ACS Publications. Most Trusted. Most Cited. Most Read
Expanding the Ambient-Pressure Phase Space of CaFe2O4-Type Sodium Postspinel Host–Guest Compounds
My Activity
  • Open Access
Article

Expanding the Ambient-Pressure Phase Space of CaFe2O4-Type Sodium Postspinel Host–Guest Compounds
Click to copy article linkArticle link copied!

  • Justin C. Hancock
    Justin C. Hancock
    Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United States
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
  • Kent J. Griffith
    Kent J. Griffith
    Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United States
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
  • Yunyeong Choi
    Yunyeong Choi
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    Department of Materials Science and Engineering, University of California, Berkeley, California 94720, United States
  • Christopher J. Bartel
    Christopher J. Bartel
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    Department of Materials Science and Engineering, University of California, Berkeley, California 94720, United States
  • Saul H. Lapidus
    Saul H. Lapidus
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    X-ray Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States
  • John T. Vaughey
    John T. Vaughey
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    Chemical Sciences and Engineering Division, Argonne National Laboratory, Lemont, Illinois 60439, United States
  • Gerbrand Ceder
    Gerbrand Ceder
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    Department of Materials Science and Engineering, University of California, Berkeley, California 94720, United States
    Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States
  • Kenneth R. Poeppelmeier*
    Kenneth R. Poeppelmeier
    Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United States
    Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United States
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Organic & Inorganic Au

Cite this: ACS Org. Inorg. Au 2022, 2, 1, 8–22
Click to copy citationCitation copied!
https://doi.org/10.1021/acsorginorgau.1c00019
Published September 1, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

CaFe2O4-type sodium postspinels (Na-CFs), with Na+ occupying tunnel sites, are of interest as prospective battery electrodes. While many compounds of this structure type require high-pressure synthesis, several compounds are known to form at ambient pressure. Here we report a large expansion of the known Na-CF phase space at ambient pressure, having successfully synthesized NaCrTiO4, NaRhTiO4, NaCrSnO4, NaInSnO4, NaMg0.5Ti1.5O4, NaFe0.5Ti1.5O4, NaMg0.5Sn1.5O4, NaMn0.5Sn1.5O4, NaFe0.5Sn1.5O4, NaCo0.5Sn1.5O4, NaNi0.5Sn1.5O4, NaCu0.5Sn1.5O4, NaZn0.5Sn1.5O4, NaCd0.5Sn1.5O4, NaSc1.5Sb0.5O4, Na1.16In1.18Sb0.66O4, and several solid solutions. In contrast to earlier reports, even cations that are strongly Jahn–Teller active (e.g., Mn3+ and Cu2+) can form Na-CFs at ambient pressure when combined with Sn4+ rather than with the smaller Ti4+. Order and disorder are probed at the average and local length-scales with synchrotron powder X-ray diffraction and solid-state NMR spectroscopy. Strong ordering of framework cations between the two framework sites is not observed, except in the case of Na1.16In1.18Sb0.66O4. This compound is the first example of an Na-CF that contains Na+ in both the tunnel and framework sites, reminiscent of Li-rich spinels. Trends in the thermodynamic stability of the new compounds are explained on the basis of crystal-chemistry and density functional theory (DFT). Further DFT calculations examine the relative stability of the CF versus spinel structures at various degrees of sodium extraction in the context of electrochemical battery reactions.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2021 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

The CaFe2O4 (calcium ferrite or CF) structure, also occasionally referred to as the CaV2O4 structure because it was identified first for this compound, (1,2) has been of increasing interest in recent years. Previously, the structure type was primarily of geological and crystal-chemical relevance. (3−8) Many spinel compounds transform to this structure type under high pressure (hence the CF structure is termed “post-spinel”), and the structure is thought be a host for various cations in the Earth’s mantle. (3−10) However, many property- and application-oriented studies have been recently published, indicating renewed interest in this class of materials. (11−28) Materials that crystallize in the CF structure (see Figure 1) have been reported with interesting magnetic and electronic properties resulting from its pseudo-1D chain structure and geometric frustration. (11−15) The large tunnels in CF compounds have led to their investigation as host structures for phosphors (16−18) and battery cathode materials. (19−28) CF host structures have been shown to function well as both Na and Li cathode materials, (19,20,23,24) and Mg ions are also predicted to be mobile in the tunnels. (26,27)

Figure 1

Figure 1. CaFe2O4 structure. Gray spheres represent Na+. Green and blue octahedra represent the two symmetrically independent framework sites comprising the two “double rutile” chains.

The calcium ferrite (CF) structure is composed of a framework of octahedrally coordinated cations, arranged into “double-rutile” chains that run parallel to the b axis. Within the chains, the octahedra are edge-sharing, and the chains connect to adjacent chains via corner-sharing. There are two crystallographically distinct metal sites in the framework, with each double chain containing only one unique site, paired by symmetry, as illustrated in Figure 1. The chains surround one-dimensional channels of 8-coordinate cations. For the CF materials synthesizable at ambient pressure, the tunnel sites have been reported to be occupied by Na+, Ca2+, Sr2+, and Ba2+. (29) CF compounds with Na+ occupying the tunnel sites (compounds hereafter referred to as Na-CFs) are the most likely to be relevant to the energy storage materials community because of the relatively high mobility of Na+. Divalent Ca2+ is expected to show low mobility, (28) and Sr2+ and Ba2+ primarily form CF structures when the framework cations are redox inactive trivalent rare-earth metals. (29) However, Na-CF materials with redox-active transition metal cations have been synthesized at ambient pressure, although the library of such compounds, detailed in the following paragraph, is limited.
The only comprehensive study of the crystal chemistry of Na-CF phases was performed by Reid, Wadsley, and Sienko, who studied compounds of the formula NaA3+B4+O4 using high-temperature solid-state synthesis. (8) They synthesized and evaluated NaScTiO4, NaFeTiO4, NaFeSnO4, NaScZrO4, NaScHfO4, and NaAlGeO4, the last of which required high pressure. Other compositions, such as NaMnTiO4, were attempted but resulted in no CF phases. Based on these results, Reid et al. concluded that “spherical” (Jahn–Teller inactive) ions favor formation of a CF phase. Several new compounds of this type have since been reported, (25,30−32) although many require high pressure for synthesis, such as NaAlSiO4, NaV2O4, NaCr2O4, NaMn2O4, and NaRh2O4. (5,6,11,12,33,34) Na-CF compounds with other stoichiometries have also been shown to exist, including NaA2+0.5B4+1.5O4 (A2+ = Co2+, Ni2+ and B4+ = Ti4+) and NaA3+1.5B5+0.5O4 (A3+ = Fe3+ and B5+ = Sb5+), all of which have been synthesized at ambient pressure. (35−37) The few Na-CFs deviating significantly from these stoichiometries were synthesized via hydrothermal synthesis, including Na0.55Fe0.28Ti1.72O4 and Na3Mn4Te2O12 (Na[Mn1.33Te0.67]O4), the latter having a superstructure owing to crystallographic ordering of Te6+ and Mn2+/3+. (38,39) Notably, some of the CF phases synthesized at ambient pressure contain ions that are weakly Jahn–Teller active, thus the spherical ion preference appears not to be a strict criterion. This would suggest many more as yet undiscovered Na-CF compounds may be stable even at ambient pressure.
In this paper, we report a large expansion of the ambient-pressure phase space of Na-CF materials and critically examine the crystal chemical relationships. Notably, we successfully synthesized several new Na-CFs with redox-active transitions metals, which are prospective Na/Li/Mg battery electrode materials, and Na1.16In1.18Sb0.66O4, the first Na-CF that contains sodium on both the tunnel and framework sites.

Experimental Section

Click to copy section linkSection link copied!

Synthesis

All compounds were synthesized via high-temperature solid-state reactions. The synthesis temperature depended on the composition of the CF phase. Unless stated otherwise, starting materials were NaHCO3 and binary metal oxides. In a typical synthesis, appropriate amounts of these reactants were mixed with a mortar and pestle then pressed into a pellet at 400 MPa. The pellet was placed in either a platinum crucible (air syntheses) or platinum boat (inert atmosphere syntheses), heated at the specified temperature for 2 days, ground into a powder, and repeated as necessary. Some compositions (those containing Cr3+, Mn2+, and Fe2+) required an inert atmosphere for synthesis. These samples were heated in a tube furnace under flowing argon, with a titanium rod placed upstream to remove any residual O2. In some cases, excess NaHCO3 was added in the subsequent annealing steps to account for sodium loss from volatilization. All reactions carried out in air (except NaMnSnO4, for which we were reproducing a reported synthesis) were quenched on the benchtop to reflect the thermodynamics of the synthesis temperature and minimize cooling rate effects. All reactions carried out in a tube furnace were cooled by shutting off the power once the dwell step completed. Detailed synthetic information for each new compound is discussed in more detail in the Results and Discussion section, and reaction conditions for combinations of cations that did not produce a CF phase are included in Table S1.

X-ray Diffraction

Phase purity was assessed by laboratory powder X-ray diffraction (PXRD) using both Rigaku Ultima IV and Rigaku SmartLab X-ray diffractometers. These data were collected over a 2θ range of 10–60° under ambient conditions. Synchrotron X-ray radiation was used for Rietveld refinements, except in the case of NaFe0.5Ti1.5O4, NaMn0.5Sn1.5O4, NaFe0.5Sn1.5O4, and NaCd0.5Sn1.5O4. These data were collected at 11-BM at the Advanced Photon Source at Argonne National Laboratory using a wavelength of 0.45788 Å (∼27 keV). This wavelength was chosen to minimize absorption of X-rays by In and Sn. Samples were packed into Kapton capillaries.
Rietveld refinement was performed using the General Structure Analysis System II (GSAS II) package. (40) Ionic scattering factors were used in all cases. All data were refined against an orthorhombic unit cell (space group Pnma). In all cases, the unit cell parameters, atomic coordinates, and Uiso values were refined. In the case of NaCd0.5Sn1.5O4 only, Uiso for the oxygen atoms was not refined and set to a reasonable value of 0.01. An 8- or 10-term Chebyshev polynomial was used to fit the backgrounds, with an added background peak to account for scattering from the Kapton capillary. To decrease the degrees of freedom, global compositions were fixed, and atomic positions and isotropic thermal parameters were constrained to be equal for all atoms sharing the same crystallographic sites. In cases where cation ordering of the framework sites seemed probable (see the Results and Discussion section), occupancies were refined, but in most cases occupation of the framework sites was assumed to be statistically distributed to simplify the model and avoid overfitting.

Solid-State NMR Spectroscopy

23Na solid-state NMR spectra were recorded at 9.4 T (νL(23Na) = 105.7 MHz) with a Bruker Avance III spectrometer and a Bruker HX probe. Samples were packed into a 4.0 mm diameter (80 μL volume) zirconia rotor with a Kel-F cap and measured at ambient temperature under 12.5 kHz magic-angle spinning (MAS), corresponding to approximately 30 °C, unless otherwise noted. One-dimensional spectra were collected with a one-pulse (Bloch decay) sequence and a 2.0 μs (π/4)liquid pulse. NaCl (aqueous, 1.0 M) was used to optimize the solution π/2 pulse and as the 23Na reference at 0 ppm. In all cases, the recycle delay was set to at least 5 × T1, where T1 is measured with a saturation-recovery pulse sequence. Typical T1 relaxation times were 0.5–8 s. The multiple-quantum magic-angle spinning (MQMAS) spectrum of Na1.16In1.18Sb0.66O4 was recorded with a z-filtered pulse sequence with excitation and conversion pulses of 9.0 and 3.0 μs followed by a 28 μs selective pulse. (41) Acquisition in the indirect dimension comprised 192 t1 increments of 25 μs. For each t1-slice, 396 scans were averaged and the recycle delay was 0.6 s, resulting in an experimental time of 12.7 h.

DFT Calculations

Calculations is this work are based on density functional theory (DFT) as implemented in the Vienna Ab Initio Simulation Package (VASP) (42) using the projector augmented-wave method (43,44) and the generalized gradient approximation as formulated by Perdew, Burke, and Ernzerhof. (45) For all calculations, the energy cutoff was set to 520 eV, and at least 1000 k-points were used per reciprocal atom. For geometry optimizations, energies were converged to 10–5 eV for electronic steps and 10–4 eV for ionic steps. The Hubbard U correction was used for the transition metal atoms to account for the self-interaction error of semilocal density functionals. (46) U parameters were chosen to be consistent with the Materials Project database, (47) as reported by Jain et al. (Co, 3.32 eV; Cr, 3.7 eV; Fe, 5.3 eV; Mn, 3.9 eV; Ni, 6.2 eV). (48)
Supercells of the spinel (Fdm) and CF postspinels (Pnma) with 32 oxygen ions (Nax(A,B)16O32) were used for all calculations, and the ionic positions, cell shape, and volume were allowed to optimize during relaxations. The cell size was chosen to provide enough ions for sampling various occupations for Na and Na vacancies on the alkali site as well as A and B on the octahedral sies. To generate ordered structures, the lowest and second lowest electrostatic energy configurations (as calculated with the Ewald method) were sampled to order the A and B cations on the octahedral sites and determine which Na sites to remove. In most compositions, two different configurations were used but in slow converging composition only the lowest configuration was used for DFT calculation. In total, 264 different structures were calculated at varying levels of sodiation (x) to determine the thermodynamic stability of each phase in the spinel and postspinel structures. Thermodynamic stability was determined using the convex hull method, where the formation energies of all competing phases in each Na-A-B-O chemical space were taken from the Materials Project database. The pymatgen library was used to set up and analyze the calculations in this work. (49)

Results and Discussion

Click to copy section linkSection link copied!

The Na-CF chemical space was systematically explored using numerous cation combinations and stoichiometries, and this approach and the results are summarized schematically in Figure 2. Sixteen new Na-CF end-member compounds were successfully synthesized, and NaMnSnO4, reported recently by Chiring et al., (25) was reexamined. Rietveld refinement data for these compounds are shown in Table 1, and an example Rietveld refinement is shown in Figure 3. Additional Rietveld refinements are shown in Figures S1–S16.

Figure 2

Figure 2. Schematic summarizing the synthetic approach and results presented in this paper. The ion combinations in green were successfully synthesized in the postspinel structure, whereas the ion combinations in red formed alternative phases.

Table 1. Rietveld Refinement Data for New Na-CF Compounds
sourcesynchrotron
chemical formulaNa0.99Cr0.99Ti1.01O4Na0.96Rh0.96Ti1.04O4NaCrSnO4NaMnSnO4Na0.96In0.96Sn1.04O4NaMg0.5Ti1.5O4
formula weight186.61234.67257.67260.61319.73170.99
temperature (K)298
wavelength (Å)0.457880
crystal systemorthorhombic
space group (no.)Pnma (62)
a (Å)9.101854(15)9.16953(5)9.26744(3)9.42779(10)9.53203(3)9.17179(2)
b (Å)2.933813(4)2.947138(15)3.048247(9)3.02517(3)3.172342(8)2.968472(6)
c (Å)10.668108(17)10.79754(5)10.93396(4)11.11389(11)11.29355(3)10.76171(2)
α = β = γ (deg)90
V3)284.872(1)291.791(3)308.878(2)316.976(7)341.504(2)290.001(1)
Z4
profile range3 ≤ 2θ ≤ 37.9963
GOF1.931.042.032.382.461.40
Rp (%)6.8210.607.719.926.556.27
Rwp (%)9.7012.889.7212.449.107.37
sourcesynchrotron
chemical formulaNaMg0.5Sn1.5O4NaCo0.5Sn1.5O4NaNi0.5Sn1.5O4NaCu0.5Sn1.5O4NaZn0.5Sn1.5O4NaSc1.5Sb0.5O4
formula weight277.17294.49294.37296.79297.71215.3
temperature (K)298
wavelength (Å)0.457880
crystal systemorthorhombic
space group (no.)Pnma (62)
a (Å)9.41987(3)9.41576(3)9.39739(3)9.47695(3)9.43720(5)9.44848(6)
b (Å)3.106399(8)3.115976(8)3.099946(7)3.101415(7)3.113964(15)3.133693(18)
c (Å)11.11941(3)11.10660(3)11.10428(3)11.08428(3)11.13601(6)11.13570(7)
α = β = γ (deg)90
V3)325.375(2)325.859(2)323.483(2)325.789(2)327.255(4)329.713(4)
Z4
profile range3 ≤ 2θ ≤ 37.9963
GOF1.341.851.542.011.662.08
Rp (%)7.486.316.966.488.799.83
Rwp (%)8.968.339.108.2612.4512.40
sourcesynchrotronCu Kα
chemical formulaNa1.16In1.18Sb0.66O4NaFe0.5Ti1.5O4NaMn0.5Sn1.5O4NaFe0.5Sn1.5O4NaCd0.5Sn1.5O4
formula weight306.51186.76292.49292.94321.23
temperature (K)298298
wavelength (Å)0.4578801.5406
crystal systemorthorhombicorthorhombic
space group (no.)Pnma (62)Pnma (62)
a (Å)9.53383(2)9.2175(3)9.48122(13)9.40435(15)9.5604(2)
b (Å)3.168104(6)2.96553(9)3.13771(4)3.11440(5)3.17096(7)
c (Å)11.28573(2)10.7674(4)11.19800(15)11.10497(18)11.2812(3)
α = β = γ (deg)9090
V3)340.876(1)294.33(2)333.133(9)325.252(11)341.995(17)
Z44
profile range3 ≤ 2θ ≤ 37.996310 ≤ 2θ ≤ 13010 ≤ 2θ ≤ 120
GOF2.161.812.262.422.55
Rp (%)8.651.843.342.833.99
Rwp (%)11.612.554.453.965.36

Figure 3

Figure 3. Rietveld refinement using synchrotron data for NaMg0.5Ti1.5O4. Blue crosses are the observed intensities, the green curve is the fitted pattern, the black curve is the difference pattern, and the red tick marks indicate the location of the CF-NaMg0.5Ti1.5O4 peaks.

Synthetic Studies

New Postspinels in the Na+-A3+-B4+-O2– System (Idealized Formula: NaA3+B4+O4)

The new CF phase NaCrTiO4 was successfully synthesized at ambient pressure. NaCrTiO4 must be synthesized in an inert atmosphere (Ar in this case) to avoid the oxidation of Cr3+ to Cr6+ and formation of Na2CrO4. Initially, reactions performed in the temperature range of 875–900 °C were used to minimize sodium volatility. For a 1:1:1 Na/Cr/Ti ratio, after 48 h a mixture of three quaternary phases formed, which included a CF phase, layered Na1–xCr1–xTixO2, probably close to the Na0.6Cr0.6Ti0.4O2 composition reported by Avdeev et al., (50) and NaxCrxTi2–xO4, with a nonstoichiometric sodium iron titanate (NSIT) structure. The NSIT phase probably has a composition close to Na0.9Cr0.9Ti1.1O4, which is the reported upper limit of x found in the NaxFexTi2–xO4 system. (51) (Note that, confusingly, NSIT phases are sometimes referred to as CaV2O4-type structures in databases, but we emphasize that the NSIT structure is distinct from the CF structure.) While the quaternary phases formed relatively quickly, they reacted very slowly with each other, and the system was slow to reach thermal equilibrium once these phases formed. Heating the mixture for an additional 48 h at 900 °C increased the percentage of CF-NaCrTiO4, but the other two phases were still present. Higher temperatures sped up the reaction, and the loss of Na from volatility appears to be minimal at 950 °C. At this temperature, the NSIT phase was no longer observed. Instead, the stoichiometric mixture (1:1:1 Na/Cr/Ti) resulted in Na0.6Cr0.6Ti0.4O2 and the CF compound (see Figure 4), the relative ratios of which changed little upon further heating. This suggests that the CF compound is nonstoichiometric and is deficient in both sodium and chromium relative to the ideal composition NaCrTiO4. Addition of ∼2% excess TiO2 to this mixture and further annealing at 900 °C led to a nearly single-phase product, suggesting an actual composition of Na0.99Cr0.99Ti1.01O4 for the CF phase. A separate mixture synthesized at 950 °C with 5% TiO2 excess still contained some Na0.6Cr0.6Ti0.4O2, suggesting the composition of the CF phase may have some degree of temperature dependency. Because the competing phases have such similar stoichiometries, the ratio of cations must be carefully controlled to avoid significant fractions of the non-CF phases.

Figure 4

Figure 4. Powder XRD patterns for the 1:1:1 and 0.99:0.99:1.01 Na/Cr/Ti samples compared to a calculated pattern for the CF phase. Asterisks highlight peaks from the secondary phase Na0.6Cr0.6Ti0.4O2 in the 1:1:1 Na/Cr/Ti sample.

The synthesis of NaCrSnO4 requires higher temperatures than that of NaCrTiO4. The formation of the CF phase was slow even at 950 °C, and NaCrO2 and SnO2 were the main phases after 48 h. Heating at 1000 °C produced CF-NaCrSnO4 as the main phase, but NaCrO2, potentially Sn-substituted, also formed, and SnO2 was still present at shorter reaction times. Reheating these mixtures again at 1000 °C resulted in elimination of the SnO2 phase, but with similar proportions of NaCrO2 and CF-NaCrSnO4. This suggests a loss of Sn. In some runs, metallic Sn was observed. Cleaning the alumina tube alleviated this problem somewhat, suggesting reductive species build up over time in the tube, potentially from the Ti rod, used as a getter, and its interaction with volatile phases. Nevertheless, phase-pure NaCrSnO4 was never obtained by this synthetic method alone, even when using excess SnO2, but the pure CF phase could likely be formed in a closed system such as a sealed metal tube. However, purification was possible. NaCrO2 was removed by treating the mixture with molten KNO3, which selectively oxidized the NaCrO2. The soluble Cr(VI) products were then dissolved in water, and the remaining solid was filtered, leaving behind only CF-NaCrSnO4.
Recently, CF-NaMnSnO4 was reported to be synthesizable under ambient pressure in air. (25) The authors reported that, when using a stoichiometric starting composition, phase purity was achieved only by slowly cooling at 0.5 °C/min after heating at 1200 °C for a day. Similarly, we observed secondary phases upon quenching stoichiometric mixtures but obtained a nearly phase-pure CF compound upon quenching a sample with a Na/Mn/Sn ratio of 0.96:0.96:1.04. The lattice parameters for the quenched and slowly cooled samples are significantly different, suggesting the difference in composition is real. These results, like those for the Na–Cr–Ti–O system, suggest a temperature dependence of the composition of the CF phase. Interestingly, both the quenched and slowly cooled samples had broad peaks with poorer Rietveld fits compared to the other compositions, probably caused by a high degree of strain. Presumably, this strain results from substituting a strongly JT-active cation (Mn3+) that prefers highly distorted octahedral environments into sites that usually contain more spherically symmetric cations.
CF-NaRhTiO4 was synthesized in air from NaRhO2 (prepared by heating NaHCO3 and metallic Rh powder at 900 °C for about 1 day) and TiO2. A layered phase similar to the one formed in the Na–Cr–Ti-O system, NaxRh1–xTixO2 forms quickly, with subsequent slow formation of the CF phase at 900 °C. Increasing the temperature to 950 °C increases the rate of formation, and the CF phase is the major phase after 48 h. However, heating at 1050 °C destabilizes the CF phase. At this temperature, no CF phase is observed, and the layered phase is the primary phase along with another unknown phase present in small amounts. As in the case of Na1–xCr1–xTi1+xO4, phase purity was not achieved with the ideal stoichiometry. In fact, the product of the mixture with a 1:1:1 ratio of cations had a nearly identical PXRD pattern to that of the 1:1:1 ratio in the Na–Cr–Ti–O system. Phase purity was achieved by using an excess of titanium relative to the ideal composition, and the phase-pure sample had a nominal composition of Na0.96Rh0.96Ti1.04O4. CF-NaRhSnO4 did not form under similar conditions up to a temperature of 1100 °C.
CF-NaInSnO4 appears to be more refractory than the CF phases containing Ti4+ and could be synthesized in the 950–1200 °C temperature range. The sample with the highest phase purity was synthesized by using a Na/In/Sn ratio of 0.98:0.96:1.04 and air-quenching from 1200 °C, the highest temperature studied, after ∼20 h. Thus, the actual composition of the CF phase is likely close to Na0.96In0.96Sn1.04O4, and presumably, the excess Na was lost through volatilization. Thus, all the new CF compounds of the type NaA3+B4+O4 deviate slightly from the ideal stoichiometry, with the possible exception of NaCrSnO4, whose precise composition was not determined owing to the difficulty in obtaining a phase-pure sample. CF-NaInZrO4 did not form under similar conditions.

New Postspinels in the Na+-A2+-B4+-O2– System (Idealized Formula: NaA2+0.5B4+1.5O4)

The compounds NaMg0.5Ti1.5O4 and NaFe0.5Ti1.5O4 can be synthesized with high phase purity. A stoichiometric mixture of NaHCO3, MgO, and TiO2 heated at 925 °C for 48 h resulted in primarily Na0.9Mg0.45Ti1.55O4 (NSIT-type structure). Phase-pure NaMg0.5Ti1.5O4 was not obtained until the mixture was further heated at 950 °C for ∼96 h with an intermediate grinding step. The CF phase decomposes after 3 h at 1050 °C into NSIT-Na0.9Mg0.45Ti1.55O4, layered Na0.68Mg0.34Ti0.66O2, and MgO. The presence of MgO suggests Na volatilization at this temperature. Interestingly, NaMg0.5Ti1.5O4 was not reported in a previous study of the Na–Mg–Ti–O system that found six phases, possibly because of the fast formation of the NSIT and layered phase and limited temperature range in which the CF phase can be formed. (52) NaFe0.5Ti1.5O4 was readily synthesized in one step at 925 °C with a 5% excess of sodium from Na8Ti5O14, FeTiO3, and TiO2 under flowing argon. The relatively fast kinetics of this synthesis could be explained by the use of different reactants. This phase seems somewhat air-sensitive even at room temperature. The sample used for synchrotron diffraction had highly asymmetric peaks skewed toward higher angles after being stored for ∼3 months, suggesting topotactic oxidation (Fe2+ to Fe3+). With this observation, a new sample was synthesized and laboratory X-ray data were collected for Rietveld refinement. It should be pointed out that a CF phase containing Fe2+ has been reported to form via hydrothermal synthesis, but the phase has a different composition (Na0.55Fe0.28Ti1.72O4) and is unusually Na-deficient. (38) In addition, a “CF-like” secondary phase was mentioned in a study of the NSIT phases NaxFe2+x/2Ti4+2–x/2O4 for higher values of x but was not discussed further. (53) No other divalent cations (Mn2+, Cu2+, or Zn2+) could be fully substituted into NaA2+0.5Ti4+1.5O4 under similar synthetic conditions.
No CF phases with the composition NaA2+0.5Sn1.5O4 have been previously reported. However, more compositions of this type could be synthesized than in the NaA2+0.5Ti1.5O4 system; Mg2+, Mn2+, Fe2+, Co2+, Ni2+, Cu2+, Zn2+, and Cd2+ can all form CF compounds when combined with Sn4+. These compounds are more refractory than the corresponding compounds with Ti4+, so a wider range of temperatures (950–1200 °C) was used in the synthesis of the NaA2+0.5Sn1.5O4 compounds. The compounds NaA2+0.5Sn1.5O4 (A2+ = Mg2+, Co2+, Ni2+) were first heated at 1200 °C for 1 day with a 5% excess of sodium, then reground with an additional 5% excess of sodium (as NaHCO3) and reheated at 1000 °C for 2 days. The excess Na decreased the amount of SnO2 observed as a minor secondary phase. The compounds NaA2+0.5Sn1.5O4 (A2+ = Mn2+, Fe2+, Zn2+, Cd2+) were synthesized at 1000 °C for 48 h, and NaCu0.5Sn1.5O4 was synthesized at 950 °C for 96 h. NaMn0.5Sn1.5O4 and NaFe0.5Sn1.5O4 were synthesized under flowing argon, and Fe2+ was introduced as Fe(C2O4)·2H2O. In most cases, near phase purity was achieved, although it was difficult to eliminate SnO2 as a secondary phase (e.g., ∼1% by weight in the case of NaCo0.5Sn1.5O4). NaZn0.5Sn1.5O4 always formed along with secondary phases, though in the best sample, the ratio of the most intense CF peak of the PXRD pattern to the most intense secondary phase peak was slightly greater than 13:1.
While a very small degree of deviation from ideal stoichiometry is possible, no systematic trends were observed during synthesis of the NaA2+0.5B4+1.5O4 compounds that would suggest sodium vacancies, in contrast to the NaA3+B4+O4 compounds.

New Postspinels in the Na+-A3+-B5+-O2– System (Idealized Formula: NaA3+1.5B5+0.5O4)

CF phases were also found in the Na–In–Sb–O and Na–Sc–Sb–O systems. Attempts to synthesize materials of the ideal compositions NaIn1.5Sb0.5O4 and NaSc1.5Sb0.5O4 always resulted in a CF phase and either In2O3 or Sc2O3. Decreasing the ratio of In2O3 to the other reactants (NaHCO3 and Sb2O3) improved the phase purity, suggesting the composition of the CF phase is better represented by the formula Na1+xIn1.5–2xSb0.5+xO4. When x = 0.158, a single-phase CF material was obtained after heating at 1200 °C for 1 day, followed by quenching. Thus, In3+ appears to be cosubstituted by Sb5+ and Na+. We were not able to make the pure CF phase in the Na–Sc–Sb–O system using the same strategy. The best sample was synthesized at 1100 °C for 96 h with an intermediate grinding and with the ideal ratio of cations (i.e., 1:1.5:0.5 ratio of Na to Sc to Sb), but this sample contained Sc2O3 and another unknown phase as secondary phases. NaCr1.5Sb0.5O4 did not form at 900 or 950 °C under flowing argon and from a mixture of NaHCO3, NaSbO3, and Cr2O3.

Solid Solutions

Owing to similarities in synthesis conditions, ability to move from ideal cation locations, and nonstoichiometry, solid solutions between the systems NaA3+B4+O4 and NaA2+0.5B4+1.5O4 should exist. Several compositions were tried, including NaCo1/3Fe1/3Ti4/3O4, NaNi1/3Fe1/3Ti4/3O4, and NaNi1/3Sc1/3Ti4/3O4, which produced CF phases with no apparent secondary phases. NaCo1/3Fe1/3Ti4/3O4 and NaNi1/3Fe1/3Ti4/3O4 were synthesized at 900 °C for 96 h with intermediate grindings. NaNi1/3Sc1/3Ti4/3O4 also formed at these temperatures and was stable at 1050 °C, in contrast to the other Na-CFs containing Ti. The flexibility of these phases with respect to compositional tuning further emphasizes their potential usefulness for applications such as intercalation cathodes for energy storage. The results summarizing the now-expanded Na-CF phase space (excluding solid solutions) are shown in Figure 5.

Figure 5

Figure 5. A combinatorial representation of the Na-CF phase space. The phase-space shown covers all Na-CF compounds and combinations known to have been attempted in previous studies or this work. Conceivable but unexplored combinations (e.g., Sc3+ and Mo4+) are not depicted.

Crystal Chemistry

Cation Distribution

The Na-CF material in the Na–In–Sb–O system was obtained phase pure with the nominal stoichiometry Na1.16In1.18Sb0.66O4 (Na1+xIn1.5–2xSb0.5+xO4, x = 0.16). This would imply a cosubstitution of Na+ and Sb5+ for In3+, which would be the first reported instance of Na+ occupying the framework sites of a CF compound. Rietveld refinements for this material provide further evidence of this new cation distribution. The CF structure contains two independent octahedral cation sites, which are referred to here to as the M1 and M2 sites. Assuming an actual composition of NaIn1.5Sb0.5O4 resulted in unreasonable (negative) thermal parameters (Uiso) for the M2 site, even when the occupancy of Na+ in the tunnel sites was refined. Refining the composition Na[Na0.16In1.18Sb0.66]O4 with complete disorder of the framework sites also resulted in a negative Uiso for the M2 site. Refining the fractional occupancies of the framework sites and constraining the composition alleviated this problem. Because three atoms distributed between two sites is an underdetermined system using only X-ray data and because In3+ and Sb5+ have nearly equal scattering factors, Sb5+ was treated as In3+ for this refinement. (Note that for the final refinement and in the CIF, Sb5+ was reintroduced such that the In/Sb ratio was the same for both sites.) By this refinement, nearly all (∼90%) of the octahedral Na+ occupies the M1 site rather than the M2 site, with a final distribution of Na[(In/Sb)0.856(2)Na0.144(2)]M1[(In/Sb)0.984(2)Na0.016(2)]M2O4. A model placing Na+ on the M2 site only and not allowing for occupancy refinement results in a significantly worse fit. The results of these refinements are summarized in Table 2. This Na-CF with Na+ occupying the framework sites is analogous to lithium-rich spinels such as Li4Mn5O12 (Li[Li1/3Mn5/3]O4) and the commercial lithium-ion battery material Li4Ti5O12 (Li[Li1/3Ti5/3]O4). (54,55)
Table 2. Refinement Statistics for Different Structural Models of Na1.16In1.18Sb0.66O4
starting cation distribution modelM1 and M2 site occupancies refined?Rwp (%)comment
Na[In0.75Sb0.25]M1[In0.75Sb0.25]M2O4no11.90negative Uiso for M2
Na[In0.59Sb0.33Na0.08]M1[In0.59Sb0.33Na0.08]M2O4no11.97negative Uiso for M2
Na[In0.92Na0.08]M1[In0.92Na0.08]M2O4yes11.61reasonable Uisos, 90% Na on M1 site
Na[In]M1[In0.84Na0.16]M2O4no13.30negative Uiso for M2
The sodium environments in Na1.16In1.18Sb0.66O4 were further investigated with 23Na solid-state magic-angle spinning (MAS) NMR spectroscopy. The 23Na NMR spectrum of Na1.16In1.18Sb0.66O4, shown in Figure 6a, contains a series of overlapping resonances at lower frequencies (−10 to +20 ppm), with an additional resonance at +34 ppm. The overlapping low-frequency resonances, which resolve into at least four signals with multiple-quantum magic-angle spinning (MQMAS) (Figure S17), likely correspond to Na in the tunnel sites. Although there is crystallographically only one tunnel site, the local environment is more complex given the multiple next-nearest neighbor possibilities. Multiple peaks for the tunnel sites are also resolved in some other CF phases (see Figure S18). The higher shift of the signal at 34 ppm is consistent with octahedral Na+, and the integration of this peak (11%) is reasonably consistent with the fraction of Na expected to occupy the octahedral sites. Another possibility that could explain the 23Na NMR spectrum is that the stoichiometry is ideal, NaIn1.5Sb0.5O4, but with Na+-In3+ antisite defects analogous to inversion in spinels. This is sensible crystal-chemically, as In3+ does occupy 8-coordinate sites in some phases. (56) If this did occur in NaIn1.5Sb0.5O4, it would also be expected to occur in NaInSnO4. However, the 23Na NMR spectrum for Na0.96In0.96Sn1.04O4, shown in Figure 6b, contains only the low-frequency signal distribution. Thus, we concluded antisite defects were unlikely to be the cause of the signal at 34 ppm for Na1.16In1.18Sb0.66O4. A third possibility is that the high frequency resonance is from an unidentified, and possibly amorphous, secondary phase. In the nominally stoichiometric NaIn1.5Sb0.5O4 sample, PXRD showed only CF and In2O3 as crystalline phases, and the 23Na NMR spectrum matched that of Na1.16In1.18Sb0.66O4. Since this sample is Na-poor, it is unlikely that the resonance at 34 ppm could come from a Na-containing secondary phase. The 23Na NMR spectra for “NaSc1.5Sb0.5O4” and “NaCd0.5Sn1.5O4” also show weak resonances at 35 ppm (Figure 6 and Figure S18), and these phases also probably contain framework sodium. While the CF phase in the Na–Sc–Sb–O system is likely also not quite stoichiometric NaSc1.5Sb0.5O4, a Rietveld refinement assuming this stoichiometry gave reasonable Uiso values when framework occupancies were refined. However, the 23Na NMR spectrum for the “NaSc1.5Sb0.5O4” sample (Figure 6c) contained a small peak at 35 ppm that integrates to ∼4.5% of the total Na. There was a secondary phase that we were unable to identify so we cannot conclusively say that this peak corresponds to framework-site Na in the CF phase, but it is likely given the parallels with the Na–In–Sb–O system. If “NaSc1.5Sb0.5O4” does deviate from the ideal stoichiometry, the difference is not as large as in Na1.16In1.18Sb0.66O4, and corefinements with neutron data may be necessary to determine the cation distribution accurately.

Figure 6

Figure 6. 23Na solid-state NMR spectra at 12.5 kHz MAS and 9.4 T: (a) Na1.16In1.18Sb0.66O4, (b) Na0.96In0.96Sn1.04O4, and (c) “NaSc1.5Sb0.5O4” central transition resonances.

While the CF structure contains two crystallographically independent octahedral cation sites, each double-chain contains only one of these sites (see Figure 1). Reid et al. found no evidence of cation ordering in the Na-CFs they synthesized. (8) A single-crystal study of NaFeRuO4 also found no site preference for the framework cations Fe3+ and Ru4+. (31) Only two studies of hydrothermally synthesized Na-CFs have found any statistically significant site preference. (38,57) In these cases, the synthesis temperature is considerably lower than typically used for solid-state synthesis. As noted by Reid et al., the intensities of most of the X-ray reflections are insensitive to ordering. However, the (101) reflection is strongly affected by site preference and has negligible intensity in the absence of ordering and/or when the framework cations have similar scattering factors. Thus, in this study, occupancies of the framework cations were only refined when the (101) peak was apparent in the PXRD pattern to avoid overfitting.
None of the compounds reported here have strong (101) reflections present, but a weak reflection (note the small peak at ∼3.5° in Figure 3) indicating some degree of site preference is apparent for NaMg0.5Ti1.5O4, NaCu0.5Sn1.5O4, Na1.16In1.18Sb0.66O4, and NaSc1.5Sb0.5O4. That the CF phase obtained in the Na–In–Sb–O system has a (101) reflection is further evidence that it is not the initially expected NaIn1.5Sb0.5O4 (In3+ and Sb5+ are essentially indistinguishable by X-rays). One might expect that cation site preference is more likely when cation radii and charge density differences are larger or when JT-active cations like Mn3+ and Cu2+ are present, since these cations are expected to prefer different coordination environments than Ti4+ or Sn4+. Of the Na-CF compounds reported here, only NaMnSnO4 and NaCu0.5Sn1.5O4 contain cations with strong Jahn–Teller distortions. The (101) reflection is essentially nonexistent in the case of NaMnSnO4, even for the slowly cooled sample, thus there is no evidence of site preference. For NaCu0.5Sn1.5O4, refinement of the occupancies of the framework sites indicated that ∼63% of the Cu2+ occupies the M1 site, with Cu occupancies of 0.317(2) and 0.183(2) for the M1 and M2 sites, respectively. Likewise, ∼63% of the Mg2+ cations occupy the M1 site in NaMg0.5Ti1.5O4, with Mg occupancies of 0.316(1) and 0.184(1) on the M1 and M2 sites, respectively. The strongest cation preference was observed for Na1.158In1.184Sb0.658O4. In this instance, ∼88% of the octahedral Na+ cations sit on the M1 site according to the Rietveld refinement. In3+ and Sb5+ could preferentially occupy either of the sites too, but this cannot be determined by X-ray diffraction. NaSc1.5Sb0.5O4 also has a weak (101) reflection, and cation site preference is likely in this compound. However, because we do not know with certainty the amount of Na+ occupying the framework sites in this compound, it is difficult to identify the mechanism of this partitioning. Given these results, ordering of the framework cations in Na-CFs seems to be driven by differences in cationic radii and charge density rather than JT activity. It would be expected, then, that NaCd0.5Sn1.5O4 would show cation site preference. However, X-rays cannot distinguish between Cd2+ and Sn4+, thus the cation distribution could not be fully examined in the present work; a detailed NMR crystallography study of this question is ongoing. Neutron diffraction studies would also be a valuable complement to the work presented here, as X-ray data alone is insufficient for occupancy studies when the framework atoms differ little in electron count (e.g., NaCrTiO4 and NaCd0.5Sn1.5O4). Furthermore, each of the samples with cation site preference were quenched from the synthesis temperature. It is possible that annealing at lower temperature would increase the degree of site preference, as lower synthesis temperatures (hydrothermal synthesis) resulted in site preference in the Na–Fe–Ti–O CFs, (38,57) which was not observed for the NaFe0.5Ti1.5O4 compound synthesized by a solid-state reaction and presented in this paper.

Phase Space and Compositional Trends

CF-NaCr2O4, NaMn2O4, and NaRh2O4 have all been synthesized in the postspinel structure but required the use of high pressure. (12,32,33) The requirement of high pressure is likely due to a combination of factors. The compounds are mixed-valent, and the oxidation states of Cr and Rh therein are unusual; Rh4+ compounds often require high-pressure and highly oxidizing environments, while Cr4+ compounds also often require high pressure to avoid disproportionation into Cr3+ and Cr6+. Mn4+ can be formed at ambient pressures, but Mn3+ is strongly Jahn–Teller active. Thus, CF-NaMn2O4 would not be expected to be stable by the criteria suggested by Reid et al., which potentially explains the necessity of high pressure. Ionic radius also appears to play a role, and it should also be noted that no Na-CF synthesized under ambient pressure contains framework cations smaller than 0.60 Å (the size of Sb5+) (58) as major components. Rh4+ (0.60 Å) is similar in size to Sb5+, but both Cr4+ (0.55 Å) and Mn4+ (0.53 Å) are smaller. However, Cr3+ (0.615 Å) and Rh3+ (0.665 Å) are JT-inactive (spherical), stable, and have ionic radii consistent with cations known to form CF compounds at ambient pressure. Thus, we reasoned that replacement of Cr4+ and Rh4+ with a more stable and larger tetravalent cation like Ti4+ (0.605 Å) or Sn4+ (0.69 Å) should increase the chance of synthesizing a CF compound at ambient pressure. This strategy was successful in the synthesis of NaV3+(V0.25Ti0.75)4+O4 and NaVSnO4. (31) Indeed, it was found that Cr3+ can form a Na-CF phase when paired with either Ti4+ or Sn4+. Rh3+ formed a Na-CF phase when combined with Ti4+, though we were unable to make CF-NaRhSnO4. Reid et al. found that CF-NaMnTiO4 did not form at ambient pressure, and this result was among the evidence that the CF structure tends to form with spherical ions. Instead, a mixture corresponding to the stoichiometry NaMnTiO4 forms Na4Mn4Ti5O18 at ambient pressure. (59) This structure contains two types of tunnels: a smaller tunnel with a shape reminiscent of those found in the CF structure, and a larger S-shaped tunnel. Half of the Mn3+ cations are 5-coordinate in this structure. Likewise, “NaMn2O4” forms the isostructural Na4Mn9O18 at ambient pressure. (60) Interestingly, replacing Ti4+ with Sn4+ results in CF-NaMnSnO4.
No Na-CF compound containing In3+ has previously been reported. Reid et al. attempted to synthesize NaInTiO4 but obtained a mixture of In2O3 and Na2Ti3O7. (8) Synthesis of NaInZrO4 was attempted in this work because NaScZrO4 was reported and the size mismatch would be decreased, but that did not form either. However, NaInSnO4 was successfully synthesized. It should be noted that the ionic radius of In3+ (0.800 Å) is larger than Sc3+ (0.745 Å), which means In3+ is the largest trivalent cation known to form a Na-CF compound at ambient pressure. Y3+ (0.900 Å) is apparently too large. The compound NaYTiO4 has a layered perovskite structure instead of the CF structure. (61) In the yttrium system, replacing Ti4+ with Sn4+ or Zr4+ still did not produce a CF phase.
Only two compounds of the type NaA2+0.5B4+1.5O4 have been reported previously: NaCo0.5Ti1.5O4 and NaNi0.5Ti1.5O4. (35,36) Given the trends established by Reid et al., other combinations of metal cations should be possible. It was found that Co2+ and Ni2+ could be replaced by Mg2+ (JT-inactive) or Fe2+ (weakly JT-active). No CF phases formed when the compositions NaA2+0.5Ti1.5O4 (A2+ = Mn2+, Cu2+, and Zn2+) were targeted. That CF-NaCu0.5Ti1.5O4 could not be synthesized at ambient pressure is consistent with the conclusion made by Reid et al., as Cu2+ is strongly JT-active. However, Mn2+ (d5 electron configuration) and Zn2+ are not JT-active. Mn2+ is quite large (0.83 Å), and Fe2+ (0.78 Å) is the largest divalent cation successfully substituted in the NaA2+0.5Ti1.5O4 compositions. It is easy to conclude that the size mismatch between the divalent cation and Ti4+ becomes too large when A2+ = Mn2+, destabilizing the CF phase, but the results of the NaA2+0.5Sn1.5O4 system cast doubt on this explanation. On the other hand, Zn2+ has an ionic radius (0.74 Å) within the range of the divalent metals substituted into NaA2+0.5Ti1.5O4. It is possible that the preference of Zn2+ for tetrahedral sites destabilizes the CF structure. A compound with lower Na and Zn content in the Na–Zn–Ti–O system, freudenbergite-type Na1.84Zn0.92Ti7.08O16, does contain octahedrally coordinated Zn2+. (62) However, the Zn2+ content is dilute compared to the Ti4+ in the freudenbergite compound. Furthermore, the increased sodium content in the CF compound would be expected to increase the covalency of the Zn–O bonds. More covalent Zn–O bonds are expected to favor tetrahedrally coordinated Zn2+.
No compounds of the type NaA2+0.5Sn1.5O4 have been reported previously. Given the existence of NaA2+0.5Ti1.5O4 compounds and Na-CFs containing Sn4+ (NaFeSnO4), it seemed probable NaA2+0.5Sn1.5O4 would also be stable. Notably, more CF-NaA2+0.5Sn1.5O4 compounds were found than CF-NaA2+0.5Ti1.5O4 compounds, with Mg2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+, and Cd2+ all forming Na-CF compounds when paired with Sn4+. As with NaMn3+SnO4, Sn4+ stabilizes the CF structure even when paired with the strongly JT-active Cu2+. Given the existence of NaMnSnO4 and NaCu0.5Sn1.5O4, it appears that spherical (JT-inactive) cations are not a necessary condition for a stable CF phase in the case of NaA2+0.5Sn1.5O4 and NaA3+SnO4. NaZn0.5Sn1.5O4 and NaMn0.5Sn1.5O4 can also be formed, unlike NaZn0.5Ti1.5O4 and NaMn0.5Ti1.5O4. It would be easy to conclude that NaMn0.5Sn1.5O4 forms because the difference in ionic radii between Sn4+ and Mn2+ is smaller than the difference between Ti4+ and Mn2+. However, we were also able to form NaCd0.5Sn1.5O4, and the difference in ionic radius between Cd2+ (0.95 Å) and Sn4+ (0.69 Å) is even larger than the difference between Mn2+ (0.83 Å) and Ti4+ (0.605 Å). The existence of NaCd0.5Sn1.5O4 makes Cd2+ the largest divalent framework cation known to occur in an Na-CF compound so far.
Clearly, Ti4+ and Sn4+ behave quite differently in the Na-CF phase space. Sn4+ stabilizes the CF structure more than Ti4+, as many of the cations that do not form a Na-CF with Ti4+ do form one when paired with Sn4+ instead, as is the case with NaMnSnO4, NaInSnO4, NaMn0.5Sn1.5O4, NaCu0.5Sn1.5O4, and NaZn0.5Sn1.5O4. The ability of Ti4+ to accommodate a high degree of octahedral distortion, owing to its unique combination of size and d0 electron configuration, (63−65) creates a large number of competing phases in the Ti systems not present in the Sn systems. For example, the Na2O–MgO–TiO2 phase diagram contains at least eight reported quaternary phases (including NaMg0.5Ti1.5O4), (52,66) whereas, to the best of our knowledge, the NaMg0.5Sn1.5O4 reported in this paper is the only known phase in the Na2O–MgO–SnO2 system. Since many of the phases containing highly distorted TiO6 octahedra do not have Sn4+ analogues, such as the freudenbergite structure, (67) it becomes more likely that the CF phase is on the thermodynamic convex hull in the Na2O–AO/A2O3–SnO2 systems. Another obvious difference between Ti4+ and Sn4+ is their ionic radii (0.605 and 0.69 Å, respectively). This difference was invoked by Chiring et al. to explain the stability of NaMnSnO4. (25) It was suggested Sn4+ exerted chemical pressure on Mn3+, stabilizing the octahedral configuration, as opposed to the square-pyramidal coordination observed in Na4Mn4Ti5O18. Alternatively, one could say the framework sites in NaMnSnO4 are larger than in the hypothetical NaMnTiO4, which enhances the stability of octahedral Mn3+. This reasoning might also be applied to NaZn0.5Sn1.5O4 to explain how pairing Zn2+ with Sn4+ achieves the desired octahedral coordination of Zn2+ instead of tetrahedral coordination. Notably, tetrahedral Zn2+ (0.60 Å) has a radius closer to that of octahedral Ti4+ (0.605 Å) than octahedral Sn4+ (0.69 Å), and octahedral Zn2+ (0.74 Å) has a radius closer to that of octahedral Sn4+ than octahedral Ti4+. Thus, mixing of Sn4+ with Zn2+ on octahedral sites seems more favorable, allowing for ZnO6 octahedra. However, it is difficult to say if the “stabilization” of the CF phase is not simply a result of the destabilization of competing phases such as Na4Mn4Ti5O18 upon Sn4+ substitution owing to the preference of Sn4+ for more symmetric octahedra or if both factors are important.
NaFe1.5Sb0.5O4 was the only reported Na-CF of the type NaA3+1.5B5+0.5O4. We found that CF phases also exist in the Na+-Sc3+-Sb5+-O2– and Na+-In3+-Sb5+-O2– systems. However, mixtures with the ideal compositions corresponding to NaSc1.5Sb0.5O4 and NaIn1.5Sb0.5O4 do not result in phase purity, at least under the synthetic conditions explored. In the case of Na1.16In1.18Sb0.66O4, some of the Na+ (1.02 Å) occupies the framework sites. It is likely the larger size of In3+ (0.800 Å) relative to Fe3+ (0.645 Å) allows the framework-site mixing. Sc3+ (0.745 Å) is somewhat smaller than In3+, which is consistent with “NaSc1.5Sb0.5O4” having a smaller degree of Na+ substitution on the octahedral sites. Interestingly, NaCr1.5Sb0.5O4 could not be prepared at either 900 or 950 °C. P3 phases in the Na1–xCr1–x/2Sbx/2O2 (0.42 ≤ x ≤ 0.5) system exist which are very close in composition to the target NaCr1.5Sb0.5O4, so it is likely the enhanced stability of highly Na-vacant layered phases is responsible for the absence of NaCr1.5Sb0.5O4. (68) In contrast, the layered Na1–xFe1–x/2Sbx/2O2 phase is stable only down to x = 0.25, and the CF phase is observed when lower x values are attempted. (69)
While solid solutions between NaScTiO4 and NaFeTiO4 have been successfully synthesized, to the best of our knowledge, no other Na-CF solid solutions have been reported to form via ambient-pressure synthesis. We successfully synthesized three NaA3+B4+O4–NaA2+0.5B4+1.5O4 compositions: NaCo1/3Fe1/3Ti4/3O4, NaNi1/3Sc1/3Ti4/3O4, and NaNi1/3Fe1/3Ti4/3O4. While other solid solutions were not attempted, these results suggest numerous solid solution series are accessible.
Given the results presented here and in the context of previous literature, we suggest that the phase space of Na-CFs might be expanded even more through hydrothermal synthesis by the introduction of more cation site order and stoichiometry nonhomogeneity. Some of the Na-CFs reported here show cation preference between the two framework sites, even though the compounds were quenched from high temperature. If these can be synthesized via hydrothermal synthesis, it is likely the degree of cation preference and sodium vacancies can be enhanced, as in the case of some of the Na0.55Fe0.28Ti1.72O4 and Na1–xFe1–xTi1+xO4. (38,57) More superstructures of the CF structure may also be discovered, such as in the case of Na3Mn4Te2O12, (39) and would be more likely when the framework cations have greater differences in charge density. Finally, some cation combinations that do not produce a CF at high temperature might be stabilized under, e.g., hydrothermal conditions.
The unit cell volumes for the new compounds range from 284.872(1) Å3 for Na0.99Cr0.99Ti1.01O4 to 341.995(17) Å3 for NaCd0.5Sn1.5O4. As would be expected, the unit cell volume generally increases as the weighted-average framework cation radius increases (see Figure 7 and Table 1). The Na–O bond lengths for the NaO8 bicapped trigonal prisms increase as the unit cell volume increases. For Na0.99Cr0.99Ti1.01O4, the Na–O bond lengths range from 2.378(1) Å to 2.568(1) Å. For NaNi0.5Sn1.5O4, with a unit cell volume of 323.483(2) Å3, the Na–O bond lengths range from 2.444(2) Å to 2.644(2) Å. For Na0.96Sn0.96Sn1.04O4, with a unit cell volume of 341.504(2) Å3, the Na–O bond lengths range from 2.472(2) Å to 2.728(3) Å (see Table S2). This may have implications for Na+ mobility.

Figure 7

Figure 7. Unit cell volume of the new Na-CFs versus the weighted-average effective ionic radius of the framework cations in octahedral coordination. (58)

Comparison to Lithium Spinels

A natural comparison to Na-CF compounds are the lithium spinels. Since Li+ is smaller than Na+, it favors lower coordination numbers and shorter bond distances to oxygen, stabilizing the spinel structure as opposed to the CF structure. The spinel structure contains two primary cation sites, the tetrahedral 8a site and the octahedral 16d site. The octahedral sites comprise the framework; a 3D series of interconnected tunnels are formed by the occupied tetrahedral sites and empty octahedral 16c sites. Because Li+ can readily accommodate coordination numbers of four and six, it can occupy both the tetrahedral 8c and octahedral 16d sites. This results in antisite defects, or inversion, in which Li+ and another cation that does not have a high octahedral site preference are statically distributed among the 8c and 16d sites according to their relative site preferences. In contrast, this does not occur in Na-CFs to a measurable extent. While Na+, like Li+, varies in its coordination number and geometry for known materials and is known to occupy both 6- and 8-coordinate sites; the smaller, more highly charged cations like Ti4+ are not expected to occupy 8-coordinate sites except at very high pressures. In addition, the larger difference in size between Na+ and the CF framework cations relative to Li+ and spinel framework cations increases the site preferences. Thus, in LiFeTiO4, significant inversion is observed, with Li+ and Fe3+ essentially being randomly distributed. (70) In contrast, in NaFeTiO4, Na+ solely occupies the 8-coordinate tunnel sites, and Fe3+ and Ti4+ exclusively occupy the octahedral framework sites. (8) For energy storage applications, the lack of antisite defects in Na-CF’s is expected to be beneficial. (71) Inversion in spinels results in highly charged, immobile cations occupying the tetrahedral sites, which blocks the lowest energy diffusion path, hindering electrochemical performance. (72−74) This likely explains why spinel-LiFeTiO4, with a high degree of inversion, has both a high activation energy for ionic conduction and a low reversible specific capacity, (70,75) whereas CF-LiFeTiO4, synthesized by ion exchange from NaFeTiO4, nearly achieves theoretical capacity. (19) Furthermore, Na+ is known to be exchangeable with Li+ topotactically for some Na-CFs, meaning Li-CFs can be accessed at ambient pressure. (19,20,23)
The ability of Li+ to occupy both 4- and 6- coordinate sites can also be exploited to synthesize Li-rich spinels such as Li4Mn5O12 and Li4Ti5O12, in which Li+ occupies all tetrahedral positions and some of the octahedral framework sites as well. (54,55) Analogous Na-CFs, in which Na would occupy all the 8-coordinate tunnel sites and some of the octahedral sites, were previously unreported. Na1.16In1.18Sb0.66O4 appears to be the first such example, with ∼8% of the framework sites occupied by Na+, with Na–Sc–Sb–O and Na–Cd–Sn–O also appearing to be further phases with framework Na+. However, framework sodium appears to be nonexistent for Na-CFs with redox-active transition metal cations. This is likely because the redox-active transition metals found in CFs are significantly smaller than In3+/Cd2+/Sc3+, which decreases the average size of the framework sites.

Thermodynamic Calculations

Density functional theory (DFT) calculations were used to further understand the stability of phases in the Na-CF chemical space. The energy of the CF phase relative to the normal spinel phase (Fdm) as well as the energy above the convex hull of stability, Ehull, for the CF phases were computed for various combinations of octahedral cations and at varying levels of sodiation. Combinations were chosen to include a range of successfully synthesized Na-CFs as well as compositions not expected to produce a CF phase (e.g., NaCrZrO4). Calculated Ehull values of the Na-CF phase at 0 K and without considering entropic effects are shown in Figure 8a. If the two octahedral sites are equivalently occupied by the A and B cations in the Na-CF phases, ideal mixing would suggest configurational entropy provides significant stabilization at 1200 K (the approximate synthesis temperature), lowering the energy by ∼21 meV/atom for the NaA3+B4+O4 compositions and ∼17 meV/atom for the NaA2+0.5B4+1.5O4 compositions. Figure 8b shows Ehull after including this configurational entropy contribution for fully sodiated compounds. Each box in this panel is colored according to the experimentally observed synthesis ability, with green boxes indicating successful synthesis, red boxes indicating failed synthesis, and no box indicating no known synthesis attempt. When including configurational entropy, 8 of the 12 synthesized postspinels are predicted to be on the convex hull (stable), with three more lying very close to the hull (<5 meV/atom). The notable exception is NaCo0.5Ti1.5O4, which is calculated to lie 13 meV/atom above the hull. Of the six compounds that could not be synthesized in the postspinel structure, NaMnTiO4 and NaRhSnO4 are predicted to be stable, while the others are predicted to be unstable with respect to decomposition into competing phases. The thermodynamic stability of NaMnTiO4 and NaRhSnO4, which could not be synthesized, and the thermodynamic instability of NaCo0.5Ti1.5O4 emphasizes the role of kinetics and metastability in the synthesis of these oxides. (76,77) We note that the calculated lack of stability of NaCo0.5Ti1.5O4 may arise from the inability to converge this structure with Co2+ in a high-spin state, as would be expected for CoO6 octahedra with Co2+ and was observed in our calculations for the other Co2+-containing CF phases (NaCo0.5Sn1.5O4 and NaCo0.5Zr1.5O4).

Figure 8

Figure 8. (a) Computed stability of NaxA3+B4+O4 and NaxA2+0.5B4+1.5O4 CF phases with three degrees of sodiation (x = 0, 0.5, 1.0). (b) Energy above hull of each xNa = 1 compound after including ideal mixing entropy for the octahedral cations at 1200 K. Green and red boxes show successful/failed syntheses at ambient pressure. (c) Energy difference between the CF and spinel structure. Note that the empty boxes correspond to In-containing compounds that failed to converge in either the CF or spinel structures and were consequently excluded from this analysis.

No new CF compounds containing Zr4+ were successfully synthesized, which is consistent with the high Ehull values calculated for the CF phases containing Zr4+ (ranging from 26 meV/atom for NaMn0.5Zr1.5O4 to 71 meV/atom for NaCrZrO4). After factoring in configurational entropy for the CF phase at 1200 K, all CF phases with Zr still remain above the hull. The stability of NaA2+0.5Zr1.5O4 increases as the size of A2+ increases, and NaMn0.5Zr1.5O4 is only 9 meV/atom above the hull when factoring in configurational entropy. The large size of Zr4+ and its ability to have coordination numbers higher than six likely destabilizes the CF structure relative to competing phases including ZrO2, in which Zr is 7-fold coordinated. In each attempted synthesis of the NaA3+ZrO4 and NaA2+0.5Zr1.5O4 phases, baddeleyite ZrO2 or a higher symmetry, partially substituted ZrO2 phase is formed as the main phase. NaScZrO4 appears to be a special case and remains the only known Na-CF containing Zr4+. Sn4+- and Ti4+-containing compounds have lower Ehull than those with Zr4+, in agreement with experiments that found many possible combinations including Sn and Ti. Considering that Sn4+ (0.69 Å) and Zr4+ (0.72 Å) have similar effective ionic radii, the increased stability of Sn4+-containing CF phases might originate because Sn4+ prefers the octahedral site more than Zr4+ or from the prevalence of low-energy zirconium oxide competing phases. Some of the newly synthesized phases contain redox-active metals and may be of interest as battery electrode materials, so thermodynamic calculations were also used to explore the (in)stability of the CF phase upon desodiation (Figure 8a). As expected, removing the Na+ cations destabilizes the CF structure to some degree, with values for the completely desodiated CF phases ranging from 68 meV/atom for MnTiO4 to 234 meV/atom for Ni0.5Zr1.5O4. The stability of the empty CF framework was compared to the empty spinel framework of the same composition (Figure 8c). Interestingly, the empty CF structure is more stable in every case examined here. Spinels are well-studied as Li-battery electrodes. While the spinel anodes like LiTi2O4, Li4Ti5O12, and LiCrTiO4 perform well when additional lithium is inserted and re-extracted, (78,79) when full removal of the already-present Li+ is attempted, many lithium spinels either have limited capacity or show irreversible phase transitions that do not preserve the spinel lattice. This has been observed in LiTi2O4, LiV2O4, LiVTiO4, and LiCrTiO4. (79−82) The calculations presented here indicate the CF framework is more stable when fully charged than the spinel framework. In fact, the charged CF-CrTiO4 phase has an Ehull about one-half that of the charged spinel-CrTiO4 phase (98 meV/atom vs 195 meV/atom). Given the variety of CF phases explored in these calculations, this is likely a general phenomenon and may extend to many more if not all CF/spinel compositions. Thus, CF phases appear to be promising as energy storage materials.

Conclusions

Click to copy section linkSection link copied!

The phase space of Na-containing CaFe2O4-type compounds has been expanded to include 16 new compositions and several additional solid-solutions. Previously it was suggested that only cations without Jahn–Teller distortions (spherical cations) form CaFe2O4-type compounds with sodium in the tunnels, but the existence of NaMnSnO4 and NaCu0.5Sn1.5O4 shows that this “requirement” is relaxed when B4+ is Sn4+. However, tetravalent cations larger than Sn4+ do not effectively stabilize the CaFe2O4 structure, and NaScZrO4 and NaScHfO4 remain the only known Na-CFs when the tetravalent cation is larger than Sn4+. In most cases, the framework cations are randomly distributed among the two framework sites, but weak site preference was observed for NaMg0.5Ti1.5O4 and NaCu0.5Sn1.5O4. Only in one case did Rietveld refinement and 23Na NMR spectroscopy indicate strong site preference: the first known CaFe2O4-type compound with Na+ occupying framework sites, Na1.16In1.18Sb0.66O4. In this material, the large Na+ cation strongly prefers one of the two symmetrically distinct framework sites, suggesting the order is driven by differences in charge density. “NaSc1.5Sb0.5O4,” which always contained secondary phases, may also contain Na+ in the framework sites, albeit to a lesser degree than Na1.16In1.18Sb0.66O4. DFT calculations revealed that most of the successfully synthesized Na-CFs were on or near their respective convex hull, with the exception of NaCo0.5Ti1.5O4, whereas the hypothetical Na-CFs containing Zr4+ were much higher in energy, suggesting the Na-CF compounds containing Zr4+ are thermodynamically unstable. Additional DFT calculations show that the charged (desodiated) CF framework is more stable than a charged spinel framework of the same composition, suggesting an opportunity for postspinel phases as Li/Na/Mg electrode materials. Given the picture described here, the richness of this phase space can likely be further expanded by synthesizing solid solutions as well as using hydrothermal and other soft chemical methods. The growing library of CaFe2O4-type materials inspires future fundamental and applied studies on these materials and related phase spaces.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsorginorgau.1c00019.

  • Table of attempted syntheses and products, Rietveld refinements, additional 23Na NMR spectra, and table of Na–O bond distances for selected compounds (PDF)

Accession Codes

CCDC 2103718, 21038252103831, 21038332103835, 2103925, 21039282103929, 2103933, 2103935, and 2103938 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Kenneth R. Poeppelmeier - Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United StatesJoint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesOrcidhttps://orcid.org/0000-0003-1655-9127 Email: [email protected]
  • Authors
    • Justin C. Hancock - Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United StatesJoint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesOrcidhttps://orcid.org/0000-0002-3610-291X
    • Kent J. Griffith - Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United StatesJoint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesOrcidhttps://orcid.org/0000-0002-8096-906X
    • Yunyeong Choi - Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesDepartment of Materials Science and Engineering, University of California, Berkeley, California 94720, United States
    • Christopher J. Bartel - Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesDepartment of Materials Science and Engineering, University of California, Berkeley, California 94720, United StatesOrcidhttps://orcid.org/0000-0002-5198-5036
    • Saul H. Lapidus - Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesX-ray Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United StatesOrcidhttps://orcid.org/0000-0002-7486-4325
    • John T. Vaughey - Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesChemical Sciences and Engineering Division, Argonne National Laboratory, Lemont, Illinois 60439, United StatesOrcidhttps://orcid.org/0000-0002-2556-6129
    • Gerbrand Ceder - Joint Center for Energy Storage Research, Argonne National Laboratory, Argonne, Illinois 60439, United StatesDepartment of Materials Science and Engineering, University of California, Berkeley, California 94720, United StatesMaterials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United StatesOrcidhttps://orcid.org/0000-0001-9275-3605
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by the Joint Center for Energy Storage Research (JCESR), an Energy Innovation Hub funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences. Use of the Advanced Photon Source at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357. This work made use of the Jerome B. Cohen X-ray Diffraction Facility supported by the MRSEC program of the National Science Foundation (Grant DMR-1720139) at the Materials Research Center of Northwestern University and the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (Grant NSF ECCS-2025633). This work made use of the IMSERC NMR facilities at Northwestern University, which have received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (Grant NSF ECCS-2025633), International Institute of Nanotechnology, and Northwestern University. This research used resources of the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility located at Lawrence Berkeley National Laboratory, operated under Contract No. DE-AC02-05CH11231. Computational resources were also provided by the Extreme Science and Engineering Discovery Environment (XSEDE) resource Stampede2 through Allocation TG-DMR970008S, which is supported by the National Science Foundation Grant Number ACI1053575.

References

Click to copy section linkSection link copied!

This article references 82 other publications.

  1. 1
    Bertraut, E.-F.; Blum, P.; Magnano, G. Structure des vanadite, chromite et ferrite monocalciques. Bull. Mineral. 1956, 79, 536561
  2. 2
    Hill, P. M.; Peiser, H. S.; Rait, J. R. The crystal structure of calcium ferrite and β calcium chromite. Acta Crystallogr. 1956, 9, 981986,  DOI: 10.1107/S0365110X56002862
  3. 3
    Irifune, T.; Fujino, K.; Ohtani, E. A new high-pressure form of MgAl2O4. Nature 1991, 349, 409411,  DOI: 10.1038/349409a0
  4. 4
    Funamori, N.; Jeanloz, R.; Nguyen, J. H.; Kavner, A.; Caldwell, W. A.; Fujino, K.; Miyajima, N.; Shinmei, T.; Tomioka, N. High-pressure transformations in MgAl2O4. J. Geophys. Res. Solid Earth 1998, 103, 2081320818,  DOI: 10.1029/98JB01575
  5. 5
    Liu, L. High pressure NaAlSiO4: the first silicate calcium ferrite isotype. Geophys. Res. Lett. 1977, 4, 183186,  DOI: 10.1029/GL004i005p00183
  6. 6
    Yamada, H.; Matsui, Y.; Ito, E. Crystal-chemical characterization of NaAlSiO4 with the CaFe2O4 structure. Mineral. Mag. 1983, 47, 177181,  DOI: 10.1180/minmag.1983.047.343.07
  7. 7
    Reid, A. F.; Wadsley, A. D.; Ringwood, A. E. High pressure NaAlGeO4, a calcium ferrite isotype and model structure of silicates at depth in the earth’s mantle. Acta Crystallogr. 1967, 23, 736739,  DOI: 10.1107/S0365110X67003627
  8. 8
    Reid, A. F.; Wadsley, A. D.; Sienko, M. J. Crystal chemistry of sodium scandium titanate, NaScTiO4, and its isomorphs. Inorg. Chem. 1968, 7, 112118,  DOI: 10.1021/ic50059a024
  9. 9
    Ishii, T.; Sakai, T.; Kojitani, H.; Mori, D.; Inaguma, Y.; Matsushita, Y.; Yamaura, K.; Akaogi, M. High-pressure phase relations and crystal structures of postpinel phases in MgV2O4, FeV2O4, and MnCr2O4: crystal chemistry of AB2O4 postspinel compounds. Inorg. Chem. 2018, 57, 66486657,  DOI: 10.1021/acs.inorgchem.8b00810
  10. 10
    Yamaura, K.; Huang, Q.; Zhang, L.; Takada, K.; Baba, Y.; Nagai, T.; Matsui, Y.; Kosuda, K.; Takayama-Muromachi, E. Spinel-to-CaFe2O4-type structural transformation in LiMn2O4 under higher pressure. J. Am. Chem. Soc. 2006, 128, 94489456,  DOI: 10.1021/ja0612302
  11. 11
    Yamaura, K.; Arai, M.; Sato, A.; Karki, A. B.; Young, D. P.; Movshovich, R.; Okamoto, S.; Mandrus, D.; Takayama-Muromachi NaV2O4: a quasi-1D metallic antiferromagnet with half-metallic chains. Phys. Rev. Lett. 2007, 99, 196601,  DOI: 10.1103/PhysRevLett.99.196601
  12. 12
    Sakurai, H.; Kolodiazhnyi, T.; Michiue, Y.; Takayama-Muromachi, E.; Tanabe, Y.; Kikuchi, H. Unconventional colossal magnetoresistance in sodium chromium oxide with a mixed-valence state. Angew. Chem. 2012, 124, 67576760,  DOI: 10.1002/ange.201201884
  13. 13
    Young, O.; Wildes, A. R.; Manuel, P.; Ouladdiaf, B.; Khalyavin, D. D.; Balakrishnan, G.; Petrenko, O. A. Highly frustrated magnetism in SrHo2O4: coexistence of two types of short-range order. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 024411,  DOI: 10.1103/PhysRevB.88.024411
  14. 14
    Dutton, S. E.; Broholm, C. L.; Cava, R. J. Divergent effects of static disorder and hole doping in geometrically frustrated β-CaCr2O4. J. Solid State Chem. 2010, 183, 17981804,  DOI: 10.1016/j.jssc.2010.05.032
  15. 15
    Arévalo-López, Á. M.; Dos santos-García, A. J.; Castillo-Martínez, E.; Durán, A.; Alario-Franco, M. Á. Spinel to CaFe2O4 transformation: Mechanism and properties of β-CdCr2O4. Inorg. Chem. 2010, 49, 28272833,  DOI: 10.1021/ic902228h
  16. 16
    Shimomura, Y.; Kurushima, T.; Kijima, N. Photoluminescence and crystal structure of green-emitting phosphor CaSc2O4:Ce3+. J. Electrochem. Soc. 2007, 154, J234,  DOI: 10.1149/1.2741172
  17. 17
    Hao, Z.; Zhang, J.; Zhang, X.; Lu, S.; Wang, X. Blue-green-emitting phosphor CaSc2O4:Tb3+: tunable luminescence manipulated by cross-relaxation. J. Electrochem. Soc. 2009, 156, H193,  DOI: 10.1149/1.3060382
  18. 18
    Hao, Z.; Zhang, J.; Zhang, X.; Wang, X. CaSc2O4:Eu3+: A tunable full-color emitting phosphor for white light emitting diodes. Opt. Mater. 2011, 33, 355358,  DOI: 10.1016/j.optmat.2010.09.035
  19. 19
    Bruno, S. R.; Blakely, C. K.; Clapham, J. B.; Davis, J. D.; Bi, W.; Alp, E. E.; Poltavets, V. V. Synthesis and electrochemical properties of novel LiFeTiO4 and Li2FeTiO4 polymorphs with the CaFe2O4-type structures. J. Power Sources 2015, 273, 396403,  DOI: 10.1016/j.jpowsour.2014.09.115
  20. 20
    Jung, Y. H.; Kim, D. K.; Hong, S.-T. Synthesis, structure, and electrochemical Li-ion intercalation of LiRu2O4 with CaFe2O4-type structure. J. Power Sources 2013, 233, 285289,  DOI: 10.1016/j.jpowsour.2013.01.119
  21. 21
    Sun, X.; Blanc, L.; Nolis, G. M.; Bonnick, P.; Cabana, J.; Nazar, L. F. NaV1.25Ti0.75O4: a potential post-spinel cathode material. Chem. Mater. 2018, 30, 121128,  DOI: 10.1021/acs.chemmater.7b03383
  22. 22
    Mukai, K.; Uyama, T.; Yamada, I. Structural and electrochemical analyses on the transformation of CaFe2O4-type LiMn2O4 from spinel-type LiMn2O4. ACS Omega 2019, 4, 64596467,  DOI: 10.1021/acsomega.9b00588
  23. 23
    Nolis, G.; Gallardo-Amores, J. M.; Serrano-Sevillano, J.; Jahrman, E.; Yoo, H. D.; Hu, L.; Hancock, J. C.; Bolotnikov, J.; Kim, S.; Freeland, J. W.; Liu, Y.-S.; Poeppelmeier, K. R.; Seidler, G. T.; Guo, J.; Alario-Franco, M. A.; Casas-Cabanas, M.; Morán, E.; Cabana, J. Factors defining the intercalation electrochemistry of CaFe2O4-type manganese oxides. Chem. Mater. 2020, 32, 82038215,  DOI: 10.1021/acs.chemmater.0c01858
  24. 24
    Liu, X.; Wang, X.; Iyo, A.; Yu, H.; Li, D.; Zhou, H. High stable post-spinel NaMn2O4 cathode of sodium ion battery. J. Mater. Chem. A 2014, 2, 1482214826,  DOI: 10.1039/C4TA03349C
  25. 25
    Chiring, A.; Senguttuvan, P. Chemical pressure-stabilized post spinel-NaMnSnO4 as potential cathode for sodium-ion batteries. Bull. Mater. Sci. 2020, 43, 306,  DOI: 10.1007/s12034-020-02203-6
  26. 26
    Ling, C.; Mizuno, F. Phase stability of post-spinel compound AMn2O4 (A = Li, Na, or Mg) and its application as a rechargeable battery cathode. Chem. Mater. 2013, 25, 30623071,  DOI: 10.1021/cm401250c
  27. 27
    Hannah, D. C.; Sai Gautam, G.; Canepa, P.; Rong, Z.; Ceder, G. Magnesium ion mobility in post-spinels accessible at ambient pressure. Chem. Commun. 2017, 53, 51715174,  DOI: 10.1039/C7CC01092C
  28. 28
    Dompablo, M. E. A.; Krich, C.; Nava-Avendaño, J.; Biškup, N.; Palacín, M. R.; Bardé, F. A joint computational and experimental evaluation of CaMn2O4 polymorphs as cathode materials for Ca ion batteries. Chem. Mater. 2016, 28, 68866893,  DOI: 10.1021/acs.chemmater.6b02146
  29. 29
    Müller-Buschbaum, H. The crystal chemistry of AM2O4 oxometallates. J. Alloys Compd. 2003, 349, 49104,  DOI: 10.1016/S0925-8388(02)00925-8
  30. 30
    Akimoto, J.; Takei, H. Synthesis and crystal structure of NaTi2O4: a new mixed-valence sodium titanate. J. Solid State Chem. 1989, 79, 212217,  DOI: 10.1016/0022-4596(89)90268-5
  31. 31
    Darriet, J.; Vidal, A. Les composés NaRu2O4 et NaFeRuO4. Structure cristalline de NaFeRuO4. Bull. Soc. Fr. Minéral Cristallogr. 1975, 98, 374377
  32. 32
    Viciu, L.; Ryser, A.; Mensing, C.; Bos, J.-W. G. Ambient-pressure synthesis of two new vanadium-based calcium ferrite-type compounds: NaV1.25Ti0.75O4 and NaVSnO4. Inorg. Chem. 2015, 54, 72647271,  DOI: 10.1021/acs.inorgchem.5b00655
  33. 33
    Akimoto, J.; Awaka, J.; Kijima, N.; Takahashi, Y.; Maruta, Y.; Tokiwa, K.; Watanabe, T. High-pressure synthesis and crystal structure analysis of NaMn2O4 with the calcium ferrite-type structure. J. Solid State Chem. 2006, 179, 169174,  DOI: 10.1016/j.jssc.2005.10.020
  34. 34
    Yamaura, K.; Huang, Q.; Moldovan, M.; Young, D. P.; Sato, A.; Baba, Y.; Nagai, T.; Matsui, Y.; Takayama-Muromachi, E. High-pressure synthesis, crystal structure determination, and a Ca substitutional study of the metallic rhodium oxide NaRh2O4. Chem. Mater. 2005, 17, 359365,  DOI: 10.1021/cm0483846
  35. 35
    Shukaev, I. L.; Volochaev, V. A. Ternary sodium and titanium oxides with cobalt(II). Russ. J. Inorg. Chem. 1995, 12, 19741980
  36. 36
    Nalbandyan, V. B.; Shukaev, I. L. Triple oxides of sodium, nickel, and titanium. Russ. J. Inorg. Chem. 1992, 37, 12311235
  37. 37
    Archaimbault, F.; Choisnet, J.; Rautureau, M. New ferriantimonates with the CaFe2O4 type structure: Na2Fe3SbO8 and isomorphous series Na2Fe2+xSn2–2xSbxO8 (0 ≤ x ≤ 1). Eur. J. Solid State Inorg. Chem. 1988, 25, 573587
  38. 38
    Ishiguro, T.; Tanaka, K.; Marumo, F.; Ismail, M. G. M. U.; Hirano, S.; Somiya, S. Non-stoichiometric sodium iron (II) titanium (IV) oxide. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1978, 34, 33463348,  DOI: 10.1107/S0567740878010870
  39. 39
    Feger, C. R.; Kolis, J. W. Na3Mn4Te2O12. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, 54, 10551057,  DOI: 10.1107/S0108270197020076
  40. 40
    Toby, B. H.; Von Dreele, R. B. GSAS-II: the genesis of a modern open-source all purpose crystallography software package. J. Appl. Crystallogr. 2013, 46, 544549,  DOI: 10.1107/S0021889813003531
  41. 41
    Amoureux, J.-P.; Fernandez, C.; Steuernagel, S. Z Filtering in MQMAS NMR. J. Magn. Reson., Ser. A 1996, 123, 116118,  DOI: 10.1006/jmra.1996.0221
  42. 42
    Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
  43. 43
    Kresse, G.; Hafner, J. Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements. J. Phys.: Condens. Matter 1994, 6, 8245,  DOI: 10.1088/0953-8984/6/40/015
  44. 44
    Blöchl, P. E.; Jepsen, O.; Andersen, O. K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 16223,  DOI: 10.1103/PhysRevB.49.16223
  45. 45
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  46. 46
    Wang, L.; Maxisch, T.; Ceder, G. Oxidation energies of transition metal oxides within the GGA+U framework. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73, 195107,  DOI: 10.1103/PhysRevB.73.195107
  47. 47
    Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; Persson, K. A. Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002,  DOI: 10.1063/1.4812323
  48. 48
    Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.; Persson, K. A.; Ceder, G. Formation enthalpies by mixing GGA and GGA + U calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 045115,  DOI: 10.1103/PhysRevB.84.045115
  49. 49
    Ong, S. P.; Richards, W. D.; Jain, A.; Hautier, G.; Kocher, M.; Cholia, S.; Gunter, D.; Chevrier, V. L.; Persson, K. A.; Ceder, G. Python Materials Genomics (pymatgen): A robust, open-source python library for materials analysis. Comput. Mater. Sci. 2013, 68, 314319,  DOI: 10.1016/j.commatsci.2012.10.028
  50. 50
    Avdeev, M. Yu; Nalbandyan, V. B.; Medvedev, B. S. Hexagonal sodium titanate chromite: a new high-conductivity solid electrolyte. Inorg. Mater. 1997, 33, 500503
  51. 51
    Mumme, W. G.; Reid, A. F. Non-stoichiometric sodium iron titanate, NaxFexTi2-xO4, 0.90> x> 0.75. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1968, 24, 625631,  DOI: 10.1107/S0567740868002955
  52. 52
    Nalbandyan, V. B.; Rykalova, S. I.; Bikyashev, E. A.; Isakova, S. Yu Sodium titanium magnesium (zinc) ternary oxides. Zh. Neorg. Khim. 1989, 34, 23812386
  53. 53
    Shukaev, I. L.; Butova, V. V.; Chernenko, S. V.; Pospelov, A. A.; Shapovalov, V. V.; Guda, A. A.; Aboraia, A. M.; Zahran, H. Y.; Yahia, I. S.; Soldatov, A. V. New orthorhombic sodium iron(+2) titanate. Ceram. Int. 2020, 46, 44164422,  DOI: 10.1016/j.ceramint.2019.10.167
  54. 54
    Thackeray, M. M.; de Kock, A.; Rossouw, M. H.; Liles, D.; Bittihn, R.; Hoge, D. Spinel electrodes from the Li-Mn-O system for rechargeable lithium battery applications. J. Electrochem. Soc. 1992, 139, 363366,  DOI: 10.1149/1.2069222
  55. 55
    Deschanvres, A.; Raveau, B.; Sekkal, Z. Mise en evidence et etude cristallographique d’une nouvelle solution solide de type spinelle Li1+xTi2–xO4 0 ⩽ x ⩽ 0, 333. Mater. Res. Bull. 1971, 6, 699704,  DOI: 10.1016/0025-5408(71)90103-6
  56. 56
    Rickert, K.; Sedefoglu, N.; Malo, S.; Caignaert, V.; Kavak, H.; Poeppelmeier, K. R. Structural, electrical, and optical properties of the tetragonal, fluorite-related Zn0.456In1.084Ge0.460O3. Chem. Mater. 2015, 27, 50725079,  DOI: 10.1021/acs.chemmater.5b01724
  57. 57
    Marinkovic, B. A.; Mancic, L.; Jardim, P. M.; Milosevic, O.; Rizzo, F. Hydrothermal synthesis of NaxFexTi2-xO4 from natural ilmenite sand: a CaFe2O4 structure type compound. Solid State Commun. 2008, 145, 346350,  DOI: 10.1016/j.ssc.2007.12.006
  58. 58
    Shannon, R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 751767,  DOI: 10.1107/S0567739476001551
  59. 59
    Mumme, W. G. The structure of Na4Mn4Ti5O18. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1968, 24, 11141120,  DOI: 10.1107/S0567740868003778
  60. 60
    Parant, J.-P.; Olazcuaga, R.; Devalette, M.; Fouassier, C.; Hagenmuller, P. Sur quelques nouvelles phases de formule NaxMnO2 (x ≤ 1). J. Solid State Chem. 1971, 3, 111,  DOI: 10.1016/0022-4596(71)90001-6
  61. 61
    Toda, K.; Kameo, Y.; Kurita, S.; Sato, M. Crystal structure determination and ionic conductivity of layered perovskite compounds NaLnTiO4 (Ln = rare earth). J. Alloys Compd. 1996, 234, 1925,  DOI: 10.1016/0925-8388(95)01969-3
  62. 62
    Bruhn, G.; Beutel, S.; Pfaff, G.; Albert, B. Low-temperature synthesis of freudenbergite-type titanate bronzes from metal halides, crystal growth from molybdate flux, and crystal structure determination of Na1.84Zn0.92Ti7.08O16. J. Alloys Compd. 2015, 644, 783787,  DOI: 10.1016/j.jallcom.2015.05.076
  63. 63
    Kunz, M.; Brown, I. D. Out-of-center distortions around octahedrally coordinated d0 transition metals. J. Solid State Chem. 1995, 115, 395406,  DOI: 10.1006/jssc.1995.1150
  64. 64
    Ok, K. M.; Halasyamani, P. S.; Casanova, D.; Llunell, M.; Alemany, P.; Alvarez, S. Distortions in octahedrally coordinated d0 transition metal oxides: a continuous symmetry measures approach. Chem. Mater. 2006, 18, 31763183,  DOI: 10.1021/cm0604817
  65. 65
    Urban, A.; Abdellahi, A.; Dacek, S.; Artrith, N.; Ceder, G. Electronic-structure origin of cation disorder in transition-metal oxides. Phys. Rev. Lett. 2017, 119, 176402,  DOI: 10.1103/PhysRevLett.119.176402
  66. 66
    Shilov, G. V.; Atovmyan, L. O.; Volochaev, V. A.; Nalbandyan, V. B. Crystal structure and ionic conductivity of a new sodium magnesium titanium oxide. Kristallografiya 1999, 44, 10291033
  67. 67
    Ishiguro, T.; Tanaka, K.; Marumo, F.; Ismail, M. G. M. U.; Hirano, S.; Somiya, S. Freudenbergite. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1978, 34, 255256,  DOI: 10.1107/S0567740878002708
  68. 68
    Pospelov, A. A.; Nalbandyan, V. B. Preparation, crystal structures and rapid hydration of P2- and P3-type sodium chromium antimony oxides. J. Solid State Chem. 2011, 184, 10431047,  DOI: 10.1016/j.jssc.2011.03.011
  69. 69
    Politaev, V. V.; Nalbandyan, V. B. Subsolidus phase relations, crystal chemistry and cation-transport properties of sodium iron antimony oxides. Solid State Sci. 2009, 11, 144150,  DOI: 10.1016/j.solidstatesciences.2008.04.014
  70. 70
    Arillo, M. A.; Lopez, M. L.; Perez-Cappe, E.; Pico, C.; Viega, M. L. Crystal structure and electrical properties of LiFeTiO4 spinel. Solid State Ionics 1998, 107, 307312,  DOI: 10.1016/S0167-2738(97)00537-7
  71. 71
    Kang, K.; Meng, Y. S.; Bréger, J.; Grey, C. P.; Ceder, G. Electrodes with high power and high capacity for rechargeable lithium batteries. Science 2006, 311, 977980,  DOI: 10.1126/science.1122152
  72. 72
    Sai Gautam, G.; Canepa, P.; Urban, A.; Bo, S.-H.; Ceder, G. Influence of inversion on Mg mobility and electrochemistry in spinels. Chem. Mater. 2017, 29, 79187930,  DOI: 10.1021/acs.chemmater.7b02820
  73. 73
    Bayliss, R. D.; Key, B.; Sai Gautam, G.; Canepa, P.; Kwon, B. J.; Lapidus, S. H.; Dogan, F.; Adil, A. A.; Lipton, A. S.; Baker, P. J.; Ceder, G.; Vaughey, J. T.; Cabana, J. Probing Mg migration in spinel oxides. Chem. Mater. 2020, 32, 663670,  DOI: 10.1021/acs.chemmater.9b02450
  74. 74
    Kwon, B. J.; Yin, L.; Park, H.; Parajuli, R.; Kumar, K.; Kim, S.; Yang, M.; Murphy, M.; Zapol, P.; Liao, C.; Fister, T. T.; Klie, R. F.; Cabana, J.; Vaughey, J. T.; Lapidus, S. H.; Key, B. High voltage Mg-ion battery cathode via a solid solution Cr-Mn spinel oxide. Chem. Mater. 2020, 32, 65776587,  DOI: 10.1021/acs.chemmater.0c01988
  75. 75
    Robertson, A. D.; Tukamoto, H.; Irvine, J. T. S. Li1+xFe1–3xTi1+2xO4 (0.0 ≤ x ≤ 0.33) based spinels: possible negative electrode materials for future Li-ion batteries. J. Electrochem. Soc. 1999, 146, 39583962,  DOI: 10.1149/1.1392576
  76. 76
    Miura, A.; Bartel, C. J.; Goto, Y.; Mizuguchi, Y.; Moriyoshi, C.; Kuroiwa, Y.; Wang, Y.; Yaguchi, T.; Shirai, M.; Nagao, M.; Rosero-Navarro, N. C.; Tadanaga, K.; Ceder, G.; Sun, W. Observing and modeling the sequential pairwise reactions that drive solid-state ceramic synthesis. Adv. Mater. 2021, 33, 2100312,  DOI: 10.1002/adma.202100312
  77. 77
    Sun, W.; Dacek, S. T.; Ong, S. P.; Hautier, G.; Jain, A.; Richards, W. D.; Gamst, A. C.; Persson, K. A.; Ceder, G. The thermodynamic scale of inorganic crystalline metastability. Sci. Adv. 2016, 2, e1600225,  DOI: 10.1126/sciadv.1600225
  78. 78
    Colbow, K.M.; Dahn, J.R.; Haering, R.R. Structure and electrochemistry of the spinel oxides LiTi2O4 and Li4/3Ti5/3O4. J. Power Sources 1989, 26, 397402,  DOI: 10.1016/0378-7753(89)80152-1
  79. 79
    Ohzuku, T.; Tatsumi, K.; Matoba, N.; Sawai, K. Electrochemistry and structural chemistry of Li[CrTi]O4 (Fd3̅m) in nonaqueous lithium cells. J. Electrochem. Soc. 2000, 147, 35923597,  DOI: 10.1149/1.1393944
  80. 80
    Capponi, J. J.; Billat, S.; Bordet, P.; Lambert-Andron, B.; Souletie, B. Structure, superconducting properties and stoichiometry of Li1-xTi2O4 spinel single crystals. Phys. C 1991, 185–189, 27212722,  DOI: 10.1016/0921-4534(91)91481-I
  81. 81
    de Picciotto, L. A.; Thackeray, M. M. Insertion/extraction reactions of lithium with LiV2O4. Mater. Res. Bull. 1985, 20, 14091420,  DOI: 10.1016/0025-5408(85)90158-8
  82. 82
    Barker, J.; Saidi, M. Y.; Swoyer, J. L. Electrochemical insertion properties of lithium vanadium titanate, LiVTiO4. Solid State Ionics 2004, 167, 413418,  DOI: 10.1016/j.ssi.2003.09.009

Cited By

Click to copy section linkSection link copied!

This article is cited by 10 publications.

  1. Louise Benincasa, Mathieu Duttine, Céline Goujon, Murielle Legendre, Matthew R. Suchomel, Marie Guignard. Exploring the Limits of Fe-Rich Chemistries in Na-Based CaFe2O4-Type Postspinel Oxides. Inorganic Chemistry 2024, 63 (22) , 10373-10385. https://doi.org/10.1021/acs.inorgchem.4c01278
  2. Justin C. Hancock, Patrick P. Ding, Yunyeong Choi, Gerbrand Ceder, John T. Vaughey, Kent J. Griffith, Kenneth R. Poeppelmeier. Electrochemistry in the Large Tunnels of Lithium Postspinel Compounds. Chemistry of Materials 2024, 36 (9) , 4616-4630. https://doi.org/10.1021/acs.chemmater.4c00389
  3. Ouail Zakary, Monique Body, Thibault Charpentier, Vincent Sarou-Kanian, Christophe Legein. Structural Modeling of O/F Correlated Disorder in TaOF3 and NbOF3–x(OH)x by Coupling Solid-State NMR and DFT Calculations. Inorganic Chemistry 2023, 62 (40) , 16627-16640. https://doi.org/10.1021/acs.inorgchem.3c02844
  4. Mazia Asghar, Hamid Ullah, Young-Han Shin, Essam A. Al-Ammar. Magneto-optic and photocatalytic response of CuCrX3 (X=Cl, Br, I) for energy storage applications. Computational Condensed Matter 2025, 42 , e00997. https://doi.org/10.1016/j.cocom.2024.e00997
  5. Masaki Akaogi, Takayuki Ishii, Kazunari Yamaura. Post-spinel-type AB2O4 high-pressure phases in geochemistry and materials science. Communications Chemistry 2024, 7 (1) https://doi.org/10.1038/s42004-024-01278-0
  6. Ghazanfar Nazir, Q Mahmood, Taharh Zelai, Adeela Rehman, Hind Albalawi, Nessrin A Kattan, S Bouzgarrou, Imed Boukhris. Control of spin on ferromagnetism and thermoelectric properties of Sr(V/Cr) 2 S 4 ; emergent spintronic aspirant. Physica Scripta 2024, 99 (5) , 055972. https://doi.org/10.1088/1402-4896/ad3c7d
  7. Amjad Salamah M. Aljaloud, A.F. Qasrawi, Latifah Hamad Khalid Alfhaid. Optical properties of chromium-selenide films designed for terahertz applications. Optik 2023, 292 , 171395. https://doi.org/10.1016/j.ijleo.2023.171395
  8. Syed Awais Rouf, Hind Albalawi, Taharh Zelai, Othman Hakami, Nessrin A. Kattan, Samah Al-Qaisi, S. Bouzgarrou, Muhammad Younas, Khaild I. Hussein, Q. Mahmood. Half-metallic ferromagnetism and thermoelectric effect in spinel chalcogenides SrX2S4 (X = Mn, Fe, Co) for spintronics and energy harvesting. Journal of Physics and Chemistry of Solids 2023, 182 , 111601. https://doi.org/10.1016/j.jpcs.2023.111601
  9. Feipeng Yang, Xuefei Feng, Zengqing Zhuo, Lauren Vallez, Yi-Sheng Liu, Scott A. McClary, Nathan T. Hahn, Per-Anders Glans, Kevin R. Zavadil, Jinghua Guo. Ca2+ Solvation and Electrochemical Solid/Electrolyte Interphase Formation Toward the Multivalent-Ion Batteries. Arabian Journal for Science and Engineering 2023, 48 (6) , 7243-7262. https://doi.org/10.1007/s13369-022-07597-5
  10. Q. Mahmood, Ghazanfar Nazir, Jameela Alzahrani, Nessrin A. Kattan, Samah Al-Qaisi, Hind Albalawi, Abeer Mera, Gaber A.M. Mersal, Mohamed M. Ibrahim, Mohammed A. Amin. Room temperature ferromagnetism and thermoelectric behavior of calcium based spinel chalcogenides CaZ2S4 (Z = Ti, V, Cr, Fe) for spintronic applications. Journal of Physics and Chemistry of Solids 2022, 167 , 110742. https://doi.org/10.1016/j.jpcs.2022.110742

ACS Organic & Inorganic Au

Cite this: ACS Org. Inorg. Au 2022, 2, 1, 8–22
Click to copy citationCitation copied!
https://doi.org/10.1021/acsorginorgau.1c00019
Published September 1, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

2617

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. CaFe2O4 structure. Gray spheres represent Na+. Green and blue octahedra represent the two symmetrically independent framework sites comprising the two “double rutile” chains.

    Figure 2

    Figure 2. Schematic summarizing the synthetic approach and results presented in this paper. The ion combinations in green were successfully synthesized in the postspinel structure, whereas the ion combinations in red formed alternative phases.

    Figure 3

    Figure 3. Rietveld refinement using synchrotron data for NaMg0.5Ti1.5O4. Blue crosses are the observed intensities, the green curve is the fitted pattern, the black curve is the difference pattern, and the red tick marks indicate the location of the CF-NaMg0.5Ti1.5O4 peaks.

    Figure 4

    Figure 4. Powder XRD patterns for the 1:1:1 and 0.99:0.99:1.01 Na/Cr/Ti samples compared to a calculated pattern for the CF phase. Asterisks highlight peaks from the secondary phase Na0.6Cr0.6Ti0.4O2 in the 1:1:1 Na/Cr/Ti sample.

    Figure 5

    Figure 5. A combinatorial representation of the Na-CF phase space. The phase-space shown covers all Na-CF compounds and combinations known to have been attempted in previous studies or this work. Conceivable but unexplored combinations (e.g., Sc3+ and Mo4+) are not depicted.

    Figure 6

    Figure 6. 23Na solid-state NMR spectra at 12.5 kHz MAS and 9.4 T: (a) Na1.16In1.18Sb0.66O4, (b) Na0.96In0.96Sn1.04O4, and (c) “NaSc1.5Sb0.5O4” central transition resonances.

    Figure 7

    Figure 7. Unit cell volume of the new Na-CFs versus the weighted-average effective ionic radius of the framework cations in octahedral coordination. (58)

    Figure 8

    Figure 8. (a) Computed stability of NaxA3+B4+O4 and NaxA2+0.5B4+1.5O4 CF phases with three degrees of sodiation (x = 0, 0.5, 1.0). (b) Energy above hull of each xNa = 1 compound after including ideal mixing entropy for the octahedral cations at 1200 K. Green and red boxes show successful/failed syntheses at ambient pressure. (c) Energy difference between the CF and spinel structure. Note that the empty boxes correspond to In-containing compounds that failed to converge in either the CF or spinel structures and were consequently excluded from this analysis.

  • References


    This article references 82 other publications.

    1. 1
      Bertraut, E.-F.; Blum, P.; Magnano, G. Structure des vanadite, chromite et ferrite monocalciques. Bull. Mineral. 1956, 79, 536561
    2. 2
      Hill, P. M.; Peiser, H. S.; Rait, J. R. The crystal structure of calcium ferrite and β calcium chromite. Acta Crystallogr. 1956, 9, 981986,  DOI: 10.1107/S0365110X56002862
    3. 3
      Irifune, T.; Fujino, K.; Ohtani, E. A new high-pressure form of MgAl2O4. Nature 1991, 349, 409411,  DOI: 10.1038/349409a0
    4. 4
      Funamori, N.; Jeanloz, R.; Nguyen, J. H.; Kavner, A.; Caldwell, W. A.; Fujino, K.; Miyajima, N.; Shinmei, T.; Tomioka, N. High-pressure transformations in MgAl2O4. J. Geophys. Res. Solid Earth 1998, 103, 2081320818,  DOI: 10.1029/98JB01575
    5. 5
      Liu, L. High pressure NaAlSiO4: the first silicate calcium ferrite isotype. Geophys. Res. Lett. 1977, 4, 183186,  DOI: 10.1029/GL004i005p00183
    6. 6
      Yamada, H.; Matsui, Y.; Ito, E. Crystal-chemical characterization of NaAlSiO4 with the CaFe2O4 structure. Mineral. Mag. 1983, 47, 177181,  DOI: 10.1180/minmag.1983.047.343.07
    7. 7
      Reid, A. F.; Wadsley, A. D.; Ringwood, A. E. High pressure NaAlGeO4, a calcium ferrite isotype and model structure of silicates at depth in the earth’s mantle. Acta Crystallogr. 1967, 23, 736739,  DOI: 10.1107/S0365110X67003627
    8. 8
      Reid, A. F.; Wadsley, A. D.; Sienko, M. J. Crystal chemistry of sodium scandium titanate, NaScTiO4, and its isomorphs. Inorg. Chem. 1968, 7, 112118,  DOI: 10.1021/ic50059a024
    9. 9
      Ishii, T.; Sakai, T.; Kojitani, H.; Mori, D.; Inaguma, Y.; Matsushita, Y.; Yamaura, K.; Akaogi, M. High-pressure phase relations and crystal structures of postpinel phases in MgV2O4, FeV2O4, and MnCr2O4: crystal chemistry of AB2O4 postspinel compounds. Inorg. Chem. 2018, 57, 66486657,  DOI: 10.1021/acs.inorgchem.8b00810
    10. 10
      Yamaura, K.; Huang, Q.; Zhang, L.; Takada, K.; Baba, Y.; Nagai, T.; Matsui, Y.; Kosuda, K.; Takayama-Muromachi, E. Spinel-to-CaFe2O4-type structural transformation in LiMn2O4 under higher pressure. J. Am. Chem. Soc. 2006, 128, 94489456,  DOI: 10.1021/ja0612302
    11. 11
      Yamaura, K.; Arai, M.; Sato, A.; Karki, A. B.; Young, D. P.; Movshovich, R.; Okamoto, S.; Mandrus, D.; Takayama-Muromachi NaV2O4: a quasi-1D metallic antiferromagnet with half-metallic chains. Phys. Rev. Lett. 2007, 99, 196601,  DOI: 10.1103/PhysRevLett.99.196601
    12. 12
      Sakurai, H.; Kolodiazhnyi, T.; Michiue, Y.; Takayama-Muromachi, E.; Tanabe, Y.; Kikuchi, H. Unconventional colossal magnetoresistance in sodium chromium oxide with a mixed-valence state. Angew. Chem. 2012, 124, 67576760,  DOI: 10.1002/ange.201201884
    13. 13
      Young, O.; Wildes, A. R.; Manuel, P.; Ouladdiaf, B.; Khalyavin, D. D.; Balakrishnan, G.; Petrenko, O. A. Highly frustrated magnetism in SrHo2O4: coexistence of two types of short-range order. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 024411,  DOI: 10.1103/PhysRevB.88.024411
    14. 14
      Dutton, S. E.; Broholm, C. L.; Cava, R. J. Divergent effects of static disorder and hole doping in geometrically frustrated β-CaCr2O4. J. Solid State Chem. 2010, 183, 17981804,  DOI: 10.1016/j.jssc.2010.05.032
    15. 15
      Arévalo-López, Á. M.; Dos santos-García, A. J.; Castillo-Martínez, E.; Durán, A.; Alario-Franco, M. Á. Spinel to CaFe2O4 transformation: Mechanism and properties of β-CdCr2O4. Inorg. Chem. 2010, 49, 28272833,  DOI: 10.1021/ic902228h
    16. 16
      Shimomura, Y.; Kurushima, T.; Kijima, N. Photoluminescence and crystal structure of green-emitting phosphor CaSc2O4:Ce3+. J. Electrochem. Soc. 2007, 154, J234,  DOI: 10.1149/1.2741172
    17. 17
      Hao, Z.; Zhang, J.; Zhang, X.; Lu, S.; Wang, X. Blue-green-emitting phosphor CaSc2O4:Tb3+: tunable luminescence manipulated by cross-relaxation. J. Electrochem. Soc. 2009, 156, H193,  DOI: 10.1149/1.3060382
    18. 18
      Hao, Z.; Zhang, J.; Zhang, X.; Wang, X. CaSc2O4:Eu3+: A tunable full-color emitting phosphor for white light emitting diodes. Opt. Mater. 2011, 33, 355358,  DOI: 10.1016/j.optmat.2010.09.035
    19. 19
      Bruno, S. R.; Blakely, C. K.; Clapham, J. B.; Davis, J. D.; Bi, W.; Alp, E. E.; Poltavets, V. V. Synthesis and electrochemical properties of novel LiFeTiO4 and Li2FeTiO4 polymorphs with the CaFe2O4-type structures. J. Power Sources 2015, 273, 396403,  DOI: 10.1016/j.jpowsour.2014.09.115
    20. 20
      Jung, Y. H.; Kim, D. K.; Hong, S.-T. Synthesis, structure, and electrochemical Li-ion intercalation of LiRu2O4 with CaFe2O4-type structure. J. Power Sources 2013, 233, 285289,  DOI: 10.1016/j.jpowsour.2013.01.119
    21. 21
      Sun, X.; Blanc, L.; Nolis, G. M.; Bonnick, P.; Cabana, J.; Nazar, L. F. NaV1.25Ti0.75O4: a potential post-spinel cathode material. Chem. Mater. 2018, 30, 121128,  DOI: 10.1021/acs.chemmater.7b03383
    22. 22
      Mukai, K.; Uyama, T.; Yamada, I. Structural and electrochemical analyses on the transformation of CaFe2O4-type LiMn2O4 from spinel-type LiMn2O4. ACS Omega 2019, 4, 64596467,  DOI: 10.1021/acsomega.9b00588
    23. 23
      Nolis, G.; Gallardo-Amores, J. M.; Serrano-Sevillano, J.; Jahrman, E.; Yoo, H. D.; Hu, L.; Hancock, J. C.; Bolotnikov, J.; Kim, S.; Freeland, J. W.; Liu, Y.-S.; Poeppelmeier, K. R.; Seidler, G. T.; Guo, J.; Alario-Franco, M. A.; Casas-Cabanas, M.; Morán, E.; Cabana, J. Factors defining the intercalation electrochemistry of CaFe2O4-type manganese oxides. Chem. Mater. 2020, 32, 82038215,  DOI: 10.1021/acs.chemmater.0c01858
    24. 24
      Liu, X.; Wang, X.; Iyo, A.; Yu, H.; Li, D.; Zhou, H. High stable post-spinel NaMn2O4 cathode of sodium ion battery. J. Mater. Chem. A 2014, 2, 1482214826,  DOI: 10.1039/C4TA03349C
    25. 25
      Chiring, A.; Senguttuvan, P. Chemical pressure-stabilized post spinel-NaMnSnO4 as potential cathode for sodium-ion batteries. Bull. Mater. Sci. 2020, 43, 306,  DOI: 10.1007/s12034-020-02203-6
    26. 26
      Ling, C.; Mizuno, F. Phase stability of post-spinel compound AMn2O4 (A = Li, Na, or Mg) and its application as a rechargeable battery cathode. Chem. Mater. 2013, 25, 30623071,  DOI: 10.1021/cm401250c
    27. 27
      Hannah, D. C.; Sai Gautam, G.; Canepa, P.; Rong, Z.; Ceder, G. Magnesium ion mobility in post-spinels accessible at ambient pressure. Chem. Commun. 2017, 53, 51715174,  DOI: 10.1039/C7CC01092C
    28. 28
      Dompablo, M. E. A.; Krich, C.; Nava-Avendaño, J.; Biškup, N.; Palacín, M. R.; Bardé, F. A joint computational and experimental evaluation of CaMn2O4 polymorphs as cathode materials for Ca ion batteries. Chem. Mater. 2016, 28, 68866893,  DOI: 10.1021/acs.chemmater.6b02146
    29. 29
      Müller-Buschbaum, H. The crystal chemistry of AM2O4 oxometallates. J. Alloys Compd. 2003, 349, 49104,  DOI: 10.1016/S0925-8388(02)00925-8
    30. 30
      Akimoto, J.; Takei, H. Synthesis and crystal structure of NaTi2O4: a new mixed-valence sodium titanate. J. Solid State Chem. 1989, 79, 212217,  DOI: 10.1016/0022-4596(89)90268-5
    31. 31
      Darriet, J.; Vidal, A. Les composés NaRu2O4 et NaFeRuO4. Structure cristalline de NaFeRuO4. Bull. Soc. Fr. Minéral Cristallogr. 1975, 98, 374377
    32. 32
      Viciu, L.; Ryser, A.; Mensing, C.; Bos, J.-W. G. Ambient-pressure synthesis of two new vanadium-based calcium ferrite-type compounds: NaV1.25Ti0.75O4 and NaVSnO4. Inorg. Chem. 2015, 54, 72647271,  DOI: 10.1021/acs.inorgchem.5b00655
    33. 33
      Akimoto, J.; Awaka, J.; Kijima, N.; Takahashi, Y.; Maruta, Y.; Tokiwa, K.; Watanabe, T. High-pressure synthesis and crystal structure analysis of NaMn2O4 with the calcium ferrite-type structure. J. Solid State Chem. 2006, 179, 169174,  DOI: 10.1016/j.jssc.2005.10.020
    34. 34
      Yamaura, K.; Huang, Q.; Moldovan, M.; Young, D. P.; Sato, A.; Baba, Y.; Nagai, T.; Matsui, Y.; Takayama-Muromachi, E. High-pressure synthesis, crystal structure determination, and a Ca substitutional study of the metallic rhodium oxide NaRh2O4. Chem. Mater. 2005, 17, 359365,  DOI: 10.1021/cm0483846
    35. 35
      Shukaev, I. L.; Volochaev, V. A. Ternary sodium and titanium oxides with cobalt(II). Russ. J. Inorg. Chem. 1995, 12, 19741980
    36. 36
      Nalbandyan, V. B.; Shukaev, I. L. Triple oxides of sodium, nickel, and titanium. Russ. J. Inorg. Chem. 1992, 37, 12311235
    37. 37
      Archaimbault, F.; Choisnet, J.; Rautureau, M. New ferriantimonates with the CaFe2O4 type structure: Na2Fe3SbO8 and isomorphous series Na2Fe2+xSn2–2xSbxO8 (0 ≤ x ≤ 1). Eur. J. Solid State Inorg. Chem. 1988, 25, 573587
    38. 38
      Ishiguro, T.; Tanaka, K.; Marumo, F.; Ismail, M. G. M. U.; Hirano, S.; Somiya, S. Non-stoichiometric sodium iron (II) titanium (IV) oxide. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1978, 34, 33463348,  DOI: 10.1107/S0567740878010870
    39. 39
      Feger, C. R.; Kolis, J. W. Na3Mn4Te2O12. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, 54, 10551057,  DOI: 10.1107/S0108270197020076
    40. 40
      Toby, B. H.; Von Dreele, R. B. GSAS-II: the genesis of a modern open-source all purpose crystallography software package. J. Appl. Crystallogr. 2013, 46, 544549,  DOI: 10.1107/S0021889813003531
    41. 41
      Amoureux, J.-P.; Fernandez, C.; Steuernagel, S. Z Filtering in MQMAS NMR. J. Magn. Reson., Ser. A 1996, 123, 116118,  DOI: 10.1006/jmra.1996.0221
    42. 42
      Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 1550,  DOI: 10.1016/0927-0256(96)00008-0
    43. 43
      Kresse, G.; Hafner, J. Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements. J. Phys.: Condens. Matter 1994, 6, 8245,  DOI: 10.1088/0953-8984/6/40/015
    44. 44
      Blöchl, P. E.; Jepsen, O.; Andersen, O. K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 16223,  DOI: 10.1103/PhysRevB.49.16223
    45. 45
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    46. 46
      Wang, L.; Maxisch, T.; Ceder, G. Oxidation energies of transition metal oxides within the GGA+U framework. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73, 195107,  DOI: 10.1103/PhysRevB.73.195107
    47. 47
      Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; Persson, K. A. Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002,  DOI: 10.1063/1.4812323
    48. 48
      Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.; Persson, K. A.; Ceder, G. Formation enthalpies by mixing GGA and GGA + U calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 045115,  DOI: 10.1103/PhysRevB.84.045115
    49. 49
      Ong, S. P.; Richards, W. D.; Jain, A.; Hautier, G.; Kocher, M.; Cholia, S.; Gunter, D.; Chevrier, V. L.; Persson, K. A.; Ceder, G. Python Materials Genomics (pymatgen): A robust, open-source python library for materials analysis. Comput. Mater. Sci. 2013, 68, 314319,  DOI: 10.1016/j.commatsci.2012.10.028
    50. 50
      Avdeev, M. Yu; Nalbandyan, V. B.; Medvedev, B. S. Hexagonal sodium titanate chromite: a new high-conductivity solid electrolyte. Inorg. Mater. 1997, 33, 500503
    51. 51
      Mumme, W. G.; Reid, A. F. Non-stoichiometric sodium iron titanate, NaxFexTi2-xO4, 0.90> x> 0.75. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1968, 24, 625631,  DOI: 10.1107/S0567740868002955
    52. 52
      Nalbandyan, V. B.; Rykalova, S. I.; Bikyashev, E. A.; Isakova, S. Yu Sodium titanium magnesium (zinc) ternary oxides. Zh. Neorg. Khim. 1989, 34, 23812386
    53. 53
      Shukaev, I. L.; Butova, V. V.; Chernenko, S. V.; Pospelov, A. A.; Shapovalov, V. V.; Guda, A. A.; Aboraia, A. M.; Zahran, H. Y.; Yahia, I. S.; Soldatov, A. V. New orthorhombic sodium iron(+2) titanate. Ceram. Int. 2020, 46, 44164422,  DOI: 10.1016/j.ceramint.2019.10.167
    54. 54
      Thackeray, M. M.; de Kock, A.; Rossouw, M. H.; Liles, D.; Bittihn, R.; Hoge, D. Spinel electrodes from the Li-Mn-O system for rechargeable lithium battery applications. J. Electrochem. Soc. 1992, 139, 363366,  DOI: 10.1149/1.2069222
    55. 55
      Deschanvres, A.; Raveau, B.; Sekkal, Z. Mise en evidence et etude cristallographique d’une nouvelle solution solide de type spinelle Li1+xTi2–xO4 0 ⩽ x ⩽ 0, 333. Mater. Res. Bull. 1971, 6, 699704,  DOI: 10.1016/0025-5408(71)90103-6
    56. 56
      Rickert, K.; Sedefoglu, N.; Malo, S.; Caignaert, V.; Kavak, H.; Poeppelmeier, K. R. Structural, electrical, and optical properties of the tetragonal, fluorite-related Zn0.456In1.084Ge0.460O3. Chem. Mater. 2015, 27, 50725079,  DOI: 10.1021/acs.chemmater.5b01724
    57. 57
      Marinkovic, B. A.; Mancic, L.; Jardim, P. M.; Milosevic, O.; Rizzo, F. Hydrothermal synthesis of NaxFexTi2-xO4 from natural ilmenite sand: a CaFe2O4 structure type compound. Solid State Commun. 2008, 145, 346350,  DOI: 10.1016/j.ssc.2007.12.006
    58. 58
      Shannon, R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 751767,  DOI: 10.1107/S0567739476001551
    59. 59
      Mumme, W. G. The structure of Na4Mn4Ti5O18. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1968, 24, 11141120,  DOI: 10.1107/S0567740868003778
    60. 60
      Parant, J.-P.; Olazcuaga, R.; Devalette, M.; Fouassier, C.; Hagenmuller, P. Sur quelques nouvelles phases de formule NaxMnO2 (x ≤ 1). J. Solid State Chem. 1971, 3, 111,  DOI: 10.1016/0022-4596(71)90001-6
    61. 61
      Toda, K.; Kameo, Y.; Kurita, S.; Sato, M. Crystal structure determination and ionic conductivity of layered perovskite compounds NaLnTiO4 (Ln = rare earth). J. Alloys Compd. 1996, 234, 1925,  DOI: 10.1016/0925-8388(95)01969-3
    62. 62
      Bruhn, G.; Beutel, S.; Pfaff, G.; Albert, B. Low-temperature synthesis of freudenbergite-type titanate bronzes from metal halides, crystal growth from molybdate flux, and crystal structure determination of Na1.84Zn0.92Ti7.08O16. J. Alloys Compd. 2015, 644, 783787,  DOI: 10.1016/j.jallcom.2015.05.076
    63. 63
      Kunz, M.; Brown, I. D. Out-of-center distortions around octahedrally coordinated d0 transition metals. J. Solid State Chem. 1995, 115, 395406,  DOI: 10.1006/jssc.1995.1150
    64. 64
      Ok, K. M.; Halasyamani, P. S.; Casanova, D.; Llunell, M.; Alemany, P.; Alvarez, S. Distortions in octahedrally coordinated d0 transition metal oxides: a continuous symmetry measures approach. Chem. Mater. 2006, 18, 31763183,  DOI: 10.1021/cm0604817
    65. 65
      Urban, A.; Abdellahi, A.; Dacek, S.; Artrith, N.; Ceder, G. Electronic-structure origin of cation disorder in transition-metal oxides. Phys. Rev. Lett. 2017, 119, 176402,  DOI: 10.1103/PhysRevLett.119.176402
    66. 66
      Shilov, G. V.; Atovmyan, L. O.; Volochaev, V. A.; Nalbandyan, V. B. Crystal structure and ionic conductivity of a new sodium magnesium titanium oxide. Kristallografiya 1999, 44, 10291033
    67. 67
      Ishiguro, T.; Tanaka, K.; Marumo, F.; Ismail, M. G. M. U.; Hirano, S.; Somiya, S. Freudenbergite. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1978, 34, 255256,  DOI: 10.1107/S0567740878002708
    68. 68
      Pospelov, A. A.; Nalbandyan, V. B. Preparation, crystal structures and rapid hydration of P2- and P3-type sodium chromium antimony oxides. J. Solid State Chem. 2011, 184, 10431047,  DOI: 10.1016/j.jssc.2011.03.011
    69. 69
      Politaev, V. V.; Nalbandyan, V. B. Subsolidus phase relations, crystal chemistry and cation-transport properties of sodium iron antimony oxides. Solid State Sci. 2009, 11, 144150,  DOI: 10.1016/j.solidstatesciences.2008.04.014
    70. 70
      Arillo, M. A.; Lopez, M. L.; Perez-Cappe, E.; Pico, C.; Viega, M. L. Crystal structure and electrical properties of LiFeTiO4 spinel. Solid State Ionics 1998, 107, 307312,  DOI: 10.1016/S0167-2738(97)00537-7
    71. 71
      Kang, K.; Meng, Y. S.; Bréger, J.; Grey, C. P.; Ceder, G. Electrodes with high power and high capacity for rechargeable lithium batteries. Science 2006, 311, 977980,  DOI: 10.1126/science.1122152
    72. 72
      Sai Gautam, G.; Canepa, P.; Urban, A.; Bo, S.-H.; Ceder, G. Influence of inversion on Mg mobility and electrochemistry in spinels. Chem. Mater. 2017, 29, 79187930,  DOI: 10.1021/acs.chemmater.7b02820
    73. 73
      Bayliss, R. D.; Key, B.; Sai Gautam, G.; Canepa, P.; Kwon, B. J.; Lapidus, S. H.; Dogan, F.; Adil, A. A.; Lipton, A. S.; Baker, P. J.; Ceder, G.; Vaughey, J. T.; Cabana, J. Probing Mg migration in spinel oxides. Chem. Mater. 2020, 32, 663670,  DOI: 10.1021/acs.chemmater.9b02450
    74. 74
      Kwon, B. J.; Yin, L.; Park, H.; Parajuli, R.; Kumar, K.; Kim, S.; Yang, M.; Murphy, M.; Zapol, P.; Liao, C.; Fister, T. T.; Klie, R. F.; Cabana, J.; Vaughey, J. T.; Lapidus, S. H.; Key, B. High voltage Mg-ion battery cathode via a solid solution Cr-Mn spinel oxide. Chem. Mater. 2020, 32, 65776587,  DOI: 10.1021/acs.chemmater.0c01988
    75. 75
      Robertson, A. D.; Tukamoto, H.; Irvine, J. T. S. Li1+xFe1–3xTi1+2xO4 (0.0 ≤ x ≤ 0.33) based spinels: possible negative electrode materials for future Li-ion batteries. J. Electrochem. Soc. 1999, 146, 39583962,  DOI: 10.1149/1.1392576
    76. 76
      Miura, A.; Bartel, C. J.; Goto, Y.; Mizuguchi, Y.; Moriyoshi, C.; Kuroiwa, Y.; Wang, Y.; Yaguchi, T.; Shirai, M.; Nagao, M.; Rosero-Navarro, N. C.; Tadanaga, K.; Ceder, G.; Sun, W. Observing and modeling the sequential pairwise reactions that drive solid-state ceramic synthesis. Adv. Mater. 2021, 33, 2100312,  DOI: 10.1002/adma.202100312
    77. 77
      Sun, W.; Dacek, S. T.; Ong, S. P.; Hautier, G.; Jain, A.; Richards, W. D.; Gamst, A. C.; Persson, K. A.; Ceder, G. The thermodynamic scale of inorganic crystalline metastability. Sci. Adv. 2016, 2, e1600225,  DOI: 10.1126/sciadv.1600225
    78. 78
      Colbow, K.M.; Dahn, J.R.; Haering, R.R. Structure and electrochemistry of the spinel oxides LiTi2O4 and Li4/3Ti5/3O4. J. Power Sources 1989, 26, 397402,  DOI: 10.1016/0378-7753(89)80152-1
    79. 79
      Ohzuku, T.; Tatsumi, K.; Matoba, N.; Sawai, K. Electrochemistry and structural chemistry of Li[CrTi]O4 (Fd3̅m) in nonaqueous lithium cells. J. Electrochem. Soc. 2000, 147, 35923597,  DOI: 10.1149/1.1393944
    80. 80
      Capponi, J. J.; Billat, S.; Bordet, P.; Lambert-Andron, B.; Souletie, B. Structure, superconducting properties and stoichiometry of Li1-xTi2O4 spinel single crystals. Phys. C 1991, 185–189, 27212722,  DOI: 10.1016/0921-4534(91)91481-I
    81. 81
      de Picciotto, L. A.; Thackeray, M. M. Insertion/extraction reactions of lithium with LiV2O4. Mater. Res. Bull. 1985, 20, 14091420,  DOI: 10.1016/0025-5408(85)90158-8
    82. 82
      Barker, J.; Saidi, M. Y.; Swoyer, J. L. Electrochemical insertion properties of lithium vanadium titanate, LiVTiO4. Solid State Ionics 2004, 167, 413418,  DOI: 10.1016/j.ssi.2003.09.009
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsorginorgau.1c00019.

    • Table of attempted syntheses and products, Rietveld refinements, additional 23Na NMR spectra, and table of Na–O bond distances for selected compounds (PDF)

    Accession Codes

    CCDC 2103718, 21038252103831, 21038332103835, 2103925, 21039282103929, 2103933, 2103935, and 2103938 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.