Laura Vanessa Parra-Mesaa,
Críspulo E. Deluque Torob,
David A. Landínez Téllezac,
Henry M. Ortiz Salamancad and
Jairo Roa-Rojas*ac
aGrupo de Física de Nuevos Materiales, Departamento de Física, Universidad Nacional de Colombia, 111321 Bogotá DC, Colombia. E-mail: jroar@unal.edu.co
bGrupo de Nuevos Materiales, Facultad de Ingeniería, Universidad del Magdalena, Santa Marta, Colombia
cGrupo de Estudios de Materiales GEMA, Departamento de Física, Universidad Nacional de Colombia, 111321 Bogotá DC, Colombia
dGrupo de Investigación en Ciencias de Materiales y Minerales, Universidad Distrital Francisco José de Caldas, Bogotá DC, Colombia
First published on 31st July 2025
The synthesis process and the characterisation of structural, morphological, compositional, optical, electrical and magnetic properties of the Ca2TiRuO6 material are described and analysed. Rietveld refinement of X-ray diffraction data showed the crystallisation of the material into a perovskite-like structure, given by the P21/n space group. Scanning electron microscopy reveals images of granular, porous surface morphology with submicrometre-sized grains. Semi-quantitative analysis of energy dispersive spectroscopy does not reveal the presence of compositional impurities. Diffuse reflectance spectroscopy analyses show a direct gap semiconductor-type optical bandgap Eg = 0.89 eV. Electrical measurements corroborate the semiconducting nature of the material with evidence of two types of transport mechanisms, variable range hopping at low temperatures and small polaron hopping at high temperatures. The I–V curves conform to varistor-type behaviour. The magnetic susceptibility as a function of temperature evidences irreversible behaviour characteristic of magnetically disordered systems and the magnetisation suggests antiferromagnetic behaviour with weak ferromagnetic hysteresis due to the canting of the spin moments of Ru4+ cations because of octahedral distortions in the crystal cell. Band structure and density of electronic states calculations exhibit half-metallic behaviour, caused by a shift of the 4d-Ru4+ orbitals from the valence band into the conduction band, crossing the Fermi level for the spin-down configuration, while the spin-up polarisation remains semiconducting. The half-metallic character is attributed to the strong spin–orbit coupling effects caused by the dequenching of the orbital angular momentum of 4d-Ru4+.
The final macroscopic physical properties also depend on the characteristic structural distortions of the perovskite, which can be predicted using several criteria, the first of which is the so-called tolerance factor, which for double perovskites is given by4
![]() | (1) |
![]() | (2) |
Based on the above, in the present manuscript the experimental and theoretical study of the material Ca2TiRuO6 is proposed, for which there is no information in the crystallographic and material bases, as a starting point to show the validity of these criteria in the design of new materials with specific physical properties. The ionic radius of Ca2+ in cuboctahedral coordination with the oxygen anions is 1.340 Å, and the ionic radii of Ti4+ and Ru4+ in octahedral coordination are 0.605 Å and 0.620 Å, respectively. These oxidation states are assumed because Ca has no other valence states, and the most stable state of Ti is +4, so Ru should also assume the 4+ state to electronically balance the system as the valence state of oxygen is 2−. Thus, the expected value of the tolerance factor for Ca2TiRuO6 should be τ = 0.9733. Meanwhile, the bonds affect the valences, so it is necessary to consider the bond valence sum around each ion, so that the tolerance factor is renamed the bond valence parameter and its final value is τ = 0.9425, showing that the system is more distorted than expected, moving away from the ideal τ = 1 value of an exactly cubic system with the octahedra TiO6 and RuO6 perfectly aligned (no tilts or rotations). The value of the global instability index calculated for a cubic structure (space group Fmm) is GII = 0.36368, suggesting that the material cannot synthesise in such a symmetric structure, leading to an expected cell with differences between lattice parameters and possible octahedral distortions. The most likely structures, with GII values closest to zero, occur for rhombohedral R
, monoclinic P21/n and tetragonal I2/m and I4/m space groups. Regarding the Jahn–Teller effect, it is necessary to mention that the electronic configuration of Ru4+ is [Ar] 3d4, so, because the crystal field splitting energy is very high, the 4 valence electrons are expected to remain in their low spin state, for which the Hund rules predict a configuration t42g(3↑, 1↓), whereby the effective magnetic moment should be 2.0μB. Meanwhile, when Ru4+ is octahedrally coordinated with oxygens in perovskite-type materials, such as SrRuO3, it has reported experimental values of around 1.2μB for bulk samples,11 suggesting that interactions with the oxygens could extend the density of states close to the Fermi level, giving rise to ferromagnetic-type responses.12 Density of states and band structure calculations will elucidate these effects in Ca2TiRuO6 in the present work. To establish the type of octahedral distortion occurring in this material, as well as its space group and crystallographic structure of the material, X-ray diffraction characterization of samples produced in the laboratory is necessary.
The aim of the present work is to corroborate that these criteria facilitate the determination of the synthesis possibility of the Ca2TiRuO6 material, analysing its structural, optical, magnetic and electronic properties, both from an experimental and theoretical point of view, by means of density functional theory (DFT) calculations.
The best refinement fit was for a double perovskite-type structure belonging to space group P21/n (#14), corresponding to a monoclinic cell, which is characterised by exhibiting screw axes in the primitive unit cell. In this structure, a double rotation of degree n takes place around the axis, followed by a translation along the same axis, plus a translation (degree 1), with respect to the value of the lattice parameter. This result is consistent with the low value of the global instability index, which for space group P21/n is GII = 0.00258. It is important to note that in the perovskite milieu, there is often confusion about the difference between a 50% partially substituted perovskite, for example CaTi0.5Ru0.5O3, and a so-called double perovskite, which in this case is Ca2TiRuO6. Although from a compositional point of view these two materials are identical, they are not structurally identical, since in the first case the Ti and Ru cations are randomly distributed in the crystalline cell, while in the second case they adopt a cationic arrangement along the crystallographic axes, forming what is called a superstructure.
The highly ordered distribution of Ti and Ru cations results in the presence of reflection planes in the [111] and [331] directions, as shown in Fig. 2a. These planes do not produce constructive interference in the case of random arrangement of these cations, although other peaks such as (133) are also a kind of fingerprint of the presence of superstructure. As expected from the reflection conditions for the space group P21/n, as shown in Fig. 1, the diffracting planes (h0l) are such that h + l is even, and (00l) occurs only for l = even. Similarly, no forbidden reflections (0k0) with odd k due to the presence of the double screw axis, nor those due to the presence of a slip plane (h0l) with odd h + l are observed.20
![]() | ||
Fig. 2 (a) Diffraction planes due to cationic ordering in the perovskite cell of Ca2TiRuO6. (b) Structural scheme of the material belonging to space group P21/n. |
A first interesting result concerns the monoclinic angle which, although extremely small, being only 12 ten-thousandths away from 90°, is a characteristic of the #14 space group (P21/n), which differentiates it from the orthorhombic space group #62 (Pnma and Pbnm), which is typical of partially substituted single perovskites. Several results in Table 1 suggest a strong distortional character in the structure of the Ca2TiRuO6 double perovskite.21 First, the non-symmetric positions of the ions, away from the exact 0, 1/4, 1/2 and 3/4 positions that are expected in cubic perovskites, suggest distortional distributions of the TiO6 and RuO6 octahedra. Another circumstance that favours the distorting aspects are the differences in interatomic distances, particularly in the Ti–O versus Ru–O bonds and, particularly, between the Ca cations with respect to the oxygen anions in their cuboctahedral co-ordinations. Additionally, the presence of angles other than 180° in the Ti–O–Ru bonds not only contributes to the structural distortion but can also include modifications in the magnetic characteristics of the material due to the possibility of canting of the electronic spins in the orbitals of the Ru4+ atoms, as will be discussed in the magnetic response section.
Atom | Wyckoff site | Atomic coordinates (±0.0001) | P21/n space group cell parameters | ||
---|---|---|---|---|---|
x | y | Z | |||
Ca | 4e | 0.512 | 0.5375 | 0.25 | a = 5.3871(1) Å |
Ti | 2c | 0 | 0.5 | 0 | b = 5.5171(0) Å |
Ru | 2d | 0.0344 | 0.25 | 0.9105 | c = 7.7076(3) Å |
O1 | 4e | 0.1939 | 0.2138 | −0.0391 | |
O2 | 4e | 0.288 | 0.6976 | −0.0391 | α = β = 90.0000° |
O3 | 4e | 0.405 | −0.0174 | 0.2511 | γ = 89.9988° |
Bond lengths (±0.0001 Å) | CaO cuboctahedric bond lengths (±0.0001 Å) | ||
---|---|---|---|
Ti–O: 1.9651 Å | 3.2869 | 2.7098 | 3.0621 |
2.7088 | 3.2874 | 2.3527 | |
Ru–O: 1.9814 Å | 2.6021 | 2.3859 | 3.0688 |
2.3873 | 2.6014 | 2.5341 |
Octahedral tilt (±0.01°): ρ = 15.22° and η = 15.09 | Octahedral distortion Glazer notation |
Bond angles (±0.01°): Ti–O–Ru: 155.35° | a−b+a− |
Reliability factors: χ2 = 1.994, R(F2) = 4.88%, Rp = 1.90% |
The structural parameters obtained from the refinement are shown in Table 1, from which the structure schematized in Fig. 2b was constructed. Finally, the occurrence of the angles ρ and η denoting inclinations in the TiO6 and RuO6 octahedra is associated with the Glazer notation22 a−b+a−, which can be clearly seen in Fig. 2b, where along the b-axis the octahedra are tilted in phase while, in the a- and c-axis directions, they are tilted out of phase.
In Fig. 3b, the Lα, Kα and Kβ represent the electronic transitions L–M, K–L and K–M orbitals for each of the atomic elements that make up the Ca2TiRuO6 perovskite-like material. The area under the curve of the spectrum allows the mass percentage of the atoms present in the compound to be obtained. Thus, the deconvolution of the curve made it possible to establish the experimental weight percentage shown in the inset of Fig. 3b for each of the chemical elements contained in the material. This value is approximately 1.5% higher than the theoretical value expected for the cations from the stoichiometric formula of the material Ca2TiRuO6. Meanwhile, for oxygen, the experimental weight percentage is slightly lower because oxygen is a large radius anion, so the possibility of scattering of incident electrons with the few 2p electrons of oxygen is lower than in the case of cations.
Γ = 7Ag + 7B1g + 5B1g + 5B2g + 8Au + 7B1u + 9B2u + 9B3u, | (3) |
![]() | ||
Fig. 4 (a) Diffuse reflectance spectrum obtained for the Ca2TiRuO6 double perovskite and (b) analysis for the determination of the optical bandgap. |
The spectrum analysis for the determination of the optical bandgap presented in Fig. 4b is carried out by means of the Kubelka–Munk equation,24
![]() | (4) |
For bulk samples, where energy gap equals the energy absorbed when αR = lnR, where R is the reflectivity measured with respect to the unit, it has been shown that the formulation of eqn (4) can be expressed in terms of the maximum and minimum reflectance values, as follows25
![]() | (5) |
The curve in this figure shows an exponential type decrease in resistivity with increasing temperature. Furthermore, the order of magnitude of the resistivity is in the range of 105 mΩ cm, suggesting a tendency of the material to adopt semiconducting feature. The analysis of this curve is carried out by means of the general equation for the transport of electric charge in non-conducting materials28
![]() | (6) |
An exponent ξ = 1 corresponds to nearest-neighbour hopping and recombination processes, ξ = 1/4 represents the Mott's variable-range hopping (VRH) conduction model, expected in strongly disordered semiconductor systems with localised charge-carrier states, while ξ = 1/2 describes the small polaron hopping, also known as Efros–Shklovskii hopping (ESH), which considers a so-called Coulomb gap, with the occurrence of a small jump in the density of states near the Fermi level caused by interactions between localised electrons.29 Fig. 6a shows an anomaly in the curve in the range 235 K < T < 270 K, which has to do with the separation between two regimes with well-defined transport mechanisms. At T < 235 K, the decrease in resistivity with increasing T is fast, fitting the VRH model, whereas at T > 270 K, where the change in resistivity with increasing T is slow, the appropriate fit is ESH, as seen in the inset of Fig. 6a. For the VRH model, a value Ea = 0.08 ± 0.01 eV was obtained, while for ESH, it was Ea = 0.20 ± 0.01 eV, indicating that the carriers in intragap states are very close to the Fermi level (at 0.08 eV) for the VRH model, justifying a higher rate of change of resistivity with increasing temperature. In the 235 K < T < 270 temperature region, the system enters a depletion zone until the charge carriers further away from the Fermi level (at 0.20 eV) are obtained, registering a smaller resistivity variation with heating of the semiconductor system.
Fig. 6b shows the current density curve as a function of the applied electric field. The response is non-ohmic, following a J = CEφ type behaviour, which characterises certain semiconductor materials,30 where the exponent φ determines the non-ohmic character through a parabolic type trend and C is a characteristic non linear coefficient of the material. The value of φ = 1.45 obtained is close to that predicted by the quasi-hydrodynamic semiconductor equations, which is φ = 1.5, and is expected to be related to the occurrence of physical regimes of the sample with evanescent electric charge carriers.31 This behaviour can be attributed to the presence of natural barriers opposing the flow of the transport current, such as grain boundaries,32 which can be imagined as Schottky barrier-type micro-junctions. Thus, the J–E curve for two grains with a boundary between them resembles an assembly of two consecutive Zener diodes. Then, the resistances due to intergranular boundaries at room temperature provide a non-linear response, as shown in Fig. 6b, which was previously characterised as varistor-like. As observed in the SEM images in Fig. 3, the grains are multiform and the contact surfaces are not always identical. Meanwhile, three-dimensional modelling of the grains in this type of ceramic material in the literature has been simulated in the form of Voronoi-type structures,33 such that the microstructural electrical transport is idealised in the form of an equivalent electrical circuit with resistive and capacitive elements associated with the intergranular and intragranular currents flowing through the material, with non-linear results such as those reported in the present work.34
First, the difference in the behaviour of the magnetic moments when the sample is cooled in the absence of a magnetic field and the field is applied to the frozen system during measurement while increasing the temperature (ZFC procedure), compared to the measurement while decreasing the temperature with field applied (FC procedure), is evident. From this difference, the occurrence of irreversibility is determined with temperature values 96 K for H = 50 Oe, 90 K for H = 100 Oe and 88 K for H = 200 Oe. Both the occurrence of irreversibility and the decrease of the respective irreversibility temperature with the increasing applied field are characteristic of spin-glass type systems, where magnetic disorder produces a clustered distribution of magnetic moments with low exchange energy, causing the difference in magnetic response between ZFC and FC measurements.35–37
Since the structural characterization (Section 4.1) evidences the formation of a superstructure with cationic ordering, the occurrence of magnetic frustration as the cause of the spin glass effect in Ca2TiRuO6 can be ruled out. Meanwhile, the geometrical arrangement of the lattice, with clear structural distortions, may give rise to a non-trivial configuration of the magnetic moments, such that the magnetic interactions are not optimized to favour a single ordered state, resulting in magnetic disorder.38 In addition, it is observed that for all three field values applied at room temperature, there is evidence of finite magnetic susceptibility. Since the curves do not follow a 1/T behaviour, it could be stated that the material is magnetised below an ordering temperature whose value exceeds the maximum measurement temperature (room temperature).
On the other hand, Fig. 6b reveals the occurrence of very weak magnetic hysteresis (see inset), plus an apparently linear unsaturated trend up to applied fields of 30 kOe. The linear trend of the curve can be associated with antiferromagnetic-type response of the spin moments of Ru4+ cations, which are located in octahedral RuO6 coordinations of the unit cell of the Ca2TiRuO6 material. Meanwhile, the octahedral distortions mentioned in Section 4.1 can give rise to spin momentum canting effects, which introduce weak orientations resulting from ferromagnetic-type magnetisation, thus evidencing the weak ferromagnetic response manifested as a small hysteresis in the curve of Fig. 6b.39
An interesting result in Fig. 7 has to do with the shape of the bands which tends to be symmetrical when comparing the two spin polarisations. Meanwhile, the valence states near the Fermi level responsible for the semiconducting character for the spin-up configuration undergo a shift towards the Fermi level for the spin-down polarisation, making the system conductive for this spin orientation.
On the other hand, transition metal oxide materials exhibit strong electron–electron correlations, as the behaviour of an electron is strongly influenced by the presence and behaviour of other electrons in the same or nearby atoms. For this reason, Hubbard U corrections in DFT calculations are crucial to accurately describe the electronic and magnetic properties of transition metals, especially in materials where electron–electron interactions are significant. These corrections address the self-interaction error present in standard DFT functionals, which tend to overestimate the delocalisation of d electrons in transition metals. When the Hubbard U-correction is introduced (Fig. 7b), although the half-metallic type distribution is maintained, there is a significant change in the semiconductor bandgap value, which is magnified to Eg = 1.56 eV. This result agrees with the experimentally measured value, in which there is no identification of the two spin polarisations, the average between the two being measured.
In order to examine closely the specific contributions of the electronic orbitals near the Fermi level, calculations of density of states were carried out for the up and down spin configurations, as shown in Fig. 8, for both GGA and GGA+U calculations.
In the figure, the partial states due to the Ca2+ orbitals are not presented because they do not show contributions in the vicinity of the Fermi level. The total density of states exemplified in Fig. 8a agrees with the band structure, clearly revealing the above-mentioned half-metallic behaviour for both calculations GGA and GGA+U. From the density of partial states shown in Fig. 8b–d, it can be inferred that the semiconducting nature comes from hybridisations between Ti4+ cations with O2− anions, which take place in the TiO6 octahedra, while the conducting characteristic is due to hybridisations between Ru4+ cations and O2− anions in the RuO6 octahedra.
For both calculations, the difference between states with different spin polarisations allowed obtaining the effective magnetic moment per unit cell, whose integer value 2μB corroborates the appearance of the half-metallic behaviour because this is catalogued as one of its essential characteristics.46 X-ray absorption spectroscopy (XAS) measurements on bulk samples of SrRuO3 reveal that the crystal field splitting energy for Ru4+ reaches ∼5 eV.47 The spin electronic moments are expected to adopt a low spin configuration in the 4d orbitals, such that the angular momentum is zero (l = 0) and the spin momentum s = 1, so that the total angular momentum is J = 1, so that with g = 1, the effective magnetic moment is expected to be . The model applied for the calculation considers the octahedral-coordinated environment of the 4d-Ru4+ cations with the 2p-O2− anions, considering the complete degeneracy of the 4d and 2p orbitals, as well as the electrostatic and exchange interactions at the transition metal sites, without spin and orbital restriction because the translational symmetry of the 4d orbitals is included. Three decades ago, the relevance of the interplay between orbital and spin ordering on the magnetic properties was shown in perovskite-type 3d transition metal oxides with partially filled t2g orbitals.48 Another interesting aspect is that in perovskites, the orbital ordering states are strongly affected by strong Jahn–Teller distortions and structural distortions.49
When comparing the total and partial densities of states between the two calculation procedures, with GGA and GGA+U, it is observed that the inclusion of the correction potential U causes a shift of the electronic orbitals towards states further away from the Fermi level, both in the valence band and in the conduction band, which consequently expands the bandgap value from 1.30 eV to 1.56 eV, with an average between up and down cannels of 0.78 eV, which is relatively close to the experimental value of 0.89 eV.
As indicated by the arrows in Fig. 8, this can be deduced from the distribution of 3d states of Ti4+, which is relevant in the conduction band, at 1.0 eV above the Fermi level, in a regime where, in addition, there are 2p electronic states of O2−. On the other hand, the 4d valence states of Ru4+ correlate with the 2p valence states of O2− for the spin-up orientation, while for spin-down polarisation, there are 4d states of Ru4+ and 2p states of O2− crossing the Fermi level. This shift is evidence for the occurrence of orbital angular momentum dequenching in Ca2TiRuO6, which introduces the strong spin–orbit coupling (SOC) in this system, which, in turn, increases the value of the effective magnetic moment.
In order to evaluate the influence of the SOC on the electronic states in the vicinity of the Fermi level, GGA-PBE and GGA+U-PBE calculations were performed, the results of which are shown in Fig. 9.
![]() | ||
Fig. 9 Total density of electronic states calculated by GGA-PBE (a) and GGA+U-PBE (b) close to the Fermi level for the double perovskite Ca2TiRuO6 with the inclusion of spin–orbit coupling. |
It is important to note that the incorporation of SOC in the DFT calculations with the VASP 6.3.2 software for the study of the electronic properties of the Ca2TiRuO6 material eliminates the distinction between the spin channels. This is because the inclusion of the SOC produces a mixture of electronic states, which are described by combinations of the orbital quantum numbers (l, m), eliminating the explicit spin separation. Consequently, the total density of states analysis is performed in terms of the total contribution of each spherical harmonic orbital. These results reveal the effect of the SOC interaction on the electronic properties of the material, which, in addition, are substantially modified with the presence of the Hubbard potential, going from an essentially conducting state (Fig. 9a) to a semi-metal-like state (Fig. 9b) with the appearance of a bandgap of 0.4 eV. Under this analysis, the resulting strong SOC may have given rise to the half-metallic nature exhibited by this material. It is known that SOC can play an important role in the emergence of half-metallicity in perovskite materials, particularly in double perovskites containing transition elements. For example, this effect is observed in the material Ba2FeReO6, where strong correlations in the d electrons, combined with the SOC, are essential to reach a half-metallic state in the material.50
CCDC 2445463 contains the supplementary crystallographic data for this paper.51
This journal is © The Royal Society of Chemistry 2025 |