Open Access Article
Boris
Politov
*a,
Igor
Shein
b and
Susana
Garcia-Martin
*a
aUniversidad Complutense de Madrid, Madrid, 28040, Spain. E-mail: bpolitov@ucm.es; sgmartin@quim.ucm.es
bInstitute of Electrophysics of the UB RAS, Yekaterinburg, 620016, Russia
First published on 23rd September 2025
Profound knowledge of the electronic structure of functional solids is essential to understand and optimize their properties. The current advances in electronic structure theory, together with the improvements in computing power, allow realization of affordable calculations of the electronic structure of complex solids with the aim of explaining or predicting properties of singular materials. This work presents a density functional theory study of the GdBa2Ca2Fe5O13 oxide, a potential air electrode for solid oxide fuel cells with a layered-perovskite-related structure, which presents ordering of three different coordination-polyhedra around the Fe3+ ion (FeO6 octahedra, FeO5 squared pyramids and FeO4 tetrahedra). The existence of these three different Fe3+-environments significantly impacts the electronic properties of this oxide leading to a narrow band gap. The band structure calculations of GdBa2Ca2Fe5O13 concludes that the FeO5 layers create the CB (conduction band), the FeO6-layers form the VB (valence band) and the FeO4 layers create insulating channels, leading to anisotropic electrical properties that coincide with the experimentally observed 2D magnetic, electrical, and structural characteristics of GdBa2Ca2Fe5O13.
Coupling of cation and anion-vacancy orderings to form perovskite-based superstructures is clearly perceived in some complex Fe3+-perovskites, where the Fe3+ cations can adopt different oxygen coordination.28,29 Thought provoking is the modulation of the nuclear and magnetic structures in the Ln0.8−xBa0.8Ca0.4+xFe2O6−δ systems (Ln = Gd, Tb).30,31 Three single-phase oxides have been isolated in the system, which are better formulated as Ln1.2Ba1.2Ca0.6Fe3O8, Ln2.2Ba3.2Ca2.5Fe8O21 and LnBa2Ca2Fe5O13 (corresponding to x = 0.00, 0.25 and 0.40 respectively) in agreement with their oxygen content; the compounds constitute the A3m+5nFe3m+5nO8m+13n homologous series with m and n values 1, 0; 1, 1; and 0, 1, respectively. These three complex perovskite-related oxides present different layered-ordering of the A-site cations in combination with different oxygen-coordination environments around the Fe3+ atoms (octahedra, tetrahedra and squared pyramids) along the stacking c-axis. In the system, substitution of Ln3+ by Ca2+ decreases the oxygen content, which is accommodated by formation of FeO5-squared pyramid layers intercalated with one A-layer containing mainly Ln3+ cations. Like in layered Co-perovskites,20,21 the isolated Ln3+ layers might favor the anion conductivity through them, explaining that LnBa2Ca2Fe5O13, with a unit cell with the highest isolated Gd3+ layers rate in the series, also presents, in addition to the highest electronic conductivity, the best electrocatalytic properties.30 The antiferromagnetic behavior of these Fe3+ containing oxides is also influenced by the ordering and type of oxygen-polyhedra around the Fe atoms, demonstrating, once again, the relationship between crystal and electronic properties.31 Interestingly, the variation of their electrical conductivity with temperature, typical for a semiconducting mechanism, shows a smooth transition in a relatively wide range of temperatures (between ∼600 K and ∼900 K). The electrical transition is coupled with a broad magnetic transition, in the same temperature range, associated with the evolution of 3D into 2D-magnetic behavior.30,31
The studies on these complex layered perovskites are primarily focused on their physical properties in connection with their crystal structure, while the electronic structure is scarcely considered. Relatively recent advances in electronic structure theory, in particular density functional theory (DFT), in parallel with the improvements in computing power, allow explaining a great variety of materials properties and predict new compounds with certain thermodynamically stable crystal structures and behaviors. We present here a thorough computational study on the layered-perovskite GdBa2Ca2Fe5O13 (GBCFO), a potential oxygen electrode material for SOFCs. The work is focused on determining the electronic structure of the oxide to understand its functional properties and, therefore, to establish adequate relationships between its phase composition, crystal, and electronic characteristics. Among the oxides of the A3m+5nFe3m+5nO8m+13n series, GdBa2Ca2Fe5O13 presents the highest electronic conductivity and oxygen reduction reaction activity.30 In addition, the crystal structure of the studied oxide contains three types of oxygen-polyhedra around the Fe3+ cations (octahedra, tetrahedra and squared pyramids), which also seem to be associated with its superior properties. The results of the calculations also elucidate the previously observed 2D magnetic nature interplayed with the electrical behavior of the GdBa2Ca2Fe5O13, suggesting new approaches for further comprehensive experimental research in layered perovskites with 2D-behaviors.
The crystal structure of the GdBa2Ca2Fe5O13 oxide was modelled based on the √2ap × √2ap × 10ap unit cell (ap refers to the lattice parameter of the cubic perovskite), with orthorhombic symmetry (Imma space group), that contains 4 formula units and 92 atoms in total. In this work, the unit cell and atomic coordinates obtained from the structural refinement of the isostructural TbBa2Ca2Fe5O13 (TBCFO) and Y0.9Ba1.7Ca2.4Fe5O13 (YBCFO) oxides using neutron diffraction data were taken as a reference.30,31,44–47 The graphic representation of the unit cell is depicted in Fig. 1;30 the respective unit cell directions in Fig. 1 are given in terms of the symmetry of a simple cubic perovskite lattice.30 Three different magnetic structures (with respect to Fe3+ 3d electrons) were considered for the calculations (Fig. 1): one ferromagnetic (FM) and two antiferromagnetic (AFM) orderings: A-AFM and G-AFM.
To assess the thermodynamic stability of various magnetic configurations of the GdBa2Ca2Fe5O13 oxide, the total electronic energies of the Gd2O3, Fe2O3, BaO and CaO oxides were calculated. Defect formation energies in GBCFO were computed for the 2ap × 2ap × 10ap supercells by utilizing the approach reported earlier.48 Oxygen interstitials were created by introducing an additional oxygen atom in the pristine 184-atomic supercell; in turn, oxygen vacancies were made by removing one of oxygen atoms from it. The scheme of defect allocations used is provided in the SI.
Total (DOS) and partial (pDOS) density of states spectra of the GdBa2Ca2Fe5O13 oxide were computed by utilizing the Gaussian smearing method, since the unit cell parameter c is almost an order of magnitude larger compared to other typical perovskite unit cell dimensions; accordingly, the k-point sampling of the Brillouin zone (BZ) was selected to be N × N × 1, where N is a natural number. During the optimization procedure N was taken to be 4, while the electronic properties were calculated assuming N = 8. The band centers of mass were computed like in BaFeO3.49 Intralayer band gaps were estimated based on the position of O 2p and Fe 3d orbitals of oxygen and Fe atoms located in the different layers. Band dispersion was evaluated for the preliminarily optimized GBCFO unit cell under the constraint of preserving orthorhombic symmetry and G-AFM magnetic ordering; the distances between two adjacent high-symmetry points in BZ were interpolated by 25 k-points. Optical spectra were computed via a common approach reported previously.50 The pre- and post-processing of the results was accomplished using vaspKIT software.51 Additional calculations were made for estimating Fe–O bond strength using the LOBSTER package.52
To study the influence of the lattice arrangement of different FeOx polyhedra on the electronic properties of the resulting compounds, several alternative Fe3+-based crystal structures were investigated using the same GGA+U methodology as described above. Namely, orthorhombic GdFeO3, tetragonal BaCaFe2O5 and hexagonal CaBaFe4O8 were utilized to represent the contribution solely from FeO6 octahedra, FeO5 square pyramids and FeO4 tetrahedra, respectively. In addition, combinations of two different polyhedra (i.e. FeO4 + FeO5; FeO4 + FeO6 and FeO5 + FeO6) were assessed with the help of NaCa2Fe3O7, GdBaCaFe3O8 and GdBa2Fe3O8 orthorhombic crystal structures, which were based on the earlier reported structural data.40 All of the respective calculation details are provided in the SI (Table S1 and Fig. S1).
The electronic conductivity of the GdBa2Ca2Fe5O13 compound was computed with the linearized Boltzmann's transport equation under the constant relaxation time (CRT) approximation using Boltztrap2 software.53 The input data on band structure were obtained from the self-consistent calculation of G-AFM ordered GBCFO unit cell with the 14 × 14 × 2 k-point sampling of its BZ (the partial occupancies of orbitals were determined within the tetrahedron method using Blöchl corrections54). The electronic conductivity tensor (
CRT) was estimated using the following equation:
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
| Magnetic ordering | Formation enthalpy, eV | |||
|---|---|---|---|---|
| Non-mixed | Model I | Model II | Model III | |
| G-AFM | −1.597 | −1.510 | −1.552 | −1.520 |
| A-AFM | −0.820 | −0.729 | −0.776 | −0.745 |
| FM | −0.339 | −0.254 | −0.291 | −0.268 |
Experimental data are a valuable tool for validating the results of Ueff corrected GGA (GGA+U) calculations. Indeed, the theoretical unit cell parameters of GdBa2Ca2Fe5O13 obtained during structural relaxation (within the G-AFM magnetic ordering) compared with those experimentally determined by X-ray diffraction in the synthesized oxide23 are in reasonable agreement (Table 2). The very close matching of the GBCFO unit cell volumes estimated by GGA+U and Rietveld methods is especially remarkable; the respective relative difference is less than 0.01%. The computed energy-volume curve of the G-AFM ordered GBCFO additionally supports the computing results – the equilibrium volume at 0 K determined by fitting the respective points with the Birch–Murnaghan equation of state56 is extremely close to that obtained by simple structural relaxation (see Fig. S3(a) in the SI). In addition, the extracted temperature dependence of GBCFO unit cell volume appears to be in good agreement with the experimental observations, Fig. S3(b);57 moreover, the obtained Debye temperature (∼430 K at room temperature) coincides well with the known data for similar oxide systems.58 In contrast, the use of the SCAN functional in conjunction with the Ueff term (SCAN+U) brings worse results – the difference between experimental and computed unit cell volumes of GBCFO is estimated to be ∼5%. Therefore, one may assume that the GGA+U approximation is more precise in terms of estimating the unit cell dimensions. Hence, GGA+U approximation was used in this study.
| Parameters | Method | ||
|---|---|---|---|
| Experiment | GGA+U | SCAN+U | |
| a, Å | 5.5222 | 5.53056 | 5.49541 |
| b, Å | 5.5625 | 5.58393 | 5.44159 |
| c, Å | 38.309 | 38.106 | 37.375 |
| α, deg | 90 | 90 | 90 |
| β, deg | 90 | 89.991 | 90 |
| γ, deg | 90 | 90 | 90.006 |
| V, Å3 | 1176.75 | 1176.79 | 1117.66 |
Despite the perfect coincidence between GGA+U computed and the experimental unit cell volume of the GBCFO, the respective values of individual unit cell parameters differ significantly although the inconsistencies lie within the typical range of ab initio errors.59 In addition, the structural relaxation of the GBCFO lattice has led to symmetry breaking, thus transforming the initially orthorhombic unit cell with the Imma space group into the monoclinic P21/n one. The application of SCAN+U approximation has also resulted in the symmetry lowering of the GBCFO lattice, which suggests insufficient precision of structural models used in the calculations (i.e. one does not account for antisite-defects). However, the respective angle deviations from 90° were found to be less than 0.01 in both cases, which implies minor distortion of the GBCFO crystallographic planes.
Although estimation of unit cell dimensions is an important probing for the computational approach, it does not give information on the chemical bonds in the solid. Therefore, it is desirable to compare the calculated bond lengths with those derived from the Rietveld refinement of the structure from the corresponding diffraction data. Since these data are not available in the case of the GdBa2Ca2Fe5O13 in the literature, experimental bond lengths of the isostructural compounds Y1.1Ba1.6Ca2.3Fe5O13 and TbBa2Ca2Fe5O13 will be used in this work.31,45 The respective bond lengths and angle distributions calculated using the GGA+U method for GBCFO are given in Table 3 (the atomic coordinates obtained can be found in Table S2 in the SI). The overall agreement between the experimental bonding distances and angles, and the DFT+U computed values for the GBCFO is satisfactory. The most noticeable difference in the two datasets presented is linked with the geometry of FeO4 tetrahedra: ab initio calculations predict more symmetric distribution of Fe–O bond lengths, both in the apical and equatorial positions of the oxygen-atoms. Contrary to that, experimental data of YBCFO and TBCFO show different equatorial Fe–O distances. Still, the degree of mismatch does not exceed 1% in most cases, thus suggesting an appropriate level of theoretical approximations used. In addition, it should be mentioned that the derived bulk modulus, B0, of the GBCFO (Fig. S3(a)) agrees with the one measured for Fe3+ containing ferrites that simultaneously contain FeO6 octahedra and FeO4 tetrahedra in the structure with ∼1.8 Å and ∼2.0 Å Fe–O bond lengths respectively.60 When only FeO6 octahedra are present, the B0 magnitude is usually higher.61 Finally, it must be stated that the in-plane Fe–O–Fe bond angles for the different oxygen-polyhedra around the Fe3+-cations in RBCFO are significantly different, which is well reproduced using the GGA+U method. This observation suggests that the degree of overlap between Fe-orbitals and O-orbitals differs depending on the coordination number of the Fe3+ cations. Hence, individual contributions from the respective 3d electronic states to DOS might vary depending on the type of Fe3+–O polyhedron, as has already been reported for the YBCFO.45
| Bond distances (Å) and angles (deg.) | Experimental | GGA+U | |
|---|---|---|---|
| R = Y | R = Tb | R = Gd | |
| FeO6 | |||
| d Fe–O apical | 2.1089 | 2.1332 | 2.1187 |
| 2.0824 | 2.032 | 2.1187 | |
| d Fe–O equatorial | 1.9462 | 1.9559 | 1.9640 |
| 1.9462 | 1.9559 | 1.9640 | |
| 1.9684 | 1.9667 | 1.9870 | |
| 1.9684 | 1.9667 | 1.9870 | |
| FeO5 | |||
| d Fe–O apical | 1.8718 | 1.9234 | 1.8610 |
| d Fe–O equatorial | 2.0127 | 1.9958 | 2.0201 |
| 2.0127 | 1.9958 | 2.0201 | |
| 2.0031 | 1.9915 | 2.0170 | |
| 2.0031 | 1.9915 | 2.0170 | |
| FeO4 | |||
| d Fe–O apical | 1.8686 | 1.7532 | 1.8410 |
| 1.8686 | 1.7532 | 1.8410 | |
| d Fe–O equatorial | 1.8556 | 1.8572 | 1.9230 |
| 1.9425 | 2.249 | 1.9230 | |
| d Fe–Fe (R layer), Å | 3.7691 | 3.724 | 3.7692 |
| ΘFe–O–Fe (FeO6) | 168.94 | 166.77 | 171.66 |
| ΘFe–O–Fe (FeO5) | 149.34 | 155.58 | 151.41 |
| ΘFe–O–Fe (FeO4) | 125.06 | 117.88 | 123.96 |
The computed DOS spectrum under the GGA+U approximation of the GBCFO oxide is presented in Fig. 2(a). The results indicate that the studied compound is a narrow-gap semiconductor with a band gap value of Eg = 0.53 eV. The corresponding estimation of partial contributions to DOS reveal typical particularities for Fe3+ containing perovskites – the severely spin-split Fe 3d band, part of which is forming the bottom of the conduction band (CB) and a wide O 2p band providing the top of the VB.64–67 One can also acknowledge the overlying Gd 4d states (∼4 eV above the bottom of the CB) which are too high in energy to participate in the electron conduction process. Accordingly, the electronic structure of the GBCFO, in general, represents the common picture of localized electrons in a transition metal oxide.49
However, the implication of the layered crystalline architecture causes some peculiarities to appear in the DOS spectrum. Namely, the distribution of electronic states for Fe and O atoms was shown to depend on the Fe3+-environment in the unit cell, in particular on the exact type of oxygen-polyhedron around the Fe3+ cations. The mentioned data are shown in more detail in Fig. 2(b) and (c) for Fe 3d and O 2p states, respectively. Considering the information presented in Fig. 2(b) one can notice the wide low-intensity Fe band ranging from −6 eV up to the Fermi level; these are the Fe 3d–O 2p hybridized states. At the same time, peak-shaped Fe 3d states located both in the VB and in the CB are split with the lower energy peak corresponding to pyramidally coordinated Fe atoms. In turn, higher energy states are provided by a combination of electrons bound to the Fe atoms located in the FeO4 tetrahedra and FeO6 octahedra. Given the large splitting energy (>1 eV), it is concluded that the FeO5 layers are mainly responsible for the formation of the CB in GBCFO. On the other hand, the pDOS of the O 2p states in equatorial oxygen positions of the different polyhedron (i.e. FeO4 tetrahedron, FeO5 square pyramid or FeO6 octahedron) shows that the top of the VB is also spatially specific (Fig. 2(c)). Namely, only those states that belong to equatorial oxygen atoms located in FeO6 octahedra do contribute to the top of the VB. The states of oxygen ions in FeO4 and FeO5 polyhedra appear to be ∼1 eV lower in energy and hence their influence on electrical transport properties can be neglected. Accordingly, GBCFO may be considered as a 2D-semiconductor with FeO5-layers acting as the CB, FeO6-layers as the VB and FeO4 ones as effectively insulating channels. This result seems to be controversial, as a previous report in the isostructural YBCFO suggested the top of VB is in the FeO4-layer.45 Still, the use of other functionals (i.e. SCAN) in this work gives similar distribution of the electronic states.
It is important to notice that the computed band gap value (0.53 eV) for the GBCFO is unusually small in comparison with the other Fe3+-perovskites.60,68,69 Considering the relationship between electric bandgap and optical bandgap, we have calculated the optical spectrum of the G-AFM GBCFO, see Fig. S5 in the SI. The color of the solid can be estimated theoretically from its reflectance spectrum.70 The one of the GBCFO (Fig. S5) shows that its reflectance is approximately constant in the visible range; the respectively estimated CIE-color71 is dark brown in agreement with the experimental color of the studied oxide,30 although the determined band gap suggests the material should be totally black.70 This counterintuitive result originates from the peculiar band dispersion in GBCFO – the calculated energy band diagram (Fig. S6 in the SI) indicates that the band gap in the studied compound is indirect in nature and hence the relative contribution of the electronic transition to the overall light absorption should be small. In contrast, the strongest impact on optical absorption is provided by direct transitions, which are higher in energy (above 2 eV). Consequently, the intense increase of optical absorbance in GBCFO is observed only at UV wavelengths, which explains the non-black color of the material considered.
Although the DOS spectra provide important information about the electronic structure, they do not allow to estimate the spatial localization of bands, which is of special importance for semiconductors. Fig. 3 displays the calculated electron charge densities in the CB and VB projected onto specific crystallographic planes of the GBCFO structure. The spatial distribution of charge within the CB and VB is significantly different. The charge distribution in the CB is revealed to be essentially localized on Fe atoms located in the FeO5 square pyramids, while in the VB the charge is more delocalized with comparable contributions of the Fe atoms and equatorial oxygens of the FeO6 octahedra. These findings fully coincide with the conclusions derived from the DOS/pDOS spectra (Fig. 2). Importantly, both bands show pronounced anisotropy of the charge distribution – they appear to be concentrated mainly in the quasi-2D layers perpendicular to the c-axis ([001]p direction), see Fig. S7 in the SI, in agreement with the evaluated band dispersions (Fig. S6). The analysis of different projections of charge density in the CB, Fig. 3(b) and (c), indicates the conducting states are mainly formed by d3z2−r2 Fe orbitals. The VB shows the Fe 3d–O 2p hybridized behavior of the dx2−y2 Fe states, Fig. 3(d) and (e). It should be stated that the derived picture of charge distribution in conjunction with the observed band curvatures (Fig. S6) suggests that the electronic transport in the CB should be more localized than in the VB in agreement with earlier studies in other Fe3+-perovskites.37,72,73
The unique coexistence of the three different types of Fe3+-environment in the crystal structure makes the RBCFO (R = Y, Gd, Tb) oxides an interesting playground for testing the influence of the oxygen coordination of Fe3+ cations on their complete electronic behavior. In this sense, the NPD experiments on TBCFO revealed different magnetic moments localized on Fe3+ ions (μFe) depending on their oxygen coordination.31 The calculated data for GBCFO and the experimental values determined for TBCFO are compared in Table 4. Regarding the individual magnetic moments, GGA+U calculations give correct (within the 6% error limit) values for Fe atoms located in the FeO5 square pyramids and FeO6 octahedra but fail to predict the significant decrease of the μFe in FeO4 tetrahedra. It is important to notice that a similar issue was detected for the Ca2Fe2O5 brownmillerite where the experimentally observed difference between the μFe in FeO4 and FeO6 polyhedra74 was not properly reproduced by DFT+U calculations.66 Nevertheless, the detailed analysis of the pDOS spectra of GBCFO has shown that the electronic states in FeO4 layers do not contribute to VB and CB edges and hence the improper μFe values cannot be considered as substantial incorrectness of the calculation method.
| Parameters (eV) | Crystallographic environment | |||
|---|---|---|---|---|
| Tot | FeO6 | FeO5 | FeO4 | |
| a The value is estimated from the temperature dependence of the GBCFO conductivity.23 b The band gaps are obtained for the compounds with only one type of FeOx polyhedron. | ||||
| |μFe|, μB | 0.0 (0.0) | 4.09 (3.9) | 4.04 (3.8) | 4.05 (3.0) |
| E g | 0.53 (0.38a) | 1.90 (2.4b) | 1.67 (1.35b) | 2.79 (3.02b) |
| U dd | +9.05 | +9.16 | +8.7 | +9.05 |
| Δ CT | −4.08 | −3.83 | −3.73 | −3.63 |
| Δ ox | +2.93 | +2.93 | +3.41 | +3.23 |
As follows from the data presented in Table 4, the partial Eg values are significantly higher (if compared to the total Eg) for each polyhedron type but Eg is especially large in the case of the FeO4 tetrahedra (∼2.8 eV), assuming these sites form insulating layers in the GBCFO lattice. Notably, a very close band gap is obtained for CaBaFe4O8 that contains entirely FeO4 structural units in its lattice, see Fig. S8(a); this result is also supported by the respective optical measurements (Fig. S8(b)). In addition, the computed Eg values for GdFeO3 and BaCaFe2O5 oxides (Fig. S9 in the SI) that contain only FeO6 and FeO5 structural units respectively, were found to be in qualitative coincidence with the data obtained for individual FeO6 and FeO5 layers in GBCFO, see Table 4. Considering all the above results, one can suppose that the combination of FeO4, FeO5 and FeO6 polyhedra in one lattice provides a synergetic effect on the resulting electronic band structure.
To support this conclusion, it is important to study the implications in the band structure due to the elimination of one of the types of the FeOx polyhedra. In this context, another compound of the A3m+5nFe3m+5nO8m+13n series was considered, the Gd1.2Ba1.2Ca0.6Fe3O8, which contains only FeO4 tetrahedra and FeO6 octahedra.30 The calculation was made for the perfectly ordered GdBaCaFe3O8. The GGA+U computed DOS/pDOS spectra for the GdBaCaFe3O8 oxide are presented in Fig. 4(a) (the corresponding representation of the unit cell is given in the inset30). In this case, the Fe-3d levels in the CB and the VB show similar positioning on the energy scale, disregarding the actual FeOx polyhedron type, which differs from the spectrum of the GdBa2Ca2Fe5O13. In addition, the band gap of GdBaCaFe3O8 is estimated to be ∼2.2 eV, thus falling in the typical range of Eg values of Fe3+-perovskites.68,69
Another possibility is the combination of either FeO5 square pyramids and FeO6 polyhedra, or FeO5 pyramids and FeO4 tetrahedra. These combinations containing only Fe3+ ions can be modeled by GdBa2Fe3O8 and NaCa2Fe3O7 structures, respectively. Although such compounds have not been reported experimentally, the corresponding unit cells were built based on the known structures of YSr2Cu2FeO8 and YSr2Cu2FeO7,40 which resemble in a certain sense the GBCFO lattice. The computed DOS/pDOS spectra of GdBa2Fe3O8 and NaCa2Fe3O7 are presented in Fig. 4(b) and (c), respectively. As shown, the copresence of FeO6 and FeO5 polyhedra results in the lowest band gap of ∼1.1 eV; importantly, this result corroborates with the noticeable splitting of Fe 3d states in the CB of GdBa2Fe3O8 among the respective polyhedra, Fig. 4(b). In turn, the combination of FeO4 tetrahedra and FeO5 pyramids provides a band gap of ∼1.6 eV, Fig. 4(c), which is in between of Eg values obtained for GdBa2Fe3O8 and GdBaCaFe3O8. Interestingly, bringing two different types of FeOx polyhedra in one lattice allows for certain Eg reduction, if compared to the case of crystals that have only octahedra (Eg = 2.41 eV), only square pyramids (Eg = 1.35 eV) or only tetrahedra (Eg = 3.02 eV). Still, none of the double combinations provide as much band gap contraction as is observed in the GBCFO case. Therefore, it seems that the co-presence of all three FeOx structural units discussed is indeed required to obtain a narrow band gap in Fe3+ containing perovskite-like compounds. The schematical depiction of the observed peculiarity is shown in Fig. 5.
![]() | ||
| Fig. 5 Graphic representation of the influence of different combinations of FeOx polyhedra on the band gap of perovskite-related lattices. | ||
The obtained DOS of GBCFO (Fig. 2) allows us to elucidate other important band structure parameters, also collected in Table 4: Udd, the magnitude of splitting of the Fe 3d states; ΔCT, the so-called charge transfer energy; and Δox, the average amount of energy required to transfer electrons from the O-2p band to the top of the VB.49 The magnitude of these parameters is schematically shown in Fig. 2(a). Determination of the respective band centers, which can be done using multiple ways, must be carried out to compute Udd, ΔCT and Δox. In this work, it was assumed that the band center of the Fe-related states should not include those located in a wide Fe 3d–O 2p hybridized band; in turn, a similar approach was utilized for the O-2p states that contribute to Fe-3d peaks placed deep in the VB and CB. The calculated values suggest that the studied compound is a negative charge transfer oxide, which is characteristic of Fe3+-perovskite-type oxides.49 The high magnitude of Udd energy (>9 eV) provides strong evidence that under oxidation of GBCFO, the increased charge will be accumulated on oxygen sites.47 In turn, the large Δox parameter (∼3 eV) implies high energy required for oxygen vacancy formation in GBCFO, which is indeed experimentally observed.30,44 Therefore, one can acknowledge the sufficient precision of GGA+U approximation utilized in this work, capable of reproducing the key experimental findings reported previously.
Considering the essential differences in the computed pDOS spectra between the alternating layers with different Fe-environments in GBCFO, individual estimation of Udd, ΔCT and Δox for these layers makes certain sense. The respective data for FeO4-, FeO5- and FeO6-containing layers are given in Table 4 (importantly, the discussed estimation of band structure parameters did not involve apical oxygen sites but only equatorial ones). The charge transfer energy remains essentially the same disregarding the polyhedron type, which means that the electrons transferred from any Fe site in GBCFO will be immediately replaced (with approximately the same speed) by other electrons from the O-2p formed VB. However, the magnitude of the Udd energy is slightly lower for the Fe atoms in the FeO5 square pyramids. Apparently, this observation may be responsible for the revealed particularity of CB formation in the GBCFO. In accord with that, the lowest Δox value is calculated for FeO6-containing layers, thus explaining why these d-orbitals form the top of the VB.
A certain alignment with the previously discussed data can be found in the LOBSTER computed partial Crystal Orbital Hamilton Population (pCOHP) curves, which basically show the bonding (if pCOHP < 0), non-bonding (pCOHP = 0) or antibonding (when pCOHP > 0) characteristic of the chemical interactions.52 The respectively estimated pCOHP curves for the Fe–O bonds in the different oxygen coordination polyhedra of GBCFO lattice are shown in Fig. 6. The results indicate that chemical interaction between Fe and oxygen atoms exhibits a strong bonding characteristic deep in the VB (<−6 eV) disregarding the type of polyhedron, apical or equatorial allocation of the oxygen ions and Fe3+ spin state. This conclusion agrees with previous findings in BaFeO3.49 Similarly, pCOHP values become positive in the CB of the GBCFO for all considered cases, which marks the anti-bonding characteristic of the interactions and hence the excited nature of these states. Interestingly, the bonding type between Fe-3d and O-2p electrons in the upper part of the VB (i.e. in the energy range from −6 to 0 eV) is found to be spin-dependent: spin-up electrons of Fe atoms form bonding type interactions, while the inverse process is observed for spin-down electrons. As discussed previously,49 this observation comes from the inversion of the respective contributions of 3d and 2p orbitals to the Fe–O chemical bond depending on the spin state of the 3d electrons. In this sense, the type of oxygen polyhedron around the Fe atoms should not change that picture, which is indeed observed in the case of GBCFO. However, the shape and relative position on the energy scale of a certain pCOHP are shown to be highly dependent on the Fe-coordination. For instance, the anti-bonding Fe–O states in the CB, that belong to FeO5 square pyramids are located much closer to the Fermi level in comparison with the FeO4 or FeO6 cases (Fig. 6(b)). The anti-bonding interactions between Fe and O atoms observed at the top of the VB correspond to the FeO6 octahedra (Fig. 6(a)). Finally, in the case of FeO4 tetrahedra, the characteristic of chemical interaction appears to be non-bonding in the vicinity of the Fermi level. All the above results suggest that the excited electronic states in the GBCFO tend to localize either in FeO5 layers (when considering electrons in the CB) or in FeO6 ones (when implying electron holes in VB). At the same time, those states located in FeO4 layers can be considered as effectively insulating (Fig. 6(c)) in complete agreement with the conclusions derived from the DOS data.
Essential information about the chemical bond strength can be obtained by the integration up to the Fermi level of the pCOHP curves (IpCOHP). These data quantitatively measure the energy of covalent bonds in eV units;52 it can often be used as a guideline for searching the most probable defect configurations in non-stoichiometric oxides.49,75 The respective results for Fe–O bonds in GBCFO are presented in Fig. 6(d). All computed IpCOHP values are negative, which marks the essential chemical stability of Fe3+-perovskites. However, the magnitude of |IpCOHP| differs drastically depending on the Fe-coordination. In particular, the magnitude of IpCOHP for the apical oxygens and Fe atoms located in FeO6 octahedra indicates low bond strength. Contrary to that, in the case of other FeOx polyhedra, the respective quantity significantly increases, thus suggesting that apical oxygen ions are much more strongly bonded to FeO4 tetrahedra and FeO5 square pyramids. It should be noted that this conclusion nicely coincides with the GBCFO bond lengths presented in Table 3 – the shorter the bond, the stronger the bonding and vice versa. Accordingly, one can expect that apical oxygen sites of the FeO5-units are unlikely to become vacant. In contrast, the equatorial positions have other distribution of IpCOHP energies with the highest one belonging to Fe–O bonds located in FeO5 square pyramids, suggesting that these positions are the most probable sites for oxygen vacancy formation. However, one should consider that the largest (among other polyhedra) Δox value in FeO5 (see Table 4) implies higher energies to transfer electrons from the hybridized Fe 3d–O 2p band to the Fermi level. Hence, one may expect the decreased |IpCOHP| value in the case of FeO5 pyramids as evidence of Fe–O bond extension induced by the absence of apical oxygen atoms in the Gd layer. Accordingly, the most probable place for oxygen vacancy formation in GBCFO is the equatorial O sites in FeO6 octahedra.
el as:![]() | (5) |
el tensor component can be determined. However, the available experimental data correspond to a polycrystalline sample, which makes a direct comparison with eqn (5) difficult. To overcome this problem one can assume that the GBCFO grains are small enough to be randomly oriented in the sample studied. Then, the experimental conductivity σexp can be expressed as a normalized trace of
el which yields the following simple result:![]() | (6) |
CRT〉 and measured σexp30 becomes possible; the respective data are given in Fig. 7(a). It is worth mentioning that the CRT approach used in this work does not allow us to properly estimate the magnitude of relaxation time τ. However, the previously published results suggest this parameter lies within the range of few dozens of femtoseconds (10−15 s).76 Therefore, one can assume τ = 10 fs for the GBCFO, which agrees with the independently evaluated τ for the donor-doped SrTiO3 perovskite.73
Prior to discussing the peculiarities of transport phenomena in the studied material it is instructive to provide a brief description of its defect structure; the respective information allows to elucidate the concentration of charge carriers in the VB and CB, which is of primary importance for electronic conductivity calculations. Given that GBCFO is a narrow band-gap semiconductor it is necessary to account for different extrinsic processes of defect formation, as well as for the intrinsic thermal ionization. For instance, one can suppose that oxygen incorporation can occur into the different structural vacancies in the GBCFO lattice, i.e. into FeO5 or FeO4 layers, thus producing interstitial sites Oi. Another possibility is that oxygen vacancies VO can be formed – these can be located in all FeOx polyhedra. The list of possible sites for hosting defects, as well as the corresponding defect formation reactions are given in the SI. The results of GGA+U calculations of the j-th defect formation enthalpies ΔHdf(j) in GBCFO oxide are presented in Table 5.
| Defect formation process | ΔHdf(j), eV | |
|---|---|---|
| VO formation | FeO6 | 3.057 |
| FeO5 | 3.287 | |
| FeO4 | 3.253 | |
| Oi formation | FeO5 | 0.654 |
| FeO4 | 1.202 | |
As can be seen, oxygen vacancies are very unlikely to form with the respective enthalpies computed to be higher than 3 eV. Still, these results align well with the experimental data77 and can be explained by high energy penalties to convert Fe3+ into Fe2+ in perovskite-oxides. In addition, the predicted trend of FeO6 octahedra to the sites which are most susceptible to hosting vacancies (the Δox parameter was found to be the lowest for FeO6 polyhedra, see Table 4) seems to be supported by actual calculations, Table 5. On the other hand, the formation of interstitial oxygen appears to be intricated, although the relative fraction of Oi defects (especially of those formed in FeO5 layers) will be higher than that of oxygen vacancies under oxidative conditions (i.e. in air atmosphere). Still, as follows from the calculated results, the overall deviation of oxygen content in GdBa2Ca2Fe5O13±δ from 13 should be small (|δ| will be less than 0.01 at 1000 K and under an air atmosphere). The available thermogravimetric data support this conclusion – the measured mass change of GBCFO when heating in air was found to be negligible.57 What is even more interesting is that the positive sign of ΔHdf of the Oi formation process implies that the fraction of interstitial oxygen will increase with temperature. Accordingly, the number of holes in the VB will grow with heating. Given that ΔHdf(Oi) < ΔHdf(VO) and ΔHdf(Oi) > 0, one can reasonably assume electron holes will be the main charge carriers in GBCFO under SOFC cathode conditions.
In agreement with the conclusions of the previous paragraph, the determination of the 〈
CRT〉 value was made assuming acceptor doping level (pd) to be of 1017–1019 cm−3 in magnitude, which roughly corresponds to δ ∈ [−10−5, −10−3] per GdBa2Ca2Fe5O13 formula unit. The estimated conductivity curves at pd < 1018 cm−3 are shown in Fig. 7(a) and reveal the transition from the extrinsic to the intrinsic regime as the temperature increases. One can acknowledge the obtained 〈
CRT〉 numerical values at pd = 1017 cm−3 are close to those (σexp) measured in the experiment. Notably, when the doping amount is higher (i.e. pd = 1019 cm−3), the shape of the conductivity plot becomes almost temperature independent. These results are in agreement with the experimental findings. Indeed, defect formation enthalpies suggest that the concentration of holes should increase with temperature, i.e. pd = f(T). And hence, the experimentally observed σ = f(T) of GBCFO originates from the intrinsic regime at low temperatures (due to the narrow band gap) and hole-doped regime, which is invoked by the incorporation of interstitial oxygen into the crystal lattice, at higher temperatures. Importantly, this mechanism qualitatively explains the flattening of the σexp = f(T) curve at T > 800 K, see Fig. 7(a).
Accounting for the discussed crystalline anisotropy of GBCFO it is interesting to consider the direction-dependent components of 〈
CRT〉. The respective data represented as σCRT/τ functions are given in Fig. 7(b). Those components of conductivity tensor that are parallel to either [110]p or [−110]p crystallographic directions (i.e. planes (001), also named planes ab) are 3–4 orders of magnitude higher than those parallel to the [001]p direction. In addition, the σCRT‖[110]p and σCRT‖[−110]p values are almost similar. Therefore, both magnitudes will be referred to as σCRT(‖), while the quantity σCRT‖[001]p will be designated as σCRT(⊥). This denomination underlines that the obtained result shows direct evidence of the anisotropy of the GBCFO electrical properties. However, the reason behind the uncovered phenomenon seems to be unclear. In this sense, the previous computational study for the YBCFO suggested that the absence of oxygen ions in the Y layer provided the main contribution to the 2D characteristic of electronic conduction.44 In turn, the present study provides a more complicated picture where all polyhedral layers in the GBCFO lattice play a certain role in providing the anisotropic behavior. But indeed, the elimination of the FeO5 square pyramids should result in a more isotropic 〈
CRT〉; the respective evaluations made for GdBaCaFe3O8 and presented as a ratio between σCRT(‖) and σCRT(⊥) components of conductivity tensor clearly confirm this statement, see Fig. 8. In fact, the corresponding pDOS data suggest that the CB of GdBaCaFe3O8 is composed from electronic states of both FeO6 and FeO4 structural units (Fig. 4(a)). Therefore, it can be deduced that in the case of the acceptor doped GBCFO the main reason for the anisotropic behavior of 〈
CRT〉 is the combination of FeO6 layers forming the top of the VB and the large slabs composed of FeO65 and FeO4 polyhedra that provide effective isolation of the conductive octahedral layers from each other. Because of that the curvature of the VB along the [001]p direction in GBCFO (b2 direction in reciprocal space, as represented in Fig. S6) is almost zero, resulting in large effective masses for charge carriers and hence making the [001]p direction highly unfavorable for electronic conduction in GBCFO. Similar behavior is observed for the acceptor doped GdBa2Fe3O8 and NaCa2Fe3O7 compounds (Fig. 8); the only difference from GBCFO is the nature of insulating layer - it is either FeO4 tetrahedra (in NaCa2Fe3O7) or FeO5 pyramids (in GdBa2Fe3O8).
| This journal is © the Owner Societies 2025 |