Next Article in Journal
pH-Dependent Crystallization of 2-, 4-, 5-, and 6-Hydroxynicotinic Acids in Aqueous Media
Previous Article in Journal
3D-Printed Bioceramic Scaffolds with High Strength and High Precision
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Observation of Possible Pressure-Induced Phase Transformation in GdAlO3 Perovskite Using In Situ X-ray Diffraction

1
Center for the Study of Matter at Extreme Conditions (CeSMEC), Department of Mechanical and Materials Engineering, Florida International University, Miami, FL 33199, USA
2
Semiconductor Electronics Department, Lviv Polytechnic National University, 12 Bandera, 79013 Lviv, Ukraine
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(7), 1060; https://doi.org/10.3390/cryst13071060
Submission received: 30 May 2023 / Revised: 28 June 2023 / Accepted: 29 June 2023 / Published: 5 July 2023
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
Gadolinium aluminate perovskite (GdAlO3) was studied at high pressures of up to 23 GPa in a diamond anvil cell (DAC) using monochromatic synchrotron X-ray powder diffraction. Evidence of a pressure-induced phase transformation from orthorhombic (Pbnm) to rhombohedral (R 3 ¯ c) structure was observed at 21 GPa and further proved by DFT calculations. Before phase transition, the volumetric ratio of polyhedron A and B (i.e., VA/VB for ABX3 general notation) in the Pbnm phase continuously increased towards the ideal value of five at the transition, indicating a pressure-induced decrease in the structural distortion as opposed to the trend in many other orthorhombic perovskites (e.g., CaSnO3, CaGeO3, MgSiO3 and NaMgF3). Pressure–volume data of the Pbnm phase were fitted to the third-order Birch–Murnaghan equation of state yielding a bulk modulus (Ko) of 216 ± 7 GPa with a pressure derivative of the bulk modulus ( K o ) of 5.8 GPa (fixed). This work confirms the pressure-induced phase transformation from orthorhombic to a higher symmetry structure previously predicted in GdAlO3 perovskite.

1. Introduction

Perovskite materials are well-known solids with the general formula of ABO3, where A is a large cation occupying the voids with coordination number 12 (in the ideal cubic perovskite structure) and B is a smaller cation in octahedral oxygen coordination. Rare-earth aluminates with perovskite structure are potential materials for optical, electronic, structural, and magnetic applications such as colossal magneto resistant materials, substrates for high-temperature superconductors, solid oxide fuel cells and microwave dielectric materials [1,2,3,4,5,6,7]. A large number of rare-earth perovskite-type oxides have been studied because of their interesting properties. Many of those ABO3 compounds crystallize into an orthorhombic structure which deviates from the ideal cubic structure via the titling and distortion of the BO6 octahedra [8,9]. The distortion degree of perovskites can be described with the following parameters: the BO6 octahedral tilting and rotation, the Goldschmidt’s tolerance factor(t) and the polyhedron volume ratio VA/VB, where VA is the volume of A-site polyhedron and VB is that of the B-site octahedron. The values for the tilting angle ( B–O1–B), the rotation angle ( B–O2–B), t and VA/VB for an ideal perovskite are 180°, 180°, 1 and 5, respectively. The nature of this distortion has been related to the existence of different ionic radii within the crystal [10,11,12,13,14,15,16].
Pressure, as one of the important thermodynamic parameters plays a critical role in altering material properties. Under pressure, shortening of the effective ionic radius may vary depending on the ion itself and its coordinating environment. Zhao et al. [17] developed a model using the bond valence concept to predict the structural behavior of oxide perovskites at high pressure, introducing the ratio MA/MB, which determines the relative compressibility of the AO12 and BO6 sites due to the change in the average bond distance, where MA and MB describe the variation of the bond valence sum at the central cation in a polyhedral site. Based on this model, the relative polyhedral compressibility (i.e., VA/VB) determines the influence of pressure on the crystal structure of perovskite [18]. Some orthorhombic rare-earth perovskites such as LaFeO3 and PrFeO3 undergo a first-order transition to high-pressure phases, among which the high-pressure phase of LaFeO3 was reported to have a tetragonal symmetry with unknown structural details [19]. First-order reversible structural phase transition from the orthorhombic to rhombohedral R 3 ¯ c structure was observed in SmAlO3 and LaGaO3 near 10 GPa [20] and 2.5 GPa [21], respectively, by X-ray synchrotron powder diffraction.
GdAlO3 orthorhombic perovskite has been studied as a potential material for industrial applications considering its magnetic, optical, electronic, and structural properties [22,23,24,25,26,27]. The crystal structure of gadolinium aluminate has been reported at ambient conditions [22,28,29]. Ross et al. [30,31] studied the structural behavior of GdAlO3 at high pressures of up to 8 GPa and found that the abovementioned structural distortion decreases with increasing pressure. By extrapolating the metrical deviation of the orthorhombic perovskite from tetragonal symmetry (proportional to (b−a)/a0 where a0 is the value of the phase with a higher symmetry), they predicted an orthorhombic to higher-symmetry transition at about 12 GPa in GdAlO3. Nevertheless, such a transition has never been directly observed in any experimental study. In this work, we compressed GdAlO3 up to 23 GPa and observed signs of the predicted transition but at a higher pressure (21 GPa). Density functional theory (DFT) calculations indicate that the rhombohedral structure (R 3 ¯ c) found in LaGaO3 is favorable at high pressure over the low-pressure orthorhombic structure (Pbnm) of GdAlO3 and the tetragonal structure (I4/mcm) found in CeAlO3 [32].

2. Materials and Methods

The GdAlO3 sample was prepared through the solid-state reaction of Gd2O3 and Al2O3 by arc-melting in argon atmosphere. Details of the procedure can be found elsewhere [9,32,33]. High pressures were generated using a Mao–Bell diamond anvil cell (DAC). Two opposed diamond anvils with 300 µm diameter culets were compressed on a 302 stainless steel gasket with an initial thickness of 250 µm. The gasket was pre-indented with the anvils to a thickness of 60 µm and a hole of a 125 µm diameter was drilled at the center of the indentation to hold the sample. The synthesized sample was grounded into fine powder in an agate mortar and loaded into the 125 µm diameter sample chamber. A methanol–ethanol (4:1) mixture was loaded together with the sample and a small ruby sphere (pressure calibrant) as a pressure transmitting medium. The in situ high-pressure powder X-ray diffraction experiments were performed at the 16-IDB beamline of HPCAT (High Pressure Collaborative Access Team) at Advanced Photon Source (APS) of Argonne National Laboratory, Lemont, IL, USA. The monochromatic X-ray beam of a 0.3757 Å wavelength was focused to a spot size of 5 µm. Diffraction patterns were recorded using a MAR345 detector. During the experiment, the sample pressure was increased stepwise; fluorescence of the ruby sphere and sample diffraction were collected at each step. The sample pressure was determined using ruby fluorescence pressure scale [34,35]. The FIT2D software [36] was used to integrate the recorded 2-dimensional diffraction patterns to obtain intensity versus 2θ plots. Rietveld refinements were performed for both the ambient-pressure phase and the high-pressure phase at 21 GPa using the GSAS software with a GUI interface [37,38]. The volumetric compression data of the ambient-pressure phase were derived through full-spectrum Le Bail analysis of the in-situ X-ray diffraction patterns.
Density functional theory (DFT) calculations were performed using the Vienna ab initio simulation package (VASP 6.3) [39,40]. The generalized gradient approximation (GGA) [41] with the Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional [42] was used to describe exchange and correlation effects. Electron wave functions were expanded by a plane wave with a cutoff energy of 520 eV. The Gaussian smearing method was used with a width of 0.05 eV. A Monkhorst–Pack [43] k-point grid was set with a 0.03 1/Å k-point separation which provides a mesh density of >1000 per reciprocal atom. Atomic relaxation was performed until the changes in the electronic and ionic steps were less than 10−6 eV and 10−5 eV, respectively.

3. Results

A sequence of in situ X-ray diffraction patterns of the sample was taken up to 23 GPa (Figure 1). All the diffraction peaks in the initial diffraction pattern were identified as the Pbnm perovskite phase in agreement with previous works [9,29,31,44]. Disappearance of (111), (021), (113), (023)/(221), and (133) peaks indicates a possible pressure-induced structural phase transition at a pressure between 16.2 GPa and 21 GPa, and this transition is reversible as all these disappeared peaks reappear when the pressure is released back to ambient. From the symmetry relationship point of view, there are three higher symmetry phases for perovskite in the pathway from its Pbnm orthorhombic to P m 3 ¯ m cubic polymorphs, i.e., P4/mbm, I4/mcm and R 3 ¯ c [16]. However, differences in symmetry operations require that the (111) and (021) peaks in the Pbnm orthorhombic phase disappear in the I4/mcm tetragonal phase (i.e., hkl: h + k + l = 2n and 0kl: k, l = 2n) and the R 3 ¯ c rhombohedral phase while they are allowed in the P4/mbm tetragonal phase. Since the (111) and (021) peaks in the Pbnm orthorhombic phase completely disappeared when sample pressure increased to 21 GPa (Figure 1), the P4/mbm tetragonal phase was not considered. Therefore, we examined three possible structural models using the diffraction pattern at 21 GPa, i.e., Pbnm (the ambient pressure phase), R 3 ¯ c (the high-pressure phase observed in SmAlO3 [20] and I4/mcm (the structure of CeAlO3). As shown in Figure 2 and Table 1, the differences in structure refinements using the three different space groups are not significant enough to conclusively determine the favorable structure for the high-pressure phase. We then calculated the total energy of the three structure models as a function of pressure using DFT (Figure 3). The calculation results indicate that Pbnm is favorable at low pressures, whereas R 3 ¯ c is favorable at high pressures, although the calculated transition pressure (75 GPa) is much higher than the experimental pressure where the disappearance of diffraction peaks was observed. We realize that the methanol–ethanol (4:1) pressure medium solidifies at a pressure above 10 GPa [45], creating a non-hydrostatic pressure environment and therefore the sample may experience significant stresses at 21 GPa. Zhao et al. [46,47] were able to quantify the stress field using single-crystal X-ray diffraction and indicate that such a non-hydrostatic pressure environment in a DAC may affect and even induce the phase transition as stress fields can influence the structural distortions in perovskites. A more recent experimental study [48] demonstrated that applying non-hydrostatic pressure may lower the transition pressure in a DAC. Taking this stress influence and possibly large uncertainty in determining transition pressure using DFT calculations into consideration, we propose that the theoretically predicted Pbnm to R 3 ¯ c phase transition may occur at a pressure between 16.2 GPa and 21 GPa promoted by non-hydrostatic pressure. As shown in Table 1, the volume per formula unit of the R 3 ¯ c structure is nearly identical to the Pbnm structure within the experimental uncertainty. To further confirm the symmetry of the high-pressure phase, higher resolution structure characterization, e.g., in situ single crystal X-ray diffraction, is needed.
The structural deviation of the perovskite (ABX3) structure from its ideal cubic structure is commonly characterized by its BX6 octahedral titling. The Pbnm perovskite has a simple tilt system of aac+ [49,50] (the superscript + and − indicate that successive octahedra along an axis have the same tilt and opposite tilt respectively). Figure 3a shows the GdAlO3 structure model based on the refinement of the X-ray diffraction at ambient condition. The tilting angles along the a-axis and the b-axis are identical, and successive octahedra along each of these axes have opposite tilt, whereas successive octahedra along the c-axis have the same tilt. The R 3 ¯ c perovskite has a tilt system of aaa [49,50]. Figure 3b shows the high-pressure phase structure model of GdAlO3 based on the refinement of the X-ray diffraction at 21 GPa. There are three identical tilts along the a, b and c-axis, and successive octahedra along each of the axes have opposite tilt. The degree of structure distortion of perovskite can be evaluated using the volumetric ratio (VA/VB) between AX12 and BX6 polyhedron with an ideal value of five for undistorted cubic structure [15]. Assuming regular BX6 octahedra throughout the Pbnm perovskite structure, VA and VB can be calculated through its unit cell dimensions [15,51]. Evolution of the VA/VB ratio of the low-pressure phase in GdAlO3 with pressure is given in Figure 4 shows the VA/VB ratio as a function of pressure (Table 2). As pressure increases, the VA/VB ratio of the Pbnm phase approaches the value of five, indicating a transition to a higher symmetry. Note that during the phase transition (between 16.2 and 21 GPa), VA/VB goes beyond five at 16.2 GPa, indicating large pre-transition straining. Detailed reasoning for these phenomena is unclear. While coexistence of two phases may compromise refinement accuracy, simultaneously fitting two phases did not solve this issue in the Pbnm phase. Although the pure high-pressure phase was observed at 21 GPa, the transition likely starts at 16.2 GPa since significant weakening of (111), (021), (113), (023)/(221), and (133) peaks occurs at this pressure, which is still higher than the transition pressure (12 GPa) predicted by Ross et al. [30,31].
The compression data (Figure 5) show that the transition from the Pbnm to R 3 ¯ c phase in GdAlO3 at a pressure between 16.2 GPa and 21 GPa accompanied nearly no volumetric discontinuity. Nevertheless, the space group–subgroup theory indicates that the Pbnm to R 3 ¯ c transformation is a first-order phase transition [16]. Likely, the volumetric discontinuity is so small that it is unrecognizable with the experimental uncertainty of this study. The orthorhombic (Pnma) to rhombohedral (R 3 ¯ c) transition observed in GdAlO3 and LaGaO3 has a similar unrecognizable volume discontinuity [20,21].
The obtained volumetric data of the Pbnm phase were fitted into the third–order Birch–Murnaghan equation of state (EoS) [52]:
P = 3 2 K 0 V V 0 7 / 3 V V 0 5 / 3 x 1 + 3 4 K o 4 ) V V 0 2 / 3 1 ,
where P is the pressure applied, V is the volume of the unit cell under pressure, K0 is the bulk modulus, subscript 0 indicates the ambient pressure, and K0′ is the pressure derivative of bulk modulus. EoSFit 7.0 software [53] was used for fitting the experimental P–V data into the EoS. The fitting was weighted according to the uncertainty of each point [53] and K0′ was fixed at 5.8 based on the previous experimental data by Ross et al. [31]. The V0 and K0 determined through the fitting are 207.29 ± 0.01 Å3 and 216 ± 7 GPa, respectively. Due to the very limited data for the high-pressure R 3 ¯ c phase (only two points), no EoS fitting was performed.

4. Discussion

While temperature influence on the structure of perovskite has been extensively studied, investigations of pressure influence on perovskite structure are still limited. Pressure influence on the Pbnm perovskite structure is twofold. Some structural studies on MgSiO3 [54,55] and NaMgF3 [11,56] perovskite showed that the VA/VB ratio of the Pbnm perovskite structure decreases further away from the ideal value of five upon compression, i.e., pressure increases the distortion of the perovskite. Experimental studies of pressure effect on CaGeO3 perovskite structure presented controversial results. In situ extended X-ray absorption fine structure (EXAFS) study inferred a trend of reduced distortion under high pressure [57], while X-ray diffraction study indicated a reversed pressure effect [58]. Controversial results about the pressure influence also exist in theoretical calculations on CaSiO3 perovskite [59,60]. CaSnO3 is another example of perovskites that show increased distortion with increasing pressure [61]. Whether pressure enhances the distortion or reduces it depends on the relative compressibility of the AX6 octahedra and the BX12 polyhedron. In the case of GdAlO3 of this study, the GdO6 octahedra are less compressible than the AlO12 polyhedron, yielding an increase in the VA/VB ratio towards five. This is consistent with previous single-crystal X-ray diffraction study on GdAlO3 by Ross et al. [30,31], although their experiment was limited to the high pressure of 8 GPa and therefore did not observe any phase transition. An ab initio calculation [62] and a single-crystal X-ray diffraction study [63] on another rare-earth aluminate perovskite, YAlO3, also show that the AX6 octahedra are less compressible than the BX12 polyhedron. All of these align nicely with the model developed by Zhao et al. [17] which shows that in GdFeO3-type orthorhombic perovskites with both cations possessing a formal charge of 3+, the structural distortion decreases with increasing pressure and the structure evolves towards a higher symmetry at high pressure.
The bulk modulus (216 GPa) of Pbnm GdAlO3 derived in this study is higher than those reported previously by Bass et al. [64] using Brillouin spectroscopy (204 GPa) and by Ross et al. [31] using single-crystal X-ray diffraction (191 GPa). In perovskite-structured alkaline earth silicates, germanates, and titanates, a simple trend of a smaller unit formula volume leading to a larger bulk modulus has been observed [65]. In the case of rare-earth aluminate perovskite, a summary of the available bulk moduli is listed in Table 3. Figure 6 shows a plot of the bulk moduli at ambient conditions as a function of ionic radius of R3+ [66] in the rare-earth aluminate perovskite RAlO3 (R = La, Pr, Nd, Sm, Eu and Gd). The linear trendline defined by the data of LaAlO3, PrAlO3, NdAlO3, SmAlO3 and EuAlO3 indicates that the bulk modulus of RAlO3 increases as the ionic radius of R3+ decreases. The bulk modulus value of GdAlO3 derived from the current study agrees well with the general trend of bulk moduli as a function of the ionic radius (Figure 6). More systematic studies on the bulk moduli of the perovskite-structured RAlO3 aluminates for the entire rare-earth family will be interesting to generalize the influence of ionic radius on bulk modulus of the rare-earth aluminate perovskite. We also realize that the methanol–ethanol pressure medium may yield non-hydrostatic environment when the pressure rises above 10 GPa due to solidification [45]. Although enlarged experimental uncertainties because of the non-hydrostatic pressure have been taken into account through weighted EoS fitting, a future study using nearly hydrostatic pressure medium such as helium or neon is more desirable for a better characterization of the compression behavior and the phase transition.
Table 3. Summary of the available bulk moduli (K) of rare-earth aluminate perovskite RAlO3.
Table 3. Summary of the available bulk moduli (K) of rare-earth aluminate perovskite RAlO3.
CompositionSpace GroupK (GPa)ReferenceNoteUse for the Trendline
LaAlO3 R 3 ¯ c 190 ± 5[67] XRD ,   fit   using   P m 3 ¯ m , up to 40 GPa (improper space group and pressure range)no
LaAlO3 R 3 ¯ c 177 ± 4[68]XRD, up to 7 GPayes
LaAlO3 R 3 ¯ c 196[69]Brillouin scatteringyes
PrAlO3Pnma193 ± 1[70]XRD, up to 7 GPayes
NdAlO3 R 3 ¯ c 204[71]Ab-initio calculationyes
SmAlO3Pnma178 ± 14[72]Ultrasound, at 0.6 GPa, 2.2% porosity no
SmAlO3Pnma198 ± 4[64]Brillouin scatteringyes
EuAlO3Pnma203 ± 15[72]Ultrasound, at 0.6 GPa, 2.4% porosityno
EuAlO3Pnma213[73]Ultrasound, sample of [72], corrected for porosityyes
GdAlO3Pnma179 ± 15[72]Ultrasound, at 0.6 GPa, 2.4% porosityno
GdAlO3Pnma203 ± 4[64]Brillouin scatteringno
GdAlO3Pnma191 ± 1[31]XRD, up to 8 GPano
GdAlO3Pnma216 ± 7this studyXRD, up to 16 GPano
Figure 6. Bulk moduli of rare-earth aluminate perovskite RAlO3 (R = La, Pr, Nd, Sm, Eu and Gd) as a function of the ionic radius of R3+. The broken line represents the linear trendline based on the data of LaAlO3 (JZ 2004 [68], MC 2010 [69]), PrAlO3 [70], NdAlO3 [71], SmAlO3 [64] and EuAlO3 [73]. Data for GdAlO3 of this study, JB 1984 [64] and NR 2004 [31] are plotted to show the agreement of these data with the linear trendline.
Figure 6. Bulk moduli of rare-earth aluminate perovskite RAlO3 (R = La, Pr, Nd, Sm, Eu and Gd) as a function of the ionic radius of R3+. The broken line represents the linear trendline based on the data of LaAlO3 (JZ 2004 [68], MC 2010 [69]), PrAlO3 [70], NdAlO3 [71], SmAlO3 [64] and EuAlO3 [73]. Data for GdAlO3 of this study, JB 1984 [64] and NR 2004 [31] are plotted to show the agreement of these data with the linear trendline.
Crystals 13 01060 g006

5. Conclusions

The previously predicted pressure-induced phase transformation from orthorhombic (Pbnm) to a higher-symmetry structure in GdAlO3 has been observed in in situ X-ray diffraction and the higher-symmetry phase is identified with space group of R 3 ¯ c with the assistant of DFT calculations. The X-ray diffraction patterns indicate that the structural phase transformation occurs between 16 and 21 GPa, although the volumetric variations VA/VB of the Pbnm phase suggest the transition to a higher symmetry may apparently start at around 14 GPa. The structural phase transformation is a first-order transition according to the space groups but the volumetric discontinuity is indistinguishable with the experimental uncertainty of the current study. A bulk modulus (K0) of 216 ± 7 GPa with a fixed derivative bulk modulus (K0) at 5.8 was obtained for the Pbnm phase according to the third-order Birch–Murnaghan equation of state. The bulk modulus derived in this study is higher than the previously reported values but is in good agreement with the linear trend of increase in bulk modulus as the ionic radius of R3+ in rare-earth aluminate perovskite RAlO3 decreases.

Author Contributions

Conceptualization, J.C., S.S. and L.V.; Data curation, M.M.; Formal analysis, M.M., V.D. and J.C.; Funding acquisition, J.C.; Investigation, M.M., A.D., V.D., S.E., J.C., S.S., X.L. and L.V.; Methodology, M.M., A.D. and V.D.; Project administration, A.D. and V.D.; Resources, J.C. and S.S.; Supervision, J.C.; Validation, M.M. and X.L.; Visualization, S.E. and J.C.; Writing—original draft, M.M.; Writing—review and editing, S.E., J.C. and L.V. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported in part by NSF Grant No. EAR-1723185.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We thank Lyci George for her assistance in data collection. Portions of this work were performed at HPCAT (Sector 16), Advanced Photon Source (APS), Argonne National Laboratory. HPCAT operations are supported by DOE-NNSA’s Office of Experimental Sciences. The Ad-vanced Photon Source is a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory. L. Vasylechko acknowledges the support of the National Research Foundation of Ukraine under grant no. 2020.02/0373 “Crystalline phosphors’ engineering for biomedical applications, energy saving lighting and contactless thermometry”.

Conflicts of Interest

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

References

  1. Cho, S.-Y.; Kim, I.-T.; Hong, K.S. Microwave dielectric properties and applications of rare-earth aluminates. J. Mater. Res. 1999, 14, 114–119. [Google Scholar] [CrossRef]
  2. Henke, M.; Perbon, J.; Kuck, S. Preparation and spectroscopy of Yb2+-doped Y3Al5O12, YAlO3, and LiBaF3. J. Lumin. 2000, 87, 1049–1051. [Google Scholar] [CrossRef]
  3. Jin, H.; Rhim, S.H.; Im, J.; Freeman, A.J. Topological oxide insulator in cubic perovskite structure. Sci. Rep. 2013, 3, 1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lepe, F.J.; Fernández-Urbán, J.; Mestres, L.; Martínez-Sarrión, M.L. Synthesis and electrical properties of new rare-earth titanium perovskites for SOFC anode applications. J. Power Sour. 2005, 151, 74–78. [Google Scholar] [CrossRef]
  5. Lybye, D.; Bonanos, N. Proton and oxide ion conductivity of doped LaScO3. Solid State Ion. 1999, 125, 339–344. [Google Scholar] [CrossRef]
  6. Lybye, D.; Poulsen, F.W.; Mogensen, M. Conductivity of A- and B-site doped LaAlO3, LaGaO3, LaScO3 and LaInO3 perovskites. Solid State Ion. 2000, 128, 91–103. [Google Scholar] [CrossRef]
  7. Pickett, W.E.; Singh, D.J. Electronic structure and half-metallic transport in the La1-xCaxMnO3 system. Phys. Rev. B 1996, 53, 1146–1160. [Google Scholar] [CrossRef]
  8. Vasylechko, L.; Senyshyn, A.; Bismayer, U. Perovskites-Types Aluminates and Gallates; Gschneidner, K.A., Jr., Bünzli, J.C.G., Pechersky, V.K., Eds.; Elesvier, B.V.: Amsterdam, The Netherlands, 2009; Volume 39, pp. 113–295. [Google Scholar]
  9. Vasylechko, L.; Shmanko, H.; Ohon, N.; Prots, Y.; Hoffmann, S.; Ubizskii, S. Lattice crossover and phase transitions in NdAlO3—GdAlO3 system. J. Solid State Chem. 2013, 198, 101–107. [Google Scholar] [CrossRef] [Green Version]
  10. Avdeev, M.; Caspi, E.N.; Yakovlev, S. On the polyhedral volume ratios VA/VB in perovskites ABX3. Acta Crystallogr. Sect. B Struct. Sci. 2007, 63, 363–372. [Google Scholar] [CrossRef]
  11. Chen, J.; Liu, H.; Martin, C.D.; Parise, J.B.; Weidner, D.J. Crystal chemistry of NaMgF3 perovskite at high pressure and temperature. Am. Mineral. 2005, 90, 1534–1539. [Google Scholar] [CrossRef]
  12. Goldschmidt, V.M. The Laws of Crystal Chemistry (Die Gesetze der Krystallochemie). Die Nat. 1926, 14, 477–485. [Google Scholar] [CrossRef]
  13. Jacobsson, T.J.; Pazoki, M.; Hagfeldt, A.; Edvinsson, T. Goldschmidt’s Rules and Strontium Replacement in Lead Halogen Perovskite Solar Cells: Theory and Preliminary Experiments on CH3NH3SrI3. J. Phys. Chem. C 2015, 119, 25673–25683. [Google Scholar] [CrossRef]
  14. Magyari-Köpe, B.; Vitos, L.; Johansson, B.; Kollár, J. Origin of octahedral tilting in orthorhombic perovskites. Phys. Rev. B 2002, 66, 092103. [Google Scholar] [CrossRef]
  15. Thomas, N.W. The compositional dependence of octahedral tilting in orthorhombic and tetragonal perovskites. Acta Crystallogr. Sect. B Struct. Sci. 1996, 52, 16–31. [Google Scholar] [CrossRef] [Green Version]
  16. Thomas, N.W. A New Global Parameterization of Perovskite Structures. Acta Crystallogr. Sect. B Struct. Sci. 1998, B54, 585–599. [Google Scholar]
  17. Zhao, J.; Ross, N.L.; Angel, R.J. New view of the high-pressure behaviour of GdFeO3-type perovskites. Acta Crystallogr. Sect. B Struct. Sci. 2004, 60, 263–271. [Google Scholar] [CrossRef] [Green Version]
  18. Angel, R.J.; Ross, N.L.; Zhao, J. The compression of framework minerals: Beyond rigid polyhedra. Eur. J. Mineral. 2005, 17, 193–200. [Google Scholar] [CrossRef]
  19. Xu, W.M.; Naaman, O.; Rozenberg, G.K.; Pasternak, M.P.; Taylor, R.D. Pressure-induced breakdown of a correlated system: The progressive collapse of the Mott-Hubbard state in RFeO3. Phys. Rev. B 2001, 64, 094411. [Google Scholar] [CrossRef]
  20. Li, H.; Ma, S.; Ye, J.; Li, N.; Wang, X. The effects of pressure on the lattice of the rare-earth-based perovskite-type oxides SmAlO3 and NdAlO3. New J. Phys. 2022, 24, 113008. [Google Scholar] [CrossRef]
  21. Kennedy, B.J.; Vogt, T.; Martin, C.D.; Parise, J.B.; Hriljac, J.A. Pressure-induced orthorhombic to rhombohedral phase transition in LaGaO3. J. Phys. Condens. Matter 2001, 13, L925. [Google Scholar] [CrossRef]
  22. Du Boulay, D.; Lshizawa, N.; Maslen, E.N. GdAlO3 perovskite. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2004, 60, i120–i122. [Google Scholar] [CrossRef] [Green Version]
  23. Mahana, S.; Manju, U.; Topwal, D. Giant magnetocaloric effect in GdAlO3 and a comparative study with GdMnO3. J. Phys. D Appl. Phys. 2017, 50, 035002. [Google Scholar] [CrossRef]
  24. Oliveira, H.H.S.; Cebim, M.A.; Da Silva, A.A.; Davolos, M.R. Structural and optical properties of GdAlO3:RE3+ (RE = Eu or Tb) prepared by the Pechini method for application as X-ray phosphors. J. Alloys Compd. 2009, 488, 619–623. [Google Scholar]
  25. Qiu, L.M.; Numazawa, T.; Thummes, G. Performance improvement of a pulse tube cooler below 4 K by use of GdAlO3 regenerator material. Cryogenics 2001, 41, 693–696. [Google Scholar] [CrossRef]
  26. Watras, A.; Pazik, R.; Dereń, P.J. Optical properties of Ce3+ doped ABO3 perovskites (A = La, Gd, y and B = Al, Ga, Sc). J. Lumin. 2013, 133, 35–38. [Google Scholar] [CrossRef]
  27. Su, H.; Shen, Z.; Ma, W.; Liu, Y.; Zhao, D.; Guo, Y. Comprehensive microstructure regularization mechanism and microstructure–property stability at 1773 K of directionally solidified Al2O3/GdAlO3 eutectic ceramic composite. Compos. Part B Eng. 2023, 256, 110647. [Google Scholar] [CrossRef]
  28. Geller, S.; Balla, V. Crystallographic studies of perovskite-like compounds. I. Rare earth orthoferrites and YFeO3, YCrO3, YAlO3. Acta Crystallogr. 1956, 9, 1019–1025. [Google Scholar] [CrossRef]
  29. Vasylechko, L.; Matkovskii, A.; Senyshyn, A.; Savytskii, D.; Knapp, M.; Bähtz, C. Crystal structure and thermal expansion of orthorhombic perovskite—Type RE aluminates. HASYLAB Ann. Rep. 2003, 1, 251–252. [Google Scholar]
  30. Ross, N.L.; Zhao, J.; Angel, R.J. High-pressure structural behavior of GdAlO3 and GdFeO3 perovskites. J. Solid State Chem. 2004, 177, 3768–3775. [Google Scholar] [CrossRef]
  31. Ross, N.L.; Zhao, J.; Burt, J.B.; Chaplin, T.D. Equations of state of GdFeO3 and GdAlO3 perovskites. J. Phys. Condens. Matter 2004, 16, 5721–5730. [Google Scholar] [CrossRef]
  32. Vasylechko, L.; Senyshyn, A.; Trots, D.; Niewa, R.; Schnelle, W.; Knapp, M. CeAlO3 and Ce1−xRxAlO3 (R = La, Nd) solid solutions: Crystal structure, thermal expansion and phase transitions. J. Solid State Chem. 2007, 180, 1277–1290. [Google Scholar] [CrossRef]
  33. Ohon, N.; Vasylechko, L.; Prots, Y.; Schmidt, M. Phase and structural behavior of SmAlO3–RAlO3 (R = Eu, Gd) systems. Mater. Res. Bull. 2014, 50, 509–513. [Google Scholar] [CrossRef]
  34. Mao, H.K.; Xu, J.A.; Bell, P.M. Calibration of the ruby pressure gauge to 800 kbar under quasi-hydrostatic conditions. J. Geophys. Res. 1986, 91, 4673–4676. [Google Scholar]
  35. Sokolova, T.S.; Dorogokupets, P.I.; Litasov, K.D. Self-consistent pressure scales based on the equations of state for ruby, diamond, MgO, B2–NaCl, as well as Au, Pt, and other metals to 4 Mbar and 3000 K. Russ. Geol. Geophys. 2013, 54, 181–199. [Google Scholar] [CrossRef]
  36. Hammersley, A. FIT2D: A multi-purpose data reduction, analysis and visualization program. J. Appl. Crystallogr. 2016, 49, 646–652. [Google Scholar] [CrossRef]
  37. Larson, A.C.; Dreele, R.B.V. General Structure Analysis System (GSAS); Los Alamos National Laboratory Report LAUR 86–748; Los Alamos National Laboratory: Los Alamos, NM, USA, 2004; p. 224. [Google Scholar]
  38. Toby, B.H. A graphical user interface for GSAS. J. Appl. Cryst. 2001, 34, 210–213. [Google Scholar]
  39. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  40. 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, 15–50. [Google Scholar] [CrossRef]
  41. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  42. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  44. Chaudhury, S.; Parida, S.C.; Pillai, K.T.; Singh Mudher, K.D. High-temperature X-ray diffraction and specific heat studies on GdAlO3, Gd3Al5O12 and Gd4Al2O9. J. Solid State Chem. 2007, 180, 2393–2399. [Google Scholar] [CrossRef]
  45. Klotz, S.; Chervin, J.C.; Munsch, P.; Le Marchand, G. Hydrostatic limits of 11 pressure transmitting media. J. Phys. D Appl. Phys. 2009, 42, 075413. [Google Scholar] [CrossRef]
  46. Zhao, J.; Angel, R.; Ross, N. Effects of deviatoric stresses in the diamond-anvil pressure cell on single-crystal samples. J. Appl. Crystallogr. 2010, 43, 743–751. [Google Scholar] [CrossRef]
  47. Zhao, J.; Angel, R.; Ross, N. The structural variation of rhombohedral LaAlO3 perovskite under non-hydrostatic stress fields in a diamond-anvil cell. J. Phys. Condens. Matter Inst. Phys. J. 2011, 23, 175901. [Google Scholar] [CrossRef]
  48. Zhuang, Y.; Wu, L.; Gao, B.; Cui, Z.; Gou, H.; Zhang, D.; Zhu, S.; Hu, Q. Deviatoric stress-induced quasi-reconstructive phase transition in ZnTe. J. Mater. Chem. C 2020, 8, 3795–3799. [Google Scholar] [CrossRef]
  49. Glazer, A.M. The classification of tilted octahedra in perovskites. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1972, 28, 3384–3392. [Google Scholar] [CrossRef]
  50. Glazer, A.M. Simple ways of determining perovskite structures. Acta Crystallogr. Sect. A 1975, 31, 756–762. [Google Scholar] [CrossRef] [Green Version]
  51. Zhao, Y.; Weidner, D.J.; Parise, J.B.; Cox, D.E. Thermal expansion and structural distortion of perovskite—Data for NaMgF3 perovskite. Part I. Phys. Earth Planet. Inter. 1993, 76, 1–16. [Google Scholar]
  52. Birch, F. Finite strain isotherm and velocities for single-crystal and polycrystalline NaCl at high pressures and 300 K. J. Geophys. Res. 1978, 83, 1257–1268. [Google Scholar]
  53. Angel, R.J.; Gonzalez-Platas, J.; Alvaro, M. EosFit7c and a Fortran module (library) for equation of state calculations. Z. Fur Krist. 2014, 229, 405–419. [Google Scholar] [CrossRef]
  54. Horiuchi, H.; Ito, E.; Weidner, D.J. Perovskite-type MgSiO3; single-crystal X-ray diffraction study. Am. Mineral. 1987, 72, 357–360. [Google Scholar]
  55. Kudoh, Y.; Ito, E.; Takeda, H. Effect of pressure on the crystal structure of Perovskite-Type MgSiO3. Phys. Chem. Miner. 1987, 14, 350–354. [Google Scholar] [CrossRef]
  56. Zhao, Y.; Parise, J.B.; Wang, Y.; Kusaba, K.; Vaughan, M.T.; Weidner, D.J.; Kikegawa, T.; Chen, J.; Shimomura, O. High-pressure crystal chemistry of neighborite, NaMgFr: An angle-dispersive diffraction study using monochromatic synchrotron X-radiation. Am. Mineral. 1994, 79, 615–621. [Google Scholar]
  57. Andrault, D.; Poirier, J.P. Evolution of the distortion of perovskites under pressure: An EXAFS study of BaZrO3, SrZrO3 and CaGeO3. Phys. Chem. Miner. 1991, 18, 91–105. [Google Scholar] [CrossRef]
  58. Wu, X.; Qin, S.; Gu, T.T.; Yang, J.; Manthilake, G. Structural and elastic properties of CaGeO3 perovskite at high pressures. Phys. Earth Planet. Inter. 2011, 189, 151–156. [Google Scholar] [CrossRef]
  59. Jung, D.Y.; Oganov, A.R. Ab initio study of the high-pressure behavior of CaSiO3 perovskite. Phys. Chem. Miner. 2005, 32, 146–153. [Google Scholar] [CrossRef] [Green Version]
  60. Li, L.; Weidner, D.J.; Brodholt, J.; Alfè, D.; Price, G.D.; Caracas, R.; Wentzcovitch, R. Phase stability of CaSiO3 perovskite at high pressure and temperature: Insights from ab initio molecular dynamics. Phys. Earth Planet. Inter. 2006, 155, 260–268. [Google Scholar] [CrossRef]
  61. Zhao, J.; Ross, N.L.; Angel, R.J. Tilting and distortion of CaSnO3 perovskite to 7 GPa determined from single-crystal X-ray diffraction. Phys. Chem. Miner. 2004, 31, 299–305. [Google Scholar] [CrossRef]
  62. Wu, X.; Qin, S.; Wu, Z. Generalized gradient approximation calculations of the pressure-induced phase transition of YAlO3 perovskite. J. Phys. Condens. Matter 2006, 18, 3907–3916. [Google Scholar] [CrossRef] [Green Version]
  63. Ross, N.L.; Zhao, J.; Angel, R.J. High-pressure single-crystal X-ray diffraction study of YAlO3 perovskite. J. Solid State Chem. 2004, 177, 1276–1284. [Google Scholar] [CrossRef]
  64. Bass, J.D. Elasticity of single-crystal SmAlO3, GdAlO3 and ScAlO3 perovskites. Phys. Earth Planet. Inter. 1984, 36, 145–156. [Google Scholar] [CrossRef]
  65. Xiao, W.; Tan, D.; Zhou, W.; Liu, J.; Xu, J. Cubic perovskite polymorph of strontium metasilicate at high pressures. Am. Miner. 2013, 98, 2096–2104. [Google Scholar] [CrossRef]
  66. Hoshino, M.; Sanematsu, K.; Watanabe, Y. Chapter 279—REE Mineralogy and Resources. In Handbook on the Physics and Chemistry of Rare Earths; Jean-Claude, B., Vitalij, K.P., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; Volume 49, pp. 129–291. [Google Scholar]
  67. Bouvier, P.; Kreisel, J. Pressure-induced phase transition in LaAlO3. J. Phys. Condens. Matter 2002, 14, 3981. [Google Scholar] [CrossRef]
  68. Zhao, J.; Ross, N.L.; Angel, R.J. Polyhedral control of the rhombohedral to cubic phase transition in LaAlO3 perovskite. J. Phys. Condens. Matter 2004, 16, 8763–8773. [Google Scholar] [CrossRef]
  69. Carpenter, M.A.; Sinogeikin, S.V.; Bass, J.D.; Lakshtanov, D.L.; Jacobsen, S.D. Elastic relaxations associated with the—Transition in LaAlO3: I. Single crystal elastic moduli at room temperature. J. Phys. Condens. Matter 2010, 22, 035403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Zhao, J.; Ross, N.L.; Angel, R.J.; Carpenter, M.; Howard, C.J.; Pawlak, D.; Lukasiewicz, T. High-pressure crystallography of rhombohedral PrAlO3 perovskite. J. Phys. Condens. Matter Inst. Phys. J. 2009, 21, 235403. [Google Scholar] [CrossRef]
  71. Verma, A.K.; Modak, P. Ab-initio investigations of R3c to Pm3m transition in RAlO3 (La, Pr and Nd) perovskites under pressure. AIP Conf. Proc. 2013, 1512, 86–87. [Google Scholar] [CrossRef]
  72. Liebermann, R.C.; Jones, L.E.A.; Ringwood, A.E. Elasticity of aluminate, titanate, stannate and germanate compounds with the perovskite structure. Phys. Earth Planet. Inter. 1977, 14, 165–178. [Google Scholar] [CrossRef]
  73. Kung, J.; Rigden, S. Oxide perovskites: Pressure derivatives of the bulk and shear moduli. Phys. Chem. Miner. 1999, 26, 234–241. [Google Scholar] [CrossRef]
Figure 1. In situ synchrotron X-ray diffraction patterns of GdAlO3 at different pressures as indicated and room temperature (λ = 0.3757 Å). Red circles indicate the characteristic peaks that disappear when Pbnm to R 3 ¯ c phase transition occurs.
Figure 1. In situ synchrotron X-ray diffraction patterns of GdAlO3 at different pressures as indicated and room temperature (λ = 0.3757 Å). Red circles indicate the characteristic peaks that disappear when Pbnm to R 3 ¯ c phase transition occurs.
Crystals 13 01060 g001
Figure 2. Result of Rietveld refinement of the diffraction pattern of GdAlO3 at 21 GPa with (a) Pbnm, (b) I4/mcm and (c) R 3 ¯ c space groups. Crosses are experimentally observed data, red solid line is calculated data, vertical bars are locations of expected diffraction peaks, and blue line at the bottom is the difference between the observed and calculated data.
Figure 2. Result of Rietveld refinement of the diffraction pattern of GdAlO3 at 21 GPa with (a) Pbnm, (b) I4/mcm and (c) R 3 ¯ c space groups. Crosses are experimentally observed data, red solid line is calculated data, vertical bars are locations of expected diffraction peaks, and blue line at the bottom is the difference between the observed and calculated data.
Crystals 13 01060 g002
Figure 3. Crystal structure of GdAlO3 at (a) ambient pressure (Pbnm space group, aac+ tilt system), and (b) 21 GPa (R 3 ¯ c space group, aaa tilt system). Solid black lines indicate unit cell. Small red dots represent oxygen atoms and large pink spheres gadolinium atoms. Grey areas denote AlO6 octahedra.
Figure 3. Crystal structure of GdAlO3 at (a) ambient pressure (Pbnm space group, aac+ tilt system), and (b) 21 GPa (R 3 ¯ c space group, aaa tilt system). Solid black lines indicate unit cell. Small red dots represent oxygen atoms and large pink spheres gadolinium atoms. Grey areas denote AlO6 octahedra.
Crystals 13 01060 g003
Figure 4. Variations of VA/VB of the Pbnm (solid symbols) and R 3 ¯ c (open symbols) phase of GdAlO3 as a function of pressure. The broken line indicates the ideal value of VA/VB for undistorted cubic perovskite structure. The dotted line indicates the pressure of the predicted structural transition to from Pbnm to a higher symmetry [30,31]. The shaded area shows the pressure range where the Pbnm to R 3 ¯ c transition may actually happen under non-hydrostatic pressure.
Figure 4. Variations of VA/VB of the Pbnm (solid symbols) and R 3 ¯ c (open symbols) phase of GdAlO3 as a function of pressure. The broken line indicates the ideal value of VA/VB for undistorted cubic perovskite structure. The dotted line indicates the pressure of the predicted structural transition to from Pbnm to a higher symmetry [30,31]. The shaded area shows the pressure range where the Pbnm to R 3 ¯ c transition may actually happen under non-hydrostatic pressure.
Crystals 13 01060 g004
Figure 5. Variation of the volume per formula unit of GdAlO3 perovskite (triangles: the Pbnm phase, squares: the R 3 ¯ c phase) as a function of pressure. The experimental errors are either indicated by the bars or less than the size of the symbols. The solid line represents the third-order Birch–Murnaghan EoS fitting for the data of the Pbnm phase (triangles).
Figure 5. Variation of the volume per formula unit of GdAlO3 perovskite (triangles: the Pbnm phase, squares: the R 3 ¯ c phase) as a function of pressure. The experimental errors are either indicated by the bars or less than the size of the symbols. The solid line represents the third-order Birch–Murnaghan EoS fitting for the data of the Pbnm phase (triangles).
Crystals 13 01060 g005
Table 1. Results of Rietveld refinement of the diffraction pattern of GdAlO3 at 21 GPa using Pbnm, I4/mcm and R 3 ¯ c space groups.
Table 1. Results of Rietveld refinement of the diffraction pattern of GdAlO3 at 21 GPa using Pbnm, I4/mcm and R 3 ¯ c space groups.
AtomxyZa (Å)b (Å)c (Å)V (Å3)V/Z (Å3)wR
PbnmGd0.998(7)0.003(4)0.255.093(1)5.146(1)7.322(1)191.89(2)47.972.9%
Al00.50
O10.5(4)0.98(18)0.25
O20.75(8)0.25(10)0.00(7)
I4/mcmGd00.50.255.123(2)5.123(2)7.25(1)190.76(4)47.691.9%
Al000
O1000.25
O20.32(5)0.82(5)0
R 3 ¯ cGd0005.111(2)5.111(2)12.713(5)287.55(21)47.932.9%
Al0.421(3)00.25
O1000.25
Table 2. Unit cell parameters of GdAlO3 measured in the diamond anvil cell from 1 bar to 23 GPa. The figures in parentheses represent 1 esd of the last decimal place shown.
Table 2. Unit cell parameters of GdAlO3 measured in the diamond anvil cell from 1 bar to 23 GPa. The figures in parentheses represent 1 esd of the last decimal place shown.
P (GPa)σ * (GPa)a (Å)b (Å)c (Å)V (Å3)Va/Vb **
Pbnm0.0005.2534(7)5.2910(6)7.458(1)207.29(1)4.87
2.70.35.2289(9)5.2823(6)7.401(2)204.4(2)4.87
6.00.55.212(1)5.2558(8)7.368(2)201.82(2)4.90
8.80.85.2050(7)5.2373(8)7.352(1)200.42(1)4.94
11.91.05.188(1)5.218(1)7.316(2)198.06(3)4.96
14.41.05.173(4)5.176(3)7.308(3)195.64(3)5.01
16.21.25.175(1)5.120(1)7.371(1)195.29(3)5.04
R 3 ¯ c20.62.05.111(2)5.111(2)12.713(5)287.55(21)4.66
23.42.05.094(1)5.094(1)12.682(3)285.02(6)4.65
* Experimental uncertainty taking the nonhydrostatic pressure due to solidification of pressure medium into account [45]. ** Va and Vb are calculated from the unit cell dimensions.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mora, M.; Durygin, A.; Drozd, V.; Esdaille, S.; Chen, J.; Saxena, S.; Liang, X.; Vasylechko, L. Experimental Observation of Possible Pressure-Induced Phase Transformation in GdAlO3 Perovskite Using In Situ X-ray Diffraction. Crystals 2023, 13, 1060. https://doi.org/10.3390/cryst13071060

AMA Style

Mora M, Durygin A, Drozd V, Esdaille S, Chen J, Saxena S, Liang X, Vasylechko L. Experimental Observation of Possible Pressure-Induced Phase Transformation in GdAlO3 Perovskite Using In Situ X-ray Diffraction. Crystals. 2023; 13(7):1060. https://doi.org/10.3390/cryst13071060

Chicago/Turabian Style

Mora, Maria, Andriy Durygin, Vadym Drozd, Shanece Esdaille, Jiuhua Chen, Surendra Saxena, Xue Liang, and Leonid Vasylechko. 2023. "Experimental Observation of Possible Pressure-Induced Phase Transformation in GdAlO3 Perovskite Using In Situ X-ray Diffraction" Crystals 13, no. 7: 1060. https://doi.org/10.3390/cryst13071060

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

Article Metrics

Back to TopTop