. (a) Unit cell volumes of rs-(LiScO2)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The solid and dashed lines are the least-squares fits (see Table 2). (b) FWHM of the 200 line of rs-(LiScO2)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The solid and dashed lines are guides for the eye. The solid circles correspond to rs-(LiScO2)0.3(ZnO)0.7, and the open circles correspond to rs-(LiScO2)0.2(ZnO)0.8. In all cases, the error bars are smaller than the symbols.
Figure 3. (a) Unit cell volumes of rs-(LiScO2)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The solid and dashed lines are the least-squares fits (see Table 2). (b) FWHM of the 200 line of rs-(LiScO2)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The solid and dashed lines are guides for the eye. The solid circles correspond to rs-(LiScO2)0.3(ZnO)0.7, and the open circles correspond to rs-(LiScO2)0.2(ZnO)0.8. In all cases, the error bars are smaller than the symbols.
Materials 16 05336 g003
Figure 4. FWHM of the 200 line of the rs-(Li2TiO3)1−x(ZnO)x solid solutions versus the composition under ambient conditions (solid squares). The error bars are smaller than the symbols. The dashed lines are guides for the eye.
Figure 4. FWHM of the 200 line of the rs-(Li2TiO3)1−x(ZnO)x solid solutions versus the composition under ambient conditions (solid squares). The error bars are smaller than the symbols. The dashed lines are guides for the eye.
Materials 16 05336 g004
Figure 5. Lattice parameters of rs-(Li2TiO3)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. The open circles are our data, and the solid circle [7,8,9] and solid square [26,27] are the literature data. The dashed line represents Vegard’s law.
Figure 5. Lattice parameters of rs-(Li2TiO3)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. The open circles are our data, and the solid circle [7,8,9] and solid square [26,27] are the literature data. The dashed line represents Vegard’s law.
Materials 16 05336 g005
Figure 6. Sequence of synchrotron powder X-ray diffraction patterns of (Li2TiO3)0.2(ZnO)0.8 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.6841 Å). The initial rock-salt phase is stable up to 800–1000 K; the decomposition products are mainly a mixture of two phases—wurtzite and rock-salt.
Figure 6. Sequence of synchrotron powder X-ray diffraction patterns of (Li2TiO3)0.2(ZnO)0.8 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.6841 Å). The initial rock-salt phase is stable up to 800–1000 K; the decomposition products are mainly a mixture of two phases—wurtzite and rock-salt.
Materials 16 05336 g006
Figure 7. Unit cell volumes of rs-(Li2TiO3)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The error bars are smaller than the symbols. The dashed lines are the least-squares fits (see Table 3).
Figure 7. Unit cell volumes of rs-(Li2TiO3)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The error bars are smaller than the symbols. The dashed lines are the least-squares fits (see Table 3).
Materials 16 05336 g007
Figure 8. Lattice parameters of rs-(Li3TaO4)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. Open circles are our data, and the solid circle [7,8,9], solid square [31], and diamond [32] are the literature data. The dashed line represents Vegard’s law.
Figure 8. Lattice parameters of rs-(Li3TaO4)1−x(ZnO)x solid solutions as a function of composition under ambient conditions. Open circles are our data, and the solid circle [7,8,9], solid square [31], and diamond [32] are the literature data. The dashed line represents Vegard’s law.
Materials 16 05336 g008
Figure 9. Sequence of synchrotron powder X-ray diffraction patterns of (Li3TaO4)0.5(ZnO)0.5 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.68805 Å). The initial rock-salt phase is stable up to 1000 K; the decomposition products are mainly a mixture of two phases—wurtzite and rock-salt.
Figure 9. Sequence of synchrotron powder X-ray diffraction patterns of (Li3TaO4)0.5(ZnO)0.5 solid solution taken during stepwise heating to 1100 K (wavelength λ = 0.68805 Å). The initial rock-salt phase is stable up to 1000 K; the decomposition products are mainly a mixture of two phases—wurtzite and rock-salt.
Materials 16 05336 g009
Figure 10. Unit cell volumes of rs-(Li3TaO4)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The error bars are smaller than the symbols. The dashed lines are the least-squares fits (see Table 4).
Figure 10. Unit cell volumes of rs-(Li3TaO4)1−x(ZnO)x solid solutions versus temperature at ambient pressure. The error bars are smaller than the symbols. The dashed lines are the least-squares fits (see Table 4).
Materials 16 05336 g010
Figure 11. Comparison of the phase stability of rock-salt ZnO-Lin−1MeOn−2 (Men+ = Sc3+, Ti4+, Ta5+) solid solutions at ambient pressure. Below the dotted lines are the kinetic stability fields of the corresponding cubic phases; above, the decomposition with formation of different phases is observed, n = 3, 4, 5 is an oxidation state of metal cations.
Figure 11. Comparison of the phase stability of rock-salt ZnO-Lin−1MeOn−2 (Men+ = Sc3+, Ti4+, Ta5+) solid solutions at ambient pressure. Below the dotted lines are the kinetic stability fields of the corresponding cubic phases; above, the decomposition with formation of different phases is observed, n = 3, 4, 5 is an oxidation state of metal cations.
Materials 16 05336 g011
Table 1. Crystallographic data for Lin−2Men+On−1 complex oxides used as precursors for the synthesis of ZnO—Lin−2MeOn−1 solid solutions.
Table 1. Crystallographic data for Lin−2Men+On−1 complex oxides used as precursors for the synthesis of ZnO—Lin−2MeOn−1 solid solutions.
PrecursorCation Radius (Å) *Crystal Structure (Space Group)References
LiScO2Sc3+ (0.74)I41/amd (#141)[19,20]
Li2TiO3Ti4+ (0.60)C2/c (#15) Fm-3m (#225) [18,21,22,23,24,25]
Li3TaO4Ta5+ (0.64)C2/c (#15) §P2/n (#13) §[26,27,28,29]
* The ionic radii of Li+ (0.76 Å) and Zn2+ (0.74 Å) are given according to the Shannon scale [30] for coordination number 6. At ambient pressure, monoclinic Li2TiO3 is thermodynamically stable in the 298–1430 K range [20]. At ambient pressure, rock-salt Li2TiO3 is thermodynamically stable between 1430 and 1820 K [21,22,23] and can be quenched to room temperature. The 300 K lattice parameter is 4.14281(5) Å [24] or 4.14276(1) Å [25], which is close to the lattice parameter of rs-ZnO (4.28 Å [5,8,9]). § At ambient pressure, the phase transition between the two monoclinic modifications of Li3TaO4 is observed at 1173 K [28)]. A metastable cubic (Fm-3m) Li3TaO4 has been reported to be kinetically stable below 1093 K [29,31]; its lattice parameter at room temperature is 4.2207(4) Å [31] or 4.214(5) Å [32].
Table 2. Parameters of Equation (2) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)LiScO2 in the 300–900 K range (R2 = 99.9%).
Table 2. Parameters of Equation (2) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)LiScO2 in the 300–900 K range (R2 = 99.9%).
Composition, xV03)α1 × 105 (K−1)α2 × 108 (K−2)
0.878.876(2)3.70(4)2.4(8)
0.779.115(4)3.35(5)3.2(1)
Table 3. Parameters of Equation (4) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)Li2TiO3 in the 300–1000 K range (R2 > 99.8%).
Table 3. Parameters of Equation (4) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)Li2TiO3 in the 300–1000 K range (R2 > 99.8%).
Composition, xV03)α × 105 (K−1)
0.874.78(2)5.42(6)
0.774.52(1)5.75(3)
0.673.592(4)6.16(2)
0.572.797(4)6.39(1)
0.472.296(3)6.20(1)
0.371.563(4)6.39(1)
0.271.221(2)6.47(1)
0.170.820(1)6.42(1)
0.0 [25]71.20(2)6.52(4)
Table 4. Parameters of Equation (4) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)Li3TaO4 in the 300–1000 K range (R2 > 99.8%).
Table 4. Parameters of Equation (4) describing the thermal expansion data of rock-salt solid solutions xZnO—(1 − x)Li3TaO4 in the 300–1000 K range (R2 > 99.8%).
Composition, xV03)α × 105 (K−1)
0.775.81(1)6.79(3)
0.675.62(1)6.81(2)
0.575.23(1)6.67(1)
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

Sokolov, P.S.; Baranov, A.N.; Solozhenko, V.L. Chemical Pressure Effect on the Stabilization of Rock-Salt ZnO—Lin−2MeOn−1 Solid Solutions Synthesized at High Pressure. Materials 2023, 16, 5336. https://doi.org/10.3390/ma16155336

AMA Style

Sokolov PS, Baranov AN, Solozhenko VL. Chemical Pressure Effect on the Stabilization of Rock-Salt ZnO—Lin−2MeOn−1 Solid Solutions Synthesized at High Pressure. Materials. 2023; 16(15):5336. https://doi.org/10.3390/ma16155336

Chicago/Turabian Style

Sokolov, Petr S., Andrey N. Baranov, and Vladimir L. Solozhenko. 2023. "Chemical Pressure Effect on the Stabilization of Rock-Salt ZnO—Lin−2MeOn−1 Solid Solutions Synthesized at High Pressure" Materials 16, no. 15: 5336. https://doi.org/10.3390/ma16155336

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