Abstract
This study demonstrates that the influence of cationic composition on the phase behavior of vanadates under high pressure must be meticulously considered. In this investigation, we report an in situ high-pressure powder X-ray diffraction investigation on triclinic Fe0.9Al0.1VO4 (space group P1̅) up to 11 GPa. The structural sequence of Fe0.9Al0.1VO4 is different than that of FeVO4. Our analysis shows that Fe0.9Al0.1VO4 undergoes a first-order structural phase transition at 2.85 GPa to another triclinic structure described by the same space group with a volume collapse of ∼9%. At 6.1 GPa, we observed the onset of a second phase transition to a monoclinic structure (space group P2/c), with coexistence of both phases until 8.55 GPa. The transformation to the second phase is completed at 9.15 GPa, with a volume collapse of ∼13%. On release of pressure to ambient conditions, we have observed the coexistence of the second and first high-pressure phases. The compressibility of the three phases of the compound has been studied too. We have observed variations in structural sequence and compressibility behavior due to Al incorporation. Since electronic properties could be modified by tuning the crystal structure, the present results could have an impact on applications of the studied compound such as photocatalysis and batteries.


1. Introduction
Among the orthovanadates, FeVO4, due to its unique electronic, optical, and catalytic properties, finds many applications in various technologically important fields. , Iron orthovanadate is considered a promising photocatalyst in environmental applications, particularly for wastewater treatment, due to its ability to efficiently degrade organic pollutants under visible light. , FeVO4 is also a potential candidate for energy storage applications as a promising electrode material for supercapacitors, lithium-ion batteries, and sodium-ion batteries due to its excellent electrochemical properties, including high specific capacitance, good cycling stability, and redox activity, making it a potential candidate for high-power energy storage devices such as wearable electronics and hybrid electric vehicles. FeVO4 is also employed in gas-sensing applications, particularly for detecting gases such us ammonia (NH3), nitrogen dioxide (NO2), and carbon monoxide (CO), owing to its high sensitivity and selectivity. Due to its semiconducting properties and light absorption capabilities, it finds applications in solar cells and optoelectronic devices. FeVO4 doped with Cr shows an antiferromagnetic magnetization, making it relevant in spintronics and advanced electronic applications. It has also been observed that considerable enhancement in the spintronics, photocatalytic, , and electrochemical properties of FeVO4 could be achieved by means of substitution of Fe with various trivalent transition and post-transition metals.
At ambient conditions, FeVO4 crystallizes in a triclinic structure known as FeVO4-I (space group (SG) P1̅). , The crystal structure is represented in Figure a. The unit cell has six formula units (Z = 6) and comprises 36 atoms, all of them occupying 2i Wyckoff positions, forming a complex network of chain-like motifs. In FeVO4, there are three distinct crystallographic sites for the Fe3+ ion. In two nonequivalent positions, it forms distorted FeO6 octahedra, whereas in the third position, it forms a distorted FeO5 trigonal bipyramid. These Fe–O polyhedral units form C-shaped chainlike networks running parallel to the crystallographic c-axis, with these chains interconnected by VO4 tetrahedra. Iron vanadate has multiple polymorphs, which differ in crystal structure and stability depending on temperature, pressure, and synthesis conditions. The well-known metastable polymorphs of FeVO4 include orthorhombic FeVO4-II (SG: Cmcm, 2 GPa −800 °C), orthorhombic FeVO4-III (SG: Pbcn, 6 GPa −750 °C), and monoclinic FeVO4-IV (SG: P2/c, 8 GPa −800 °C). These polymorphs have been experimentally recovered under ambient conditions. The monoclinic FeVO4-IV polymorph also has been reported to be synthesized at a much lower pressure of 5–5.5 GPa and 800 °C. In the past, numerous studies on triclinic FeVO4 have been carried out, which elucidated the structural phase transitions in this material under high-pressure conditions up to 12.3 GPa. − It is worth mentioning a recent high-pressure single-crystal X-ray diffraction (XRD) and density functional theory (DFT) study in FeVO4 which has presented pioneering findings with substantial evidence of structural phase transitions and provided valuable insights into the systematics of structural phase transitions in this material. According to this study, the FeVO4-I triclinic phase transforms at 2.1 GPa to another triclinic structure FeVO4-I′ (SG: P1̅) which was detected uniquely in this investigation. Additionally, a further transition to previously unknown monoclinic structure, FeVO4-II′ (SG: C2/m), was found at 4.8 GPa.
1.
Crystal structure of phases I, I′, and IV of FeVO4 (Fe0.9Al0.1VO4). V (Fe/Al) coordination polyhedra are shown in blue (brown). Oxygen atoms are shown in red.
As discussed earlier, partially substituting Fe by a trivalent transition and post-transition elements enhances the properties of FeVO4. Till date, there is no high-pressure investigation reporting on partially substituted FeVO4 compounds. It is known that atomic substitution (chemical pressure) could affect the high-pressure behavior and properties of oxides. An investigation combining chemical and hydrostatic pressure contributes to deepen the understanding of the influence of pressure in vanadates, and it is useful for developing routes for synthesizing and tuning novel materials. − Thus, in this investigation, we have carried out high-pressure synchrotron powder XRD measurements on Al-substituted FeVO4 up to 11 GPa to understand the nature of structural phase transitions in this material. Our findings could be valuable to tune the electronic, magnetic, and vibrational properties of this compound for practical applications.
2. Experimental Details
Single crystals of Fe0.9Al0.1VO4, up to 0.5 mm in size, were obtained by the flux method, using V2O5 as the self-flux. High-purity reagents Fe2O3, Al2O3, and V2O5 in powder form (Sigma-Aldrich) were mixed in a 0.9:0.1:2 molar ratio in a platinum crucible, which was sealed using a platinum cap. The closed crucible was heated in a furnace at 750 °C for 24 h, after which the furnace was cooled to 640 °C at 1.2 °C/h. Afterward, the crucible was removed from the furnace, allowed to cool at room temperature, and then uncapped. The crystals embedded in the flux were then separated from them by submerging and repeatedly flushing the whole crucible with hot nitric acid (1.5 M). Carefully selected crystals without evident flux inclusions were then used for our experiments. A finely ground powder was obtained from the crystals to perform high-pressure (HP) studies.
A membrane-type diamond-anvil cell (DAC), with diamond culets 400 μm in diameter, was used to generate the HP conditions. A stainless-steel gasket was first preindented to a 40 μm thickness, and after that, a 180 μm-diameter hole was drilled, in the center of the indentation, to serve as the sample chamber. The powdered sample was loaded together with a grain of copper (Cu) which was used to determine the pressure by means of the calibration provided by Dewaele et al. Pressure was determined with an accuracy better than 0.05 GPa. A 16:3:1 methanol–ethanol–water mixture was used as the pressure-transmitting medium (PTM). This PTM remains quasi-hydrostatic up to the highest pressure covered by this study. Room-temperature HP-XRD measurements were performed at the Xpress beamline of the Elettra synchrotron using a monochromatic wavelength of 0.4956 Å and a PILATUS 3S 6M detector. The powder XRD measurements were carried out in an angular dispersion configuration. The instrument was calibrated using cerium dioxide as a standard. The two-dimensional diffraction rings obtained from the detector were integrated using Dioptas to obtain the conventional intensity versus 2θ one-dimensional diffractograms. The structural analysis was performed by employing the Rietveld technique using GSAS. The background was adjusted using a Chebyshev polynomial function of the first kind, comprising eight coefficients, while the peak profiles were represented through a pseudo-Voigt function. In the refinements, we assumed that Al and Fe are distributed uniformly in all of the Fe positions of the structures of undoped FeVO4.
3. Result and Discussion
The XRD pattern of Al-substituted FeVO4 (Fe0.9Al0.1VO4) at 0.1 GPa recorded in the DAC together with Rietveld refinement is shown in Figure . The low-pressure phase can be assigned to a structure isomorphic to the FeVO4–I type-triclinic structure of pristine FeVO4 (space group P1̅, Z = 6). The unit-cell parameters at 0.1 GPa are found to be a = 6.698(5) Å, b = 8.020(7) Å, c = 9.328(5) Å, α = 96.55(1)°, β = 106.71(3)°, and γ = 101.48(3)°. The complete structural information, including atomic positions, is reported in Table . The goodness-of-fit parameters obtained were R wp = 1.6% and R p = 1.0%. The unit-cell parameters determined are slightly different than the values previously for the same structure in undoped FeVO4 ,, and by recent single-crystal measurements reported by Gonzalez-Platas et al. (CCDC 1987953). The unit-cell parameters determined in the previous study for FeVO4-I in the undoped sample are a = 6.7137(5) Å, b = 8.0609(5) Å, c = 9.3530(6) Å, α = 96.671(5)°, β = 106.645(6)°, and γ = 101.524(5)°. Thus, the substitution of 10% of Fe by Al induces a change in unit-cell parameters smaller than 0.5%, and a change smaller than 1% in the unit-cell volume.
2.

Rietveld refinement of the XRD pattern of triclinic Fe0.9Al0.1VO4 at ambient conditions (λ = 0.4956 Å). Data are shown as black crosses (×), while the red solid line represents the result of the refinement. The blue (green) line represents the residuals of the fitting (the background). Vertical pink (black) bars identify the Bragg reflections of Fe0.9Al0.1VO4 (copper).
1. Crystal Structure of the Triclinic Phase I of Fe0.9Al0.1VO4 at 0.1 GPa and Room Temperature.
| a = 6.698(5) Å, b = 8.020(7) Å, c = 9.328(5) Å, α = 96.55(1)°, β = 106.71(3)°, and γ = 101.48(3)°, V = 462.4(4) Å3, Z = 6 | |||||
|---|---|---|---|---|---|
| atom | site | x | y | z | Uiso (Å2) |
| Fe1/Al1 | 2i | 0.7566(8) | 0.6914(8) | 0.9077(8) | 0.0568 |
| Fe2/Al2 | 2i | 0.9793(8) | 0.2996(8) | 0.5144(8) | 0.0756 |
| Fe3/Al3 | 2i | 0.4690(8) | 0.8969(8) | 0.7110(8) | 0.0296 |
| V1 | 2i | 0.9973(8) | 0.9970(8) | 0.7560(8) | 0.1040 |
| V2 | 2i | 0.2087(8) | 0.6010(8) | 0.8438(8) | 0.0567 |
| V3 | 2i | 0.5266(8) | 0.3036(8) | 0.6351(8) | 0.0551 |
| O1 | 2i | 0.064(3) | 0.526(3) | 0.673(3) | 0.0594 |
| O2 | 2i | 0.771(3) | 0.879(3) | 0.762(3) | 0.1712 |
| O3 | 2i | 0.132(3) | 0.903(3) | 0.675(3) | 0.0347 |
| O4 | 2i | 0.460(3) | 0.753(3) | 0.890(3) | 0.0207 |
| O5 | 2i | 0.624(3) | 0.300(3) | 0.459(3) | 0.0983 |
| O6 | 2i | 0.933(3) | 0.146(3) | 0.665(3) | 0.0184 |
| O7 | 2i | 0.257(3) | 0.434(3) | 0.926(3) | 0.0245 |
| O8 | 2i | 0.176(3) | 0.089(3) | 0.932(3) | 0.0900 |
| O9 | 2i | 0.072(3) | 0.738(3) | 0.964(3) | 0.0168 |
| O10 | 2i | 0.247(3) | 0.296(3) | 0.545(3) | 0.0792 |
| O11 | 2i | 0.634(3) | 0.463(3) | 0.749(3) | 0.0235 |
| O12 | 2i | 0.535(3) | 0.150(3) | 0.749(3) | 0.0542 |
Figure shows the diffraction profiles at the selected pressures. There are no clearly visible changes in the diffraction patterns up to 2.35 GPa, and all the diffraction peaks could be indexed to the low-pressure triclinic phase, isomorphic to FeVO4-I. The diffraction peaks due to the pressure standard, i.e., copper, are marked as “Cu” in the diffraction profiles measured at 0.1, 6.1, and 0.1 MPa. The peaks from Cu are sharper than those from the sample and can be easily followed with pressure changes. A systematic shift was observed in all diffraction peaks due to lattice compression up to 2.35 GPa. However, at 2.85 GPa, we observed evident changes in the diffraction profile, indicative of the onset of a structural phase transition in Fe0.9Al0.1VO4.
3.
X-ray powder diffraction patterns of Fe0.9Al0.1VO4 at selected pressures (λ = 0.4956 Å). The patterns measured under pressure release are identified by (rel). Cu indicates the diffraction peaks of copper at 0.1 GPa, 6.1 GPa, and 0.1 MPa. At 6.1 GPa, (+) symbols identify diffraction peaks assigned to the triclinic HP phase, while (*) symbols identify emerging diffraction peaks assigned to the monoclinic HP phase.
As can be seen from Figure , the diffraction profile is completely transformed when the pressure is increased from 2.35 to 2.85 GPa with the appearance of several additional diffraction peaks. It is interesting to highlight that the replacement of Fe with Al (Fe0.9Al0.1VO4) leads to two opposite effects. On the one hand, it produces a decrease in the zero-pressure volume when compared with the pure FeVO4 crystal (1% decrease), equivalent to the effect of applying a hydrostatic pressure of 0.75 GPa. On the other hand, the transition pressures observed in the solid solution are observed at higher pressure than in FeVO4. The first is related to the chemical pressure induced by replacing Fe with Al. The second one is apparently contradictory. However, the increase in the pressure stability range due to the presence of impurities is not unexpected. Previous studies have shown that even minimal impurity concentrations, as low as 1 mol %, can enhance the stability range of various compounds. This seems to be related to the local disorder caused by impurities, which favor the resistance of materials to external stresses including pressure.
The Rietveld refinement analysis presented in Figure supports that the high-pressure phase has a triclinic structure isomorphic to FeVO4–I′ (space group P1̅, Z = 6), which in the following we will name simply as phase I’. The unit-cell parameters at 2.85 GPa are found to be a = 6.257(5) Å, b = 7.485(7) Å, c = 9.232(9) Å α = 98.87(7)°, β = 104.09(2)°, and γ = 95.60(7)°. This crystal structure is represented in Figure b. The complete structural information, including atomic positions, is tabulated in Table . The goodness-of-fit parameters obtained were R wp = 2.0% and R p = 1.2%.
4.

Rietveld refinement of the XRD pattern of the HP triclinic Fe0.9Al0.1VO4 (phase I′) at 2.85 GPa (λ = 0.4956 Å). Data are shown as black crosses (×), while the red solid line represents the result of the refinement. The blue (green) line represents the residuals of the fitting (the background). Vertical pink (black) bars identify the Bragg reflections of Fe0.9Al0.1VO4 (copper).
2. Crystal Structure of the Triclinic Phase I′ of Fe0.9Al0.1VO4 at 2.85 GPa and Room Temperature.
| a = 6.257(5) Å, b = 7.485(7) Å, c = 9.232(9) Å α = 98.87(7)°, β = 104.09(2)°, and γ = 95.60(7)°. V = 410.2(4) Å3, Z = 6 | |||||
|---|---|---|---|---|---|
| atom | site | x | y | z | Uiso (Å2) |
| Fe1/Al1 | 2i | 0.262(1) | 0.360(1) | 0.662(1) | 0.0221 |
| Fe2/Al2 | 2i | 0.568(1) | 0.044(1) | 0.755(1) | 0.0332 |
| Fe3/Al3 | 2i | 0.903(1) | 0.380(1) | 0.092(1) | 0.0544 |
| V1 | 2i | 0.935(1) | 0.042(1) | 0.323(1) | 0.0231 |
| V2 | 2i | 0.596(1) | 0.707(1) | 0.986(1) | 0.0196 |
| V3 | 2i | 0.765(1) | 0.305(1) | 0.558(1) | 0.0162 |
| O1 | 2i | 0.899(5) | 0.640(5) | 0.912(5) | 0.0750 |
| O2 | 2i | 0.576(5) | 0.282(5) | 0.642(5) | 0.0235 |
| O3 | 2i | 0.161(5) | 0.953(5) | 0.305(5) | 0.0052 |
| O4 | 2i | 0.939(5) | 0.392(5) | 0.694(5) | 0.0094 |
| O5 | 2i | 0.401(5) | 0.531(5) | 0.865(5) | 0.0688 |
| O6 | 2i | 0.704(5) | 0.839(5) | 0.198(5) | 0.0434 |
| O7 | 2i | 0.379(5) | 0.932(5) | 0.623(5) | 0.0029 |
| O8 | 2i | 0.697(5) | 0.250(5) | 0.956(5) | 0.0157 |
| O9 | 2i | 0.714(5) | 0.443(5) | 0.462(5) | 0.0177 |
| O10 | 2i | 0.927(5) | 0.193(5) | 0.210(5) | 0.0284 |
| O11 | 2i | 0.046(5) | 0.190(5) | 0.520(5) | 0.0201 |
| O12 | 2i | 0.605(5) | 0.870(5) | 0.886(5) | 0.0385 |
The phase transition detected at 2.85 GPa is consistent with the earlier reported single-crystal XRD investigation in FeVO4. However, the transition pressure was reported as 2.1 GPa in FeVO4, which is 0.75 GPa lower as compared with our measurements. Thus, the partial substitution of Fe by Al extends the high-pressure stability of the low-pressure phase. It is worth noting that the observed structural phase transition involves changes in coordination of Fe/Al and V cations, and both the cations became octahedrally coordinated after the phase transition. This means that all the distorted FeO5/AlO5 trigonal bipyramids are converted to regular distorted octahedra. On the other hand, VO4 tetrahedra are also converted to distorted octahedra. The HP structure can be described by zigzag chains of edge-sharing FeO6 octahedra, which are connected via edge- and corner-sharing VO6 octahedra; see Figure c.
The change in coordination of Fe and V cations can affect the crystal field splitting, which in turn can modify the electronic energy levels of Fe and V. A higher coordination number can lower the crystal field stabilization energy, influencing bandgap tuning in semiconductor applications. , Recent high-pressure studies on optical properties of FeVO4 and InVO4 attribute lowering of bandgap to change in coordination number due to alteration in the hybridization state of O 2p and V 3d and O 2p and Fe 3d orbitals, respectively. Given the fact that FeVO4 has a bandgap of 2.2 eV, it would not be surprising that compression could induce the metallization of Fe0.9Al0.1VO4. Our results suggest that it could be interesting to investigate in the future the effect of pressure on the band structure of Fe0.9Al0.1VO4.
On further increase of pressure until 5.7 GPa, there are no evident changes in the diffraction profile other than the shift of peaks to higher angles due to lattice contraction. However, at 6.1 GPa, noticeable changes in the diffraction profile suggest the onset of a second phase transition. The extra peaks that indicate the existence of a second high-pressure phase are marked with the symbol “*” in Figure , while those originated by phase I′ are marked with the symbol “+”. Both phases coexist up to 8.55 GPa, and the transformation completes only at 9.15 GPa. This indicates that substitution of 10% Fe by Al has extended the range of stability of phase I′ from 4.8 to 8.55 GPa.
The partial substitution of Fe by Al affects not only the stability of phase I′ but also the structural sequence. Instead of the I′–II′ transition found in FeVO4, in Fe0.9Al0.1VO4, we found a different second HP phase. The Rietveld refinement presented in Figure supports that the second high-pressure phase could be assigned to a monoclinic structure isomorphic to FeVO4-IV (space group P2/c, Z = 2), which we will call phase IV in the following, for consistence with the notation of previous studies in FeVO4. , The unit-cell parameters determined at 9.15 GPa are found to be a = 4.385(1) Å, b = 5.452(3) Å, c = 4.796(7) Å, and β = 90.2(1)°. This HP structure resembles the wolframite phase previously found in CrVO4. The complete structural information, including atomic positions, is reported in Table . The goodness-of-fit parameters obtained were R wp = 2.33% and R p = 0.94%.
5.

Rietveld refinement of the XRD pattern of the monoclinic phase IV structure at 9.15 GPa (λ = 0.4956 Å). Data are shown as black crosses (×), while the red solid line represents the result of the refinement. The blue (green) line represents the residuals of the fitting (the background). Vertical pink (black) bars identify the Bragg reflections of Fe0.9Al0.1VO4 (copper).
3. Crystal Structure of the Monoclinic Phase IV of Fe0.9Al0.1VO4 at 9.15 GPa and Room Temperature.
| a = 4.385(1) Å, b = 5.452(3) Å, c = 4.796(7) Å, and β = 90.2(1)°. V = 114.6(2) Å3, Z = 2 | |||||
|---|---|---|---|---|---|
| atom | site | x | y | z | Uiso (Å2) |
| Fe/Al | 2f | 0.5000 | 0.670(2) | 0.2500 | 0.0104 |
| V | 2e | 0.0000 | 0.185(2) | 0.2500 | 0.0149 |
| O1 | 4g | 0.217(9) | 0.105(9) | 0.941(9) | 0.0139 |
| O2 | 4g | 0.248(9) | 0.368(9) | 0.408(9) | 0.0161 |
An earlier single-crystal XRD and DFT calculation investigation on FeVO4 has reported a similar transition at 4.8 GPa from FeVO4-I′ to FeVO4-II′ before the transformation to FeVO4-IV. Thus, in this work, we have demonstrated that substituting Fe by Al in FeVO4 Al considerably increases the pressure range of stability of phase I′ and alters the structural phase sequence which is I–I′–IV in Fe0.9Al0.1VO4 instead of I–I′–II′–IV as in FeVO4. In the second structural transition, I′–IV, there is no change in the coordination of the cations. The Fe/Al cation is octahedrally coordinated to six oxygen atoms, forming nearly regular octahedra. The FeO6 and VO6 octahedra share edges with octahedral units of the same cations and are connected via edge-sharing with the octahedra of the other cation forming alternating zigzag chains. The HP monoclinic phase IV remained stable until 11 GPa, which is the highest pressure of these measurements. On release of the pressure, the structure reverts to the HP triclinic phase I′ at 3 GPa. On further release of pressure to ambient conditions, both phases I′ and IV coexisted, suggesting that phase transition is not reversible. The observed nonreversibility of the transitions, along with the discontinuity in the unit-cell volume identified at each transition (as will be demonstrated below), suggests that both transitions exhibit first-order characteristics.
Figure a–d shows the pressure dependence of the lattice parameters of Fe0.9Al0.1VO4 until 11 GPa. Since the crystal structures of the three phases of this compound are triclinic and monoclinic, to analyze their compressibility, we have used the eigenvalues and eigenvectors (principal axes) of the compressibility tensor obtained using the only tool PASCal. For this analysis, we used results from the compression and decompression cycles. The results are given in Table , where eλi represents the direction of the principal axes of compressibility and κλi is the magnitude of the corresponding linear compressibility.
6.
Pressure dependence of unit-cell parameters of Fe0.9Al0.1VO4. Black, red, and blue circles are used for phases I, I′, and IV. Solid symbols are from compression experiments and empty symbols from decompression experiments. The symbols used for different angles are identified in the legend.
4. Principal Axes of Compressibility and the Corresponding Compressibility for the Three Polymorphs of Fe0.9Al0.1VO4 .
| phase | phase I | phase I′ | phase IV |
|---|---|---|---|
| eλ1 | (16̅8) | (0110) | (100) |
| κλ1 (in GPa–1) | 6.6(8) 10–3 | 1.2(6) 10–3 | 1.8(1) 10–3 |
| eλ2 | (193̅) | (85̅1) | (010) |
| κλ2 (in GPa–1) | 3.6(1) 10–3 | 2.1(2) 10–3 | 1.4(1) 10–3 |
| eλ3 | (332) | (772) | (001) |
| κλ3 (in GPa–1) | 0.5(4) 10–3 | 2.1(5) 10–3 | 1.2(1) 10–3 |
We found that in phase I, Fe0.9Al0.1VO4 compression is highly anisotropic. The most compressible direction is (16̅8) with a compressibility 1 order of magnitude larger than the less compressible axis (332), which surprisingly has a linear compressibility as low as that of diamond. The isomorphic structure FeVO4-I has been also found to have a highly anisotropic compressibility; however, the principal axes of compressibility differ from those we determined for phase I in Fe0.9Al0.1VO4. In the HP triclinic phase of Fe0.9Al0.1VO4 (phase I′), the anisotropy of compressibility is reduced with two axes with a compressibility of 2.1 × 10–3 GPa–1 and the less compressible axis with a compressibility of 1.2 × 10–3 GPa–1. In the HP monoclinic phase of Fe0.9Al0.1VO4, the anisotropy is even more reduced. It is noticeable that in this phase, the β angle is always within 90.2 and 90° being the structure of compounds isomorphic to phase IV considered as pseudo-orthorhombic in the literature. A consequence of the small deviation of the β angle from 90° is the fact that the principal axes of compressibility are aligned with the crystallographic axes.
Figure shows the P–V data for the three polymorphs of Fe0.9Al0.1VO4. In the figure, the compressibility of the studied compound decreases in the successive phase transitions. This observation is consistent with the successive collapses of the volume (9% and 13%, respectively). The abrupt changes in the volume and the irreversibility of the transitions indicate that the two transitions are first order in nature. The P–V data were fitted to the second-order Birch–Murnaghan equation of state (K 0 = 4) using EosFit7-GUI and considering both the bulk modulus (K o) and volume (V o) at 0 GPa as free parameters. The results of the fitting are summarized in Table . For the triclinic phase I of Fe0.9Al0.1VO4, a bulk modulus of 67(4) GPa is found, which is at least 12% smaller compared to the bulk modulus of FeVO4-I (see Table ). , For the first high-pressure phase, the EOS fit gives a bulk modulus of 132(11) GPa, which is at least 21% smaller than the bulk modulus of FeVO4-I′ (see Table ). Finally, for the second high-pressure phase, the EOS fit gives a bulk modulus of 150(9) GPa, which is at least 14% smaller than the bulk modulus of FeVO4-IV. Thus, substituting Fe by Al has altered not only the structural sequence of the compound but also its elastic properties. This observation is consistent with the fact that Al–O bonds are more compressible than Fe–O, as shown by the fact that Al2O3 has a bulk modulus of 250 GPa and γ-Fe2O3 a bulk modulus of 303 GPa.
7.

Volume versus pressure results for Fe0.9Al0.1VO4. Black, red, and blue circles are used for phases I, I′, and IV, respectively. Solid symbols are from compression experiments and empty symbols from decompression experiments. Solid lines represent Birch–Murnaghan fits to the data. The unit-cell volume of phase IV was multiplied by three to facilitate the comparison.
5. Bulk Modulus (K o) and Volume (V o) at 0 GPa for the Three Polymorphs of Fe0.9Al0.1VO4 and the Same Polymorphs of FeVO4 .
| phase |
Fe0.9Al0.1VO4
|
FeVO4
|
||
|---|---|---|---|---|
| V o present study (Å3) | Ko present study (GPa) | Ko previous studies (GPa) | reference | |
| phase I | 468.6(6) | 67(4) | 80(4) | single-crystal XRD |
| 87.3 | DFT calculations | |||
| 76(3) | powder XRD | |||
| phase I′ | 421(2) | 132(11) | 181(17) | single-crystal XRD |
| 167.5 | DFT calculations | |||
| phase IV | 363.3(9) | 150(9) | 174(8) | powder XRD |
| 206.7 | DFT calculations | |||
4. Conclusions
Our high-pressure synchrotron powder XRD experiments on Al-substituted triclinic FeVO4 (Fe0.9Al0.1VO4) suggest that at room temperature, the low-pressure triclinic phase undergoes a first-order phase transition at 2.85 GPa to another triclinic structure isomorphic to FeVO4-I′ with ∼9% volume collapse and a change in coordination of Fe and V cations. The onset of a second phase transition is observed at 6.1 GPa with coexistence of both phases until 8.55 GPa. The second high-pressure phase is identified as a monoclinic structure isomorphic to FeVO4-IV. The transition completes at 9.15 GPa, and the second high-pressure phase remains stable up to 11 GPa. These transitions are found to be irreversible as the two high-pressure phases coexist at ambient pressure. The structural sequence induced by pressure deviates from the one found in FeVO4, as the transition pressures are different, and no intermediate phases are found between phase I′ and IV. In addition, we have observed an alteration of the compressibilities in all the polymorphs of Al-substituted FeVO4. In particular, the three polymorphs of Fe0.9Al0.1VO4 are more compressible than the isomorphous polymorphs of FeVO4. In summary, a 10% substitution of Fe with Al clearly affects the high-pressure behavior of FeVO4. The inclusion of crystal in the crystal structure affects the structural sequence, which is different in Fe0.9Al0.1VO4 than in FeVO4, and modifies the mechanical properties of FeVO4. The present study shows that the role of cationic composition in the phase behavior of vanadates at high pressures should be carefully considered when modeling them.
Acknowledgments
The authors gratefully acknowledge the financial support from the Spanish Ministerio de Ciencia, Innovación y Universidades (DOI: 10.13039/501100011033) under projects PID2022-138076NB-C41 and RED2022-134388-T. D.E. thanks the financial support of Generalitat Valenciana through grants PROMETEO CIPROM/2021/075-GREENMAT, MFA/2022/007, and MFA/2022/025. D.E. and P.B. thank the financial support of Generalitat Valenciana through grant CIAPOS/2023/406. This study forms part of the Advanced Materials program and is supported by MCIN with funding from the European Union Next Generation EU (PRTR-C17.I1) and by the Generalitat Valenciana. The authors thank Elettra Sincrotrone Trieste (proposal no. 20235003).
The data that support the findings of this study are available from the corresponding author upon reasonable request.
V.P. formal analysis, methodology, writing original draft, and writingreview and editing. P.B., N.B., and F.G.A. methodology, data acquisition, and writingreview and editing. E.B. and M.B. synthesized and characterized the samples and writingreview and editing. D.E. conceptualization, formal analysis, writing original draft, and writing review and editing. All authors participated in the writing and editing of the manuscript. All authors have given approval to the final version of the manuscript.
The authors declare no competing financial interest.
References
- Setayesh S., Nazari P., Maghbool R.. Engineered FeVO4/CeO2 nanocomposite as a two-way superior electro-Fenton catalyst for model and real wastewater treatment. J. Eng. Sci. 2020;97:110–119. doi: 10.1016/j.jes.2020.04.035. [DOI] [PubMed] [Google Scholar]
- Yang M., Ma G., Yang H., Xiaoqiang Z., Yang W., Hou H.. Advanced strategies for promoting the photocatalytic performance of FeVO4 based photocatalysts: A review of recent progress. J. Alloys Compd. 2023;941:168995. doi: 10.1016/j.jallcom.2023.168995. [DOI] [Google Scholar]
- Pardeshi O., Naeem S., Patil A.. Synthesis of FeVO4 nanoparticles using sol-gel auto-combustion method and their application in supercapacitors. Energy Storage. 2024;6(5):683. doi: 10.1002/est2.683. [DOI] [Google Scholar]
- Xu H., Fan J., Pang D., Zheng Y., Chen G., Du F., Gogotsi Y., Dall’Agnese Y., Gao Y.. Synergy of ferric vanadate and MXene for high performance Li- and Na-ion batteries. Chem. Eng. J. 2022;436:135012. doi: 10.1016/j.cej.2022.135012. [DOI] [Google Scholar]
- Lu Q., Huang L., Li W., Wang T., Yu H., Hao X., Liang X., Liu F., Sun P., Lu G.. Mixed-potential ammonia sensor using Ag decorated FeVO4 sensing electrode for automobile in-situ exhaust environment monitoring. Sens. Actuators, B. 2021;348:130697. doi: 10.1016/j.snb.2021.130697. [DOI] [Google Scholar]
- Zhang M., Ma Y., Friedrich D., van de Krol R., Wong L., Abdi F.. Elucidation of opto-electronic and photoelectrochemical properties of FeVO4 photoanodes for solar water oxidation. J. Mater. Chem. A. 2018;6(2):548–555. doi: 10.1039/c7ta08923f. [DOI] [Google Scholar]
- Bera G., Surampalli A., Mishra A., Mal P., Reddy V., Banerjee A., Sagdeo A., Das P., Turpu G.. Magnetolattice coupling, magnetic frustration, and magnetoelectric effect in the Cr-doped FeVO4 multiferroic material and their correlation with structural phase transitions. Phys. Rev. B. 2019;100(1):014436. doi: 10.1103/PhysRevB.100.014436. [DOI] [Google Scholar]
- Dutta D. P., Ramakrishnan M., Roy M., Kumar A.. Effect of transition metal doping on the photocatalytic properties of FeVO4 nanoparticles. J. Photochem. Photobiol., A. 2017;335:102–111. doi: 10.1016/j.jphotochem.2016.11.022. [DOI] [Google Scholar]
- Braganza C., Salker A.. Al-Doped FeVO4 Nanoparticles for Vapour Phase Methylation of Phenol. ChemistrySelect. 2018;3(26):7602–7607. doi: 10.1002/slct.201800231. [DOI] [Google Scholar]
- Meng L., Guo R., Sun X., Li F., Peng J., Chen C., Li T., Deng J.. Enhanced electrochemical performance of a promising anode material FeVO4 by tungsten doping. Ceram. Int. 2020;46(13):21360–21366. doi: 10.1016/j.ceramint.2020.05.232. [DOI] [Google Scholar]
- Robertson B., Kostiner E.. Crystal structure and Mössbauer effect investigation of FeVO4 . J. Solid State Chem. 1972;4(1):29–37. doi: 10.1016/0022-4596(72)90128-4. [DOI] [Google Scholar]
- He Z., Yamaura J., Udea Y.. Flux growth and magnetic properties of FeVO4 single crystals. J. Solid State Chem. 2008;181(9):2346–2349. doi: 10.1016/j.jssc.2008.05.041. [DOI] [Google Scholar]
- Muller J., Joubert J.. Synthese sous haute pression d’oxygene d’une forme dense ordonnée de FeVO4 et mise en evidence d’une variétéallotropique de structure CrVO4 . J. Solid State Chem. 1975;14(1):8–13. doi: 10.1016/0022-4596(75)90355-2. [DOI] [Google Scholar]
- Hotta Y., Ueda Y., Nakayama N., Kosuge K., Kachi S., Shimada M., Koizumi M.. Pressure-products diagram of FexV1–xO2 system (0 ≤ x ≤ 0.5) J. Solid State Chem. 1984;55(3):314–319. doi: 10.1016/0022-4596(84)90283-4. [DOI] [Google Scholar]
- Lopez-Moreno S., Errandonea D., Pellicer-Porres J., Martínez-García D., Patwe S., Achary S., Tyagi A., Rodríguez-Hernandez P., Munoz A., Popescu C.. Stability of FeVO4 under Pressure: An X-ray Diffraction and First-Principles Study. Inorg. Chem. 2018;57(13):7860–7876. doi: 10.1021/acs.inorgchem.8b00984. [DOI] [PubMed] [Google Scholar]
- Gonzalez-Platas J., Lopez-Moreno S., Bandiello E., Bettinelli M., Errandonea D.. Precise Characterization of the Rich Structural Landscape Induced by Pressure in Multifunctional FeVO4 . Inorg. Chem. 2020;59(9):6623–6630. doi: 10.1021/acs.inorgchem.0c00772. [DOI] [PubMed] [Google Scholar]
- Lopez-Moreno S., Sánchez-Martín J., Bandiello E., Bettinelli M., Bull C., Ridley C., Errandonea D.. The effect of pressure on the band-gap energy in FePO4 and FeVO4 . J. Phys. Chem. Solids. 2023;183:111604. doi: 10.1016/j.jpcs.2023.111604. [DOI] [Google Scholar]
- Badding J. V.. High-pressure Synthesis, Characterization and Tuning of Solid State Materials. Annu. Rev. Mater. Sci. 1998;28:631–658. doi: 10.1146/annurev.matsci.28.1.631. [DOI] [Google Scholar]
- Brazhkin V. V.. High-pressure Synthesized Materials: Treasures and Hints. High Press. Res. 2007;27:333–351. doi: 10.1080/08957950701546956. [DOI] [Google Scholar]
- Li X., Fu Y., Pedesseau L., Guo P., Cuthriell S., Hadar I., Even J., Katan C., Stoumpos C. C., Schaller R. D.. et al. Negative Pressure Engineering with Large Cage Cations in 2D Halide Perovskites Causes Lattice Softening. J. Am. Chem. Soc. 2020;142:11486–11496. doi: 10.1021/jacs.0c03860. [DOI] [PubMed] [Google Scholar]
- Wang Y., Zhang L., Wang J., Li Q., Wang H., Gu L., Chen J., Deng J., Lin K., Huang L.. et al. Chemical-Pressure-Modulated BaTiO3 Thin Films with Large Spontaneous Polarization and High Curie Temperature. J. Am. Chem. Soc. 2021;143:6491–6497. doi: 10.1021/jacs.1c00605. [DOI] [PubMed] [Google Scholar]
- Lin K., Li Q., Yu R., Chen J., Attfield J. P., Xing X.. Chemical Pressure in Functional Materials. Chem. Soc. Rev. 2022;51:5351–5364. doi: 10.1039/D1CS00563D. [DOI] [PubMed] [Google Scholar]
- Hilleke K., Zurek E.. Rational Design of Superconducting Metal Hydrides via Chemical Pressure Tuning. Angew. Chem., Int. Ed. 2022;61:e202207589. doi: 10.1002/anie.202207589. [DOI] [PubMed] [Google Scholar]
- Gomes E. O., Gouveia A. F., Gracia L., Lobato A. ´., Recio J. M., Andrés J.. A chemical-pressure-induced phase transition controlled by lone electron pair activity. J. Phys. Chem. Lett. 2022;13:9883–9888. doi: 10.1021/acs.jpclett.2c02582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dewaele A., Loubeyre P., Mezouar M.. Equations of state of six metals above 94 GPa. Phys. Rev. B. 2004;70:094112. doi: 10.1103/PhysRevB.70.094112. [DOI] [Google Scholar]
- Klotz S., Chervin J., Munsch P., Le Marchand G.. Hydrostatic limits of 11 pressure transmitting media. J. Phys. D: Appl. Phys. 2009;42(7):075413. doi: 10.1088/0022-3727/42/7/075413. [DOI] [Google Scholar]
- Prescher C., Prakapenka C.. DIOPTAS: a program for reduction of two-dimensional X-ray diffraction data and data exploration. High Press. Res. 2015;35(3):223–230. doi: 10.1080/08957959.2015.1059835. [DOI] [Google Scholar]
- Toby B.. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001;34(2):210–213. doi: 10.1107/S0021889801002242. [DOI] [Google Scholar]
- Garg A. B., Muñoz A., Anzellini S., Sánchez-Martín J., Turnbull R., Díaz-Anichtchenko D., Popescu C., Errandonea D.. Role of GdO addition in the structural stability of cubic Gd2O3 at high pressures: Determination of the equation of states of GdO and Gd2O3 . Materialia. 2024;34:102064. doi: 10.1016/j.mtla.2024.102064. [DOI] [Google Scholar]
- Turnbull R., González-Platas J., Rodríguez F., Liang A., Popescu C., He Z., Santamaría-Pérez D., Rodríguez-Hernández P., Muñoz A., Errandonea D.. Pressure-Induced Phase Transition and Band Gap Decrease in Semiconducting β-Cu2V2O7 . Inorg. Chem. 2022;61(8):3697–3707. doi: 10.1021/acs.inorgchem.1c03878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Monteseguro V., Ruiz-Fuertes J., Contreras-García J., Rodríguez-Hernández P., Muñoz A., Errandonea D.. High pressure theoretical and experimental analysis of the bandgap of BaMoO4, PbMoO4, and CdMoO4 . Appl. Phys. Lett. 2019;115(1):012102. doi: 10.1063/1.5109780. [DOI] [Google Scholar]
- Errandonea D., Gomis O., García-Domene B., Pellicer-Porres J., Katari V., Achary S., Tyagi A., Popescu C.. New Polymorph of InVO4: A High-Pressure Structure with Six-Coordinated Vanadium. Inorg. Chem. 2013;52(21):12790–12798. doi: 10.1021/ic402043x. [DOI] [PubMed] [Google Scholar]
- Errandonea D.. High pressure crystal structures of orthovanadates and their properties. J. Appl. Phys. 2020;128:040903. doi: 10.1063/5.0016323. [DOI] [Google Scholar]
- Botella P., López-Moreno S., Errandonea D., Manjón F. J., Sans J. A., Vie D., Vomiero A.. High-pressure characterization of multifunctional CrVO4 . J. Phys.: Condens. Matter. 2020;32(38):385403. doi: 10.1088/1361-648X/ab9408. [DOI] [PubMed] [Google Scholar]
- Nye, J. F. Physical Properties of Crystals; Oxford University Press, 1957. [Google Scholar]
- Cliffe M. J., Goodwin A. L.. PASCal: a principal axis strain calculator for thermal expansion and compressibility determination. J. Appl. Crystallogr. 2012;45:1321–1329. doi: 10.1107/S0021889812043026. [DOI] [Google Scholar]
- Gnanasekar K. I., Jayaraman V., Prabhu E., Gnanasekaran T., Periaswami G.. Electrical and sensor properties of FeNbO4: a new sensor material. Sens. Actuators, B. 1999;55(2–3):170–174. doi: 10.1016/S0925-4005(99)00051-9. [DOI] [Google Scholar]
- Birch F.. Finite Elastic Strain of Cubic Crystals. Phys. Rev. 1947;71(11):809–824. doi: 10.1103/PhysRev.71.809. [DOI] [Google Scholar]
- Gonzalez-Platas J., Alvaro M., Nestola F., Angel R.. EosFit7-GUI: A new graphical user interface for equation of state calculations, analyses and teaching. J. Appl. Crystallogr. 2016;49(4):1377–1382. doi: 10.1107/s1600576716008050. [DOI] [Google Scholar]
- Errandonea D., Garg A.. Recent progress on the characterization of the high-pressure behaviour of AVO4 orthovanadates. Prog. Mater. Sci. 2018;97:123–169. doi: 10.1016/j.pmatsci.2018.04.004. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.



