Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Anisotropic conductivity in brownmillerite-type Ca2Fe2O5 observed by measurements on oriented floating-zone grown single crystals

Esther M. Curtisa, Josie E. Auckettab and Ivana Radosavljević Evans*a
aDepartment of Chemistry, Durham University, Lower Mountjoy, South Road, Durham, DH1 3LE, UK. E-mail: ivana.radosavljevic@durham.ac.uk
bAustralian Synchrotron, ANSTO, 800 Blackburn Road, Clayton, VIC 3168, Australia

Received 1st October 2025 , Accepted 9th November 2025

First published on 10th November 2025


Abstract

Brownmillerite-type Ca2Fe2O5 is a mixed ionic and electronic conductor with applications as an electrode material for solid oxide fuel cells and solid oxide electrolyser cells. Long-range oxide ion migration in Ca2Fe2O5 has been computationally predicted to be predominantly two-dimensional, restricted to the (ac) crystallographic plane. We have used the floating zone method to grow large high-quality single crystals of Ca2Fe2O5 and determine its conductivity on two differently oriented single crystal samples in order to directly probe the anisotropy of transport properties. Impedance spectroscopy measurements have shown the conductivity in the (ac)-plane to be up to one and a half order of magnitude higher than that parallel to the crystallographic b-axis. This degree of anisotropy of the conductivity is higher than that observed experimentally in most other oxide ion or mixed conductors belonging to the apatite and melilite structural family, highlighting the suitability of brownmillerite-type materials for applications in devices requiring components in oriented crystal or thin-film forms.


Introduction

Ca2Fe2O5 is a mixed ionic and electronic conductor with applications in solid oxide fuel cells (SOFCs) and solid oxide electrolyser cells (SOECs). Recent studies on Ca2Fe2O5 derivatives have shown that they can be used for the air electrode in SOECs, with Ca2Fe0.8Co1.2O5 and Ca2Fe1.95Mg0.05O5 having area specific resistances of 0.29 Ω cm2 at 700 °C and 0.17 Ω cm2 at 800 °C, respectively.1,2 Furthermore, Co-doped Ca2Fe2O5 has been demonstrated to be a cathode material for SOFCs3,4 and studies of Ca2Fe2O5 as an electrode material for Li ion batteries have been carried out.1,5–7 A deeper understanding of the anisotropic nature of ionic conductivity in Ca2Fe2O5 could allow for a tuning of the ionic contribution of the conductivity in applications utilising oriented thin films, similar to tuning conductivity in thin films using strain engineering.8–10

Ca2Fe2O5 adopts the layered brownmillerite-type structure. Brownmillerite-type materials have a general formula A2B2O5 and can be described as oxygen-deficient perovskites (A2B2O6) where one sixth of the oxygen atoms are replaced by ordered vacancies. The structure consists of alternating layers of corner-sharing BO6 octahedra and BO4 tetrahedra, such that the inter-layer and intra-layer ordering of the tetrahedra connected into chains determines the space group adopted by the material. The Ca2+ cations are located on the sites between the layers. At room-temperature, Ca2Fe2O5 crystallises in space group Pnma, with tetrahedral chains strictly ordered (Fig. 1). Around 700 °C, it undergoes a phase transition to an incommensurately-modulated phase, Imma(00γ)s00,11–14 followed by two further phase transitions at 1100 °C and 1300 °C.15 Ca2Fe2O5 melts congruently at 1447 °C.16


image file: d5ce00949a-f1.tif
Fig. 1 The room-temperature structure of Ca2Fe2O5 showing FeO4 tetrahedra (green), FeO6 octahedra (blue), Ca atoms (purple) and O atoms (red).

Transport properties of a number of brownmillerite-type materials have been reported in the literature. Ba2In2O5 is a modest oxide ion conductor before it transforms into a cubic disordered perovskite phase; its total conductivity is 9.6 × 10−5 S cm−1 at 600 °C, but reaches 0.1 S cm−1 at 1100 °C.17 Sr2Fe2O5 is a mixed conductor with oxide ion conductivity of 0.22 S cm−1 at 900 °C, above the structural phase transition to a disordered cubic perovskite phase.18,19 The anisotropic oxide ion migration in the brownmillerite form of Sr2Fe2O5 has been described in computational studies by both classical and ab initio molecular dynamics (AIMD), while in the high temperature form the material has a three-dimensional network of equivalent oxide ion migration pathways.20–23 Ca2Fe2O5 is also a mixed conductor with oxide ion conductivity of 4 × 10−3 S cm−1 at 900 °C.24 A number of chemical modifications have been attempted in order to raise its conductivity, e.g. partial substitution of Fe3+ by Al3+, Mn4+, Co3+ and Ga3+ cations.25–29 Only Co-doped Ca2Fe2O5 was found to have a higher total conductivity (5.6 S cm−1 at 900 °C) compared to Ca2Fe2O5 (total conductivity of 0.56 S cm−1 at 900 °C) and other brownmillerite-type mixed conductors.28 The layered structure of Ca2Fe2O5 (Fig. 1), in particular the intrinsic vacancies concentrated in the (ac)-plane suggest that ionic migration in this material could be anisotropic. This has also been computationally predicted, albeit without being quantified.22,30 In order to directly demonstrate and quantify the anisotropic nature of conductivity, measurements on oriented single crystal samples need to be performed. This has been reported for a small number of apatite-type (Ln9.33(XO4)6O2, Ln = La, Pr, Sm, Nd; X = Si, Ge and Nd(28+x)/3AlxSi6−xO26, x = 0, 1, 2)31–36 and melilite-type ([La1+xCa1−x]2[Ga]2[Ga2O7+x/2]2, 0.1 ≤ x ≤ 0.6)37 materials. In addition, the oxide ion diffusivity and surface exchange was measured on oriented single crystals of A2NiO4 (A = Nb, La, Pr).38,39

In this study, we experimentally demonstrated the highly anisotropic nature of oxide ion conductivity in Ca2Fe2O5 by conductivity measurements on oriented single crystals. We grew large, high-quality single crystals of Ca2Fe2O5 using the floating-zone (FZ) method. Oriented samples were characterised using Laue X-ray diffraction, and conductivity was measured by impedance spectroscopy both parallel to the b-axis and within the (ac)-plane. These measurements showed that oxide ion conductivity in Ca2Fe2O5 is predominantly two-dimensional, occurring preferentially within the layers containing the FeO4 tetrahedral groups.

Experimental

Synthesis of polycrystalline Ca2Fe2O5

20 g of polycrystalline Ca2Fe2O5 were prepared by conventional solid-state synthesis. CaCO3 (≥99.0%, Sigma-Aldrich) and Fe2O3 (≥99%, Sigma-Aldrich) were dried overnight (400 °C, 12 h) and stoichiometric amounts were weighed out. The powder mixture was ground, pressed into 13 mm diameter pellets and fired at 1100 °C for 36 h. After firing, the powder was ball milled (300 rpm, 30 min) and powder X-ray diffraction (PXRD) was utilised to confirm phase purity.

Floating-zone crystal growth

To prepare the feed and seed rods, Ca2Fe2O5 powder was packed into rubber tubes, evacuated, and pressed at 40 MPa using a hydrostatic press to give suitable rods for single crystal growth. The rods were sintered (1200 °C, 12 h) to increase their density. The sintered rods had an approximate diameter of 4.5 mm and lengths between 4–6 cm. A Crystal Systems Corporation optical FZ furnace (model FZ-T-10000-H-VII-VPO-PC) equipped with four 300 W halogen lamps was used for single crystal growth. The growth atmosphere was 20[thin space (1/6-em)]:[thin space (1/6-em)]80 O2[thin space (1/6-em)]:[thin space (1/6-em)]Ar flowing at 0.5 L min−1. The Ca2Fe2O5 feed rod was pre-melted in the FZ furnace to reduce bubble formation in the molten zone during the growth,40 by passing the feed rod through the lamp focus at 20 mm h−1, with the feed and seed rods counter-rotated at 30 rpm. Subsequently, single crystal of Ca2Fe2O5 was grown using the pre-melted feed rod at 1 mm h−1, with the feed and seed rods counter-rotated (30 rpm). A lamp power of ∼58% was required. The tip of the pre-melted feed rod was cleaved off and used as the seed. Once the molten zone was formed, the growth proceeded in a controlled manner for 48 h, requiring no manual interference.

Orientation and cutting of single crystal Ca2Fe2O5

To prepare oriented Ca2Fe2O5 single crystals for impedance measurements, the whole crystal was mounted in Crystalbond 509-3 resin and cut using a Logitech Model 15 diamond wire saw. Excess resin remaining on the cut slices was removed with acetone. Single crystal slices for conductivity measurements were cut from the final 2 cm of the crystal grown. Laue patterns were collected on these faces using a Real-Time Back-Reflection Laue Camera (Multiwire Laboratories Model MWL12). The Laue patterns were indexed using the NorthStar7.0 software and compared to a simulated Laue pattern for Ca2Fe2O5, generated in the program WinLaue. These Laue patterns confirmed that the slice for measurements parallel to the b-axis had been cut with the faces (where the electrodes would be applied) indexed as the (010) plane, and the slice for measurements within the (ac)-plane had been cut with the faces (where the electrodes would be applied) indexed as the (101) plane.

Powder X-ray diffraction

PXRD patterns were collected on a Bruker D8 Advance diffractometer with a Lynx-Eye PSD detector and Cu Kα radiation (λ = 1.5406 Å). All PXRD patterns used to monitor the progress of the solid-state reaction were collected at room temperature in the 2θ range 10–90° with a data collection time of 20–30 minutes. The pattern of the final polycrystalline sample was collected at room temperature in the 2θ range 10–120° with a data collection time of 10 hours.

Variable temperature X-ray diffraction patterns were recorded in 20 °C intervals on heating (50–990 °C) and cooling (980–40 °C). An Anton Parr HTK1200 high-temperature attachment was used, and the furnace temperature was calibrated with an external alumina standard. Patterns were recorded over a 2θ range of 10–120° with a collection time of an hour at each temperature. The resulting diffraction patterns were analysed by sequential Rietveld fitting.

All diffraction data were analysed using the Rietveld method41 implemented in the TOPAS Academic software.42,43 Refined parameters included unit-cell parameters, a single isotropic atomic displacement parameter, sample displacement parameter, background polynomial terms and pseudo-Voigt peak shape function terms.

Single crystal X-ray diffraction

Single crystal X-ray diffraction data were collected on a Bruker D8 VENTURE diffractometer, using an ImS-microsource, Mo Kα radiation (λ = 0.7107 Å) and a PHOTON III C7 MM CPAD detector. An Oxford Cryosystems Cryostream 700+ was used to cool the sample to 120 K. Data reduction was carried out in Apex5 software, including a multi-scan absorption correction with an additional spherical absorption correction, reducing Rint from 18% to 9%. The structure refinement was performed in Jana2006 (ref. 44 and 45) using the published structure of Ca2Fe2O5 as a starting model.14 The atomic positions and anisotropic atomic displacement parameters for all atoms were refined. All structure visualisation was carried out in the VESTA software.46

Impedance spectroscopy

Impedance spectroscopy data on polycrystalline Ca2Fe2O5 were collected on a 10 mm diameter pellet sintered at 1200 °C for 12 h, with a relative density of 89%. Data were also collected on the FZ grown single crystal samples of Ca2Fe2O5 oriented to measure parallel to the b-axis or in the (ac)-plane. The indexed (010) or (101) faces, respectively, of the oriented single crystal slices, and the flat faces of the sintered pellets, were coated with platinum ink (Melator Technologies) for use as electrodes. The samples were mounted on a Probostat A-6 cell and the electrodes were fired at 1000 °C for 30 min. Impedance measurements were carried out in air and dry N2 in 20 °C intervals between 200–1000 °C with a ramp rate of 2 °C min−1 and a frequency range of 10−1–107 Hz. Data were analysed by simple circuit fitting using the ZView/Zplot software (Scribner Associates).

Results and discussion

Powder X-ray diffraction of Ca2Fe2O5

PXRD patterns collected on the powder used for crystal growth were fitted using the brownmillerite structural model in space group Pnma.14 An excellent fit was obtained with no peaks unaccounted for, confirming the phase purity. The agreement factor Rwp = 6.01% was achieved, and the final Rietveld fit obtained is shown in Fig. 2.
image file: d5ce00949a-f2.tif
Fig. 2 Rietveld fit of the PXRD pattern of polycrystalline Ca2Fe2O5 used to prepare the feed and seed rods. The blue, red and grey curves show the observed pattern, calculated pattern and the difference curve, respectively.

Variable temperature PXRD data collected on Ca2Fe2O5 between 40 °C and 990 °C were analysed by sequential Rietveld refinement, using the brownmillerite structural model in space group Pnma. The temperature evolution of the unit cell parameters is shown in Fig. 3 and a surface plot of the diffraction data from which this dependence was extracted is given in Fig. S1. A subtle change of slope can be seen at around ∼400 °C. The onset of oxide ion mobility in Ca2Fe2O5 has been observed by 18O/16O isotope exchange reactions to occur from 477 °C,22 and this is presumably related to the subtle change of slope in the unit cell parameter trends observed from our X-ray diffraction data at around 400 °C, as well as in the changes of slope in the Arrhenius conductivity curves (vide infra). Similarly subtle changes in unit cell parameters as a function of temperature were correlated with similar changes in properties in brownmillerite-type Sr2ScGaO5.47 The temperature evolution of the a and c unit cell parameters (red and blue curves in Fig. 3) shows a small discontinuity and change in slope at 670 °C, coinciding with the structural phase transition to the Imma(00γ)s00 phase, where the incommensurate modulation is related to the tetrahedral chain ordering within the (ac)-plane.11,12,15


image file: d5ce00949a-f3.tif
Fig. 3 The variation of the unit cell parameters as a function of temperature (heating and cooling). Error bars are smaller than the data markers.

Oriented single crystals of Ca2Fe2O5

The crystal obtained by the FZ growth was cylindrical with a length of 6.5 cm and a diameter of 0.4 cm (Fig. 4). The last 3.5 cm of the crystal had large facets on opposite sides along the length of the crystal, previously attributed to the (ac)-plane using Laue diffraction.40 No visible grain boundaries were observed on the outside of the crystal past the first 0.5 cm of growth, with the relatively quick FZ domain selection process attributed to the use of the cleaved tip of the pre-melted feed as the seed.40 To confirm the crystal quality, a small fragment (10−4 mm3) of the large single crystal was characterised by single crystal X-ray diffraction. The structure was refined with unit cell parameters a = 5.4188(4), b = 14.7045(9), c = 5.5866(4) and space group Pnma, with further crystallographic data given in Table 1 and the structure shown in Fig. 1. Furthermore, a section of the single crystal was ground and PXRD was collected, and a good fit was obtained with an agreement factor Rwp = 9.31%. The Rietveld fit is given in Fig. S2. Single-domain slices of Ca2Fe2O5 from the FZ grown crystal were subsequently prepared and X-ray Laue back reflection patterns collected on the faces (Fig. 5). The sample oriented with the flat faces perpendicular to the b-axis (used to measure conductivity along the b-axis) is indexed with the (010) reflection almost in the centre of the pattern; this confirms that the faces are the (010) plane. The sample cut to measure conductivity within the (ac)-plane is indexed with the (101) reflection near the centre of the Laue pattern and the faces were confirmed to be the (101) plane. The sharp, distinct reflections present in the Laue patterns also confirm that the crystal quality is sufficient for conductivity measurements. The single domain nature of the sample was confirmed by collecting multiple Laue patterns across the surface of the crystal; these were identical at all points on the surface.
image file: d5ce00949a-f4.tif
Fig. 4 The Ca2Fe2O5 crystal mounted in resin in preparation of cutting with a diamond wire saw. The scale bar is in centimetres. The red arrow marks the seeding point.
Table 1 The crystallographic parameters from the structure refinement carried out on a crystal of Ca2Fe2O5 grown by the FZ method
Unit cell parameters  
Space group Pnma
a 5.4188(4)
b 14.7045(9)
c 5.5866(4)
V 445.1(1)
Number of reflections  
All 11[thin space (1/6-em)]961
Observed 6138
Merged 759
Number of refined parameters 46
Rsym/% 8.06
R/% 3.55
wR/% 2.69



image file: d5ce00949a-f5.tif
Fig. 5 The Laue patterns collected on the flat surfaces of the Ca2Fe2O5 single crystal slices (left) and simulated Laue patterns used to confirm indexing (right). a) The sample used for oriented conductivity measurements parallel to the b-axis and b) the sample used for oriented conductivity measurements within the (ac)-plane. The small square-shaped markings, circled in red, near the central beam hole in both images are detector artifacts.

Transport properties

Impedance spectroscopy data were recorded on a sintered pellet of Ca2Fe2O5 and the two oriented single crystal slices to compare the conductivity in a polycrystalline sample with that parallel to the b-axis and within the (ac)-plane. Complex plane plots for all three samples show a single semicircular response at low temperatures. The capacitance was calculated from the apex of the arc using the RC time constant relationship 1/ω = RC.48–50 The capacitance was used to attribute the response to a grain boundary response for the polycrystalline pellet (10−10–10−8 F) or a surface response for the single crystal samples (10−11–10−9 F).48 In all cases, the electrode response is flat at low temperatures, suggesting dominant electronic conductivity.50 At higher temperatures (above ∼500 °C), a Warburg response is observed, indicative of the ionic conductivity contribution.48 The total resistivity was estimated as the high intercept of the semicircular response. With increasing temperature, the grain boundary and surface responses disappeared, and above 318 °C, only the electrode response remained for all samples. In this region, the total resistivity was estimated as the high frequency intercept of the electrode response. Nyquist plots of the polycrystalline sample are given in Fig. S3.

Fig. 6 shows the total conductivity of oriented single crystal samples of Ca2Fe2O5, probing the properties within the (ac)-plane and along the b-axis, and that of the polycrystalline sample for comparison. The activation energy values calculated in three temperature regions (<460 °C, 460–650 °C, >700 °C) using the Arrhenius relationship given in eqn (1), where σ is the conductivity, σ0 a preexponential constant, Ea the activation energy, T the temperature and R the ideal gas constant.

 
image file: d5ce00949a-t1.tif(1)
The activation energies are given in Table 2. The onset of oxide ion mobility, observed by 18O/16O isotope exchange reactions, was reported by Paulus et al. to occur above 477 °C.22 Hence, the changes of slope around 460 °C seen in all conductivity curves in Fig. 6 can therefore be attributed to the contribution of this new process to the total conductivity. Furthermore, subtle change in the slope of the temperature dependence of the unit cell parameters around 400 °C, observed in our X-ray diffraction data (Fig. 3), suggests that structural changes they reflect are related to this new process contributing to the total conductivity. Further changes of slope particularly notable in the b-direction, occur at 700 °C, the temperature of the phase transition of Ca2Fe2O5 to an incommensurately modulated Imma(00γ)s00 structure.11,12 This phase transition introduces aperiodic ordering of the tetrahedral chains in the layers parallel to the (ac)-plane, which likely makes the exchange of oxide ions between these layers and the apical sites in the octahedral layers, observed in AIMD simulations by Paulus et al.,22 less facile, leading to pronounced increase in the activation energy along the b-direction. Impedance spectroscopy data collected in dry N2 show a lower total conductivity for all samples. This provides evidence for the p-type electronic conductivity in Ca2Fe2O5,24 with the lower oxygen partial pressure suppressing the electronic part of the conductivity. Above the phase transition to the incommensurate structure, the total conductivity measured in the two atmospheres starts to converge, indicating a diminishing contribution of electronic conductivity with increasing temperature. Transference numbers show an increase in the ionic component of the conductivity with increasing temperature, especially above the phase transition (Table 3). Similar behaviour has been reported for other ferrite mixed ionic and electronic conductors such as Sr4Fe6O13 and Gd3Fe5O12.51,52


image file: d5ce00949a-f6.tif
Fig. 6 Arrhenius plots of the conductivity of Ca2Fe2O5 measured on a polycrystalline sample (purple) and two single-crystal samples, oriented to measure conductivity parallel to the b-axis (red) and in the (ac)-plane (blue). Data collection was carried out in air (left) and flowing nitrogen (right). The dashed grey lines separate the three temperature regions for which the activation energies were calculated.
Table 2 The activation energies of Ca2Fe2O5 calculated in three different regions: below 460 °C, between 460–650 °C and above 700 °C. These were calculated for single crystal samples measured parallel to the b-axis, within the (ac)-plane, and a polycrystalline sample. Activation energies were calculated for data collected in both air and dry N2
Temperature/°C Activation energies/eV
Single crystal, parallel to the b-axis Single crystal, within the (ac)-plane Polycrystalline
<460 Air 0.501(6) 0.548(4) 0.54(3)
N2 0.85(2) 0.835(1) 0.84(7)
460–650 Air 0.37(1) 0.235(4) 0.25(2)
N2 0.44(4) 0.32(1) 0.38(10)
>700 Air 0.53(2) 0.164(4) 0.151(5)
N2 0.7(1) 0.27(2) 0.4(1)


Table 3 The transference numbers of Ca2Fe2O5 calculated from conductivity data collected in air and N2
Temperature/°C tion/%
Polycrystalline (ac)-Plane b-Axis
300 3.7(7) 10.4(9) 2.5(4)
400 10(1) 23(2) 6.3(6)
500 18(1) 32(2) 10.2(5)
600 24(2) 36(1) 11.6(9)
700 32(2) 38(1) 14(1)
800 41(4) 38(2) 16.5(8)


Directionality of oxide ion diffusion in Ca2Fe2O5

The diffusion mechanism of oxide ions in Ca2Fe2O5 has been modelled by AIMD simulations, predicting long-range oxide ion diffusion predominantly within the (ac)-plane.22 AIMD simulations at temperatures above 797 °C show a zigzag-to-zagzig switching of the orientation of the tetrahedral chains in Ca2Fe2O5 occurring on the picosecond time-scale.22 At these temperatures, the apical oxide ions within the octahedral layers start migrating into the oxygen vacancies within the tetrahedral layers.22 A study of Ca2Fe2O5 thin films reduction into quasi-2D layered CaFeO2 (ref. 30) showed that the reduction of the b-axis-oriented film took about four times longer than of the film oriented perpendicular to this direction. Since it was assumed that oxygen was lost predominantly from the face of the film, it was proposed that long-range oxide ion diffusion within Ca2Fe2O5 occurred predominantly within the (ac)-plane of the brownmillerite structure; however, the degree of anisotropy could not be quantified.

Through impedance spectroscopy measurements on oriented single crystal samples of Ca2Fe2O5, we demonstrated experimentally that the total conductivity within the (ac)-plane is up to one and a half orders of magnitude higher than that parallel to the b-axis (Fig. 6). The trends in conductivity between the samples are maintained above the phase transition into the incommensurately modulated structure; this is expected, as the high temperature Ca2Fe2O5 retains the layers present in the low temperature form.11,12,14,40 Previous studies of apatite-type materials (Ln9.33(XO4)6O2, Ln = La, Pr, Sm, Nd; X = Si, Ge and Nd(28+x)/3AlxSi6−xO26 for x = 0, 1, 2) and melilite-type [La1+xCa1−x]2[Ga]2[Ga2O7+x/2]2 show an increase of around half an order of magnitude in most cases, with only La9.33(SiO4)6O2 displaying a higher anisotropy of conductivity, of up to two orders of magnitude.31–37

Conclusions

Centimetre-sized high-quality single crystals of Ca2Fe2O5 were grown by the floating zone method and two differently oriented samples were prepared. The quality and orientation of the single crystals were confirmed using X-ray Laue diffraction. Impedance spectroscopy measurements were carried out on the oriented single crystals to probe the conductivity within the (ac)-plane and parallel to the b-axis, and additionally on a polycrystalline sample as a reference. Our study experimentally demonstrates that long-range oxide ion diffusion in brownmillerite-type Ca2Fe2O5 occurs predominantly in the (ac)-plane, consistent with theoretical predictions based on ab initio molecular dynamics simulations.20–23 The transport properties exhibit a significant degree of anisotropy, with the total conductivity measured within the (ac)-plane found to be up to one and a half orders of magnitude higher than that parallel to the b-axis. This suggests that Ca2Fe2O5 is a good candidate for solid state electrochemistry applications where thin films or oriented single crystals are required.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data supporting this article are available at https://doi.org/10.15128/r2j9602067k and in the supplementary information (SI) document.

Supplementary information is available. See DOI: https://doi.org/10.1039/d5ce00949a.

Acknowledgements

The authors thank the Leverhulme Trust for the provision of a PhD research project grant for EMC (RPG-2023-080). IRE thanks Durham University for research leave.

Notes and references

  1. E. Squizzato, G. Carollo and A. Glisenti, Int. J. Hydrogen Energy, 2021, 46, 24555–24566 CrossRef.
  2. S. J. Lee, S. M. Yong, D. S. Kim and D. K. Kim, Int. J. Hydrogen Energy, 2012, 37, 17217–17224 CrossRef.
  3. V. Cascos, R. Martínez-Coronado, J. A. Alonso and M. T. Fernández-Díaz, Int. J. Hydrogen Energy, 2015, 40, 5456–5468 CrossRef.
  4. Baijnath, P. Tiwari and S. Basu, Int. J. Hydrogen Energy, 2019, 44, 10059–10070 CrossRef.
  5. S. K. Sundriyal and Y. Sharma, Appl. Surf. Sci., 2021, 560, 150055 CrossRef.
  6. P. Zhu, G. Yang, X. Sun, Q. Cao, Y. Zhao, R. Ding, E. Liu and P. Gao, New J. Chem., 2023, 47, 11102–11109 RSC.
  7. G. Spanu, A. Celeste, F. Bozza, E. Serra, P. Torelli, L. Braglia, S. Brutti, P. Reale and L. Silvestri, Batteries Supercaps, 2024, 7, e202400063 CrossRef.
  8. G. Yang, M. El Loubani, H. R. Chalaki, J. Kim, J. K. Keum, C. M. Rouleau and D. Lee, ACS Appl. Electron. Mater., 2023, 5, 4556–4563 CrossRef PubMed.
  9. P. Arunkumar, R. Ramaseshan, S. Dash and K. S. Babu, Sci. Rep., 2017, 7, 3450 CrossRef PubMed.
  10. P. Arunkumar, P. Panda, M. Sribalaji, R. Ramaseshan, A. K. Keshri and K. S. Babu, Electrochim. Acta, 2017, 240, 437–446 CrossRef.
  11. H. Krüger and V. Kahlenberg, Acta Crystallogr., Sect. B: Struct. Sci., 2005, 61, 656–662 CrossRef PubMed.
  12. H. Krüger, V. Kahlenberg, V. Petříček, F. Phillipp and W. Wertl, J. Solid State Chem., 2009, 182, 1515–1523 CrossRef.
  13. T. Labii, M. Ceretti, A. Boubertakh, W. Paulus and S. Hamamda, J. Therm. Anal. Calorim., 2013, 112, 865–870 CrossRef.
  14. P. Berastegui, S.-G. Eriksson and S. Hull, Mater. Res. Bull., 1999, 34, 303–314 CrossRef.
  15. E. Asenath-Smith, S. Misture and D. Edwards, J. Solid State Chem., 2011, 184, 2167–2177 CrossRef.
  16. S. Cheng, E. Nekhoroshev, M. Shevchenko, D. Shishin and E. Jak, Ceram. Int., 2024, 50, 45438–45452 CrossRef.
  17. J. B. Goodenough, J. E. Ruiz-Diaz and Y. S. Zhen, Solid State Ion., 1990, 44, 21–31 CrossRef.
  18. I. A. Leonidov, M. V. Patrakeev, J. A. Bahteeva, K. V. Poholok, D. S. Filimonov, K. R. Poeppelmeier and V. L. Kozhevnikov, J. Solid State Chem., 2006, 179, 3045–3051 CrossRef.
  19. M. Schmidt and S. J. Campbell, J. Solid State Chem., 2001, 156, 292–304 CrossRef.
  20. S. Ambaum, N. L. Allan, R. Dittmann and R. A. De Souza, Chem. Mater., 2024, 36, 2039–2048 CrossRef.
  21. J. E. Auckett, W. T. Lee, K. C. Rule, A. Bosak and C. D. Ling, Inorg. Chem., 2019, 58, 12317–12324 CrossRef PubMed.
  22. W. Paulus, H. Schober, S. Eibl, M. Johnson, T. Berthier, O. Hernandez, M. Ceretti, M. Plazanet, K. Conder and C. Lamberti, J. Am. Chem. Soc., 2008, 130, 16080–16085 CrossRef PubMed.
  23. J. E. Auckett, A. J. Studer, E. Pellegrini, J. Ollivier, M. R. Johnson, H. Schober, W. Miiller and C. D. Ling, Chem. Mater., 2013, 25, 3080–3087 CrossRef.
  24. A. L. Shaula, Y. V. Pivak, J. C. Waerenborgh, P. Gaczyñski, A. A. Yaremchenko and V. V. Kharton, Solid State Ionics, 2006, 177, 2923–2930 CrossRef.
  25. Q. E. Stahl, G. J. Redhammer, G. Tippelt and A. Reyer, Phys. Chem. Miner., 2019, 46, 271–298 CrossRef.
  26. R. K. Hona, A. Huq and F. Ramezanipour, Mater. Chem. Phys., 2019, 238, 121924 CrossRef.
  27. R. K. Hona, A. Huq and F. Ramezanipour, Ionics, 2019, 25, 1315–1321 CrossRef.
  28. A. L. Shaula, A. A. Markov, E. N. Naumovich, J. C. Waerenborgh, Y. V. Pivak and V. V. Kharton, Solid State Ionics, 2012, 225, 206–210 CrossRef.
  29. V. V. Kharton, I. P. Marozau, N. P. Vyshatko, A. L. Shaula, A. P. Viskup, E. N. Naumovich and F. M. B. Marques, Mater. Res. Bull., 2003, 38, 773–782 CrossRef.
  30. S. Inoue, M. Kawai, N. Ichikawa, H. Kageyama, W. Paulus and Y. Shimakawa, Nat. Chem., 2010, 2, 213–217 CrossRef PubMed.
  31. S. Nakayama and M. Highchi, J. Mater. Sci. Lett., 2001, 20, 913–915 CrossRef.
  32. S. Nakayama, M. Sakamoto, M. Higuchi, K. Kodaira, M. Sato, S. Kakita, T. Suzuki and K. Itoh, J. Eur. Ceram. Soc., 1999, 19, 507–510 CrossRef.
  33. M. Higuchi, K. Kodaira and S. Nakayama, J. Cryst. Growth, 2000, 216, 317–321 CrossRef.
  34. S. Nakayama and M. Sakamoto, Solid State Ionics, 2013, 253, 47–52 CrossRef.
  35. T. An, T. Baikie, F. Wei, S. S. Pramana, M. K. Schreyer, R. O. Piltz, J. F. Shin, J. Wei, P. R. Slater and T. J. White, Chem. Mater., 2013, 25, 1109–1120 CrossRef.
  36. Y. Masubuchi, M. Higuchi, S. Kikkawa, K. Kodaira and S. Nakayama, Solid State Ionics, 2004, 175, 357–360 CrossRef.
  37. F. Wei, H. Gasparyan, P. J. Keenan, M. Gutmann, Y. Fang, T. Baikie, J. B. Claridge, P. R. Slater, C. L. Kloc and T. J. White, J. Mater. Chem. A, 2015, 3, 3091–3096 RSC.
  38. J. M. Bassat, M. Burriel, O. Wahyudi, R. Castaing, M. Ceretti, P. Veber, I. Weill, A. Villesuzanne, J. C. Grenier, W. Paulus and J. A. Kilner, J. Phys. Chem. C, 2013, 117, 26466–26472 CrossRef.
  39. J. M. Bassat, P. Odier, A. Villesuzanne, C. Marin and M. Pouchard, Solid State Ionics, 2004, 167, 341–347 CrossRef.
  40. J. E. Auckett, A. J. Studer and C. D. Ling, Aust. J. Chem., 2014, 67, 1824–1828 CrossRef.
  41. H. M. Rietveld, J. Appl. Crystallogr., 1969, 2, 65 CrossRef.
  42. R. E. Dinnebier, A. Leineweber and J. S. O. Evans, Rietveld Refinement: Practical Powder Diffraction Pattern Analysis using TOPAS, De Gruyter, 2010 Search PubMed.
  43. A. A. Coelho, J. Evans, I. Evans, A. Kern and S. Parsons, Powder Diffr., 2011, 26, S22 CrossRef.
  44. V. Petříček, L. Palatinus, J. Plášil and M. Dušek, Z. Kristallogr. - Cryst. Mater., 2023, 238, 271–282 CrossRef.
  45. V. Petříček, M. Dušek and L. Palatinus, Z. Kristallogr. - Cryst. Mater., 2014, 229, 345–352 CrossRef.
  46. K. Momma and F. Izumi, J. Appl. Crystallogr., 2008, 41, 653–658 CrossRef.
  47. C. A. Fuller, Q. Berrod, B. Frick, M. R. Johnson, S. J. Clark, J. S. O. Evans and I. R. Evans, Chem. Mater., 2019, 31, 7395–7404 CrossRef.
  48. J. T. S. Irvine, D. C. Sinclair and A. R. West, Adv. Mater., 1990, 2, 132–138 CrossRef.
  49. Impedance Spectroscopy Theory, Experiment and Applicatons, ed. E. Barsoukov and J. R. Macdonald, Wiley-Interscience, New Jersey, 2nd edn, 2005 Search PubMed.
  50. A. Ch. Lazanas and M. I. Prodromidis, ACS Meas. Sci. Au, 2023, 3, 162–193 CrossRef PubMed.
  51. V. V. Kharton, A. L. Shaula, E. N. Naumovich, N. P. Vyshatko, I. P. Marozau, A. P. Viskup and F. M. B. Marques, J. Electrochem. Soc., 2003, 150, J33–J42 CrossRef.
  52. M. Y. Avdeev, M. V. Patrakeev, V. V. Kharton and J. R. Frade, J. Solid State Electrochem., 2002, 6, 217–224 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.