DOI:
10.1039/D5RA06712J
(Paper)
RSC Adv., 2025,
15, 44116-44124
Exploring lead free Rb2AlInX6 halide double perovskites for advanced energy harvesting applications
Received
6th September 2025
, Accepted 4th November 2025
First published on 12th November 2025
Abstract
Halide double perovskites have recently attracted attention as stable and environmentally benign alternatives to lead based perovskites for optoelectronic and energy applications. However, detailed insights into their stability, electronic structure, and multifunctional properties remain limited. In this study, the physical properties of Rb2AlInX6 (X = Cl, Br) were systematically examined by first-principles calculations. The structural stability of both compounds was confirmed through formation enthalpy, tolerance factor (τG), octahedral factor (µ), and octahedral misfit (Δµ), all of which fall within the accepted stability ranges. Both Rb2AlInCl6 and Rb2AlInBr6 crystallize in the cubic Fm3m phase with optimized lattice constants of 20.37 and 21.43 bohr, respectively. Electronic structure analysis identifies both Rb2AlInCl6 and Rb2AlInBr6 as semiconducting with calculated bandgaps of 2.85 eV and 1.90 eV, respectively, underscoring their potential for optoelectronic applications. Mechanical stability, verified via Born criteria, was further supported by elastic tensor analysis, demonstrating isotropic and robust mechanical behavior. The Rb2AlInBr6 exhibits moderate absorption extending into the visible region, while the Rb2AlInCl6 primarily absorbs in the near UV. These features suggest potential for optoelectronic or UV photodetection applications. While thermoelectric analysis shows notable power factor values at 800 K, pointing toward possible thermoelectric applications. These findings provide a comprehensive understanding of Rb2AlInX6 halides, offering valuable insights into their multifunctional prospects in next generation optoelectronic and thermoelectric devices.
Introduction
Scientists are finding novel materials that may be used to address the technical issues of the present time as technology advances, such as rechargeable battery packs that may be utilized at room temperature or above, and high stress bearing ceramics for the aviation industry.1–3 Owing to their remarkable potential and versatility, halide double perovskites are a major topic of interest for material scientists, who have made amazing strides in developing novel materials. The formula for HDPs is A2BB′X6, where X is a halogen ion, B and B′ represent ions from magnetic or other trivalent metals, and A stands for alkali or alkaline earth compounds (or related).4,5 Additionally, several studies have been provided to grasp the electronic, magnetic, and morphological features of HDPs. Rb2AgAlX6 (X = Br, I) demonstrated bandgap (Eg) of 2.60 eV and 1.08 eV, which specifies that they may absorb light energy from the UV to VIS spectrum. Their potential for incorporation into cutting-edge solar energy systems is further highlighted by the absorption band of Rb2AgAsI6, from 1.7 eV to 3.4 eV.6 The direct Eg of K2InSbCl6 (1.31 eV) and K2InSbBr6 (1.22 eV) enables efficient absorption in the UV-vis regions, which makes these perfect for solar uses. Also, their excellent transport properties suggest strong potential for thermoelectric (TE) applications at 300 K.7 Additionally, it has been effectively forecast that many stable and ecologically friendly HDPs will be viable and sustainable substitutes for solar power and thermoelectric techniques, including Rb2YInX6 (X = Cl, Br, I),8 Cs2AgInX6 (X = F, Cl, Br, I),9 and Cs2InAgCl6.10 Cs2AgBi(Br, Cl)6 have indirect Eg of 2.19 and 2.77 eV, and show air stability with minimal degradation. These environmentally friendly semiconductors offer a suitable perovskite for sustainable devices.11 Several other related HDPs like Cs2ErXCl6 (X = Ag, Au).12 X2ScHgCl6 (X = Cs, Rb)13 and A2YHgCl6 (A = Cs, K)14 were also suggested as sustainable applications.
The lack of previous research on Rb2AlInX6 (X = Cl, Br) offers a chance to examine their physical characteristics. This study aims to provide valuable details into the fundamental properties of both compounds and their potential applications in advanced technologies through an in-depth analysis. Such pioneering work not only expands the knowledge base of halide double perovskites but also sets the stage for their integration into future technological innovations.
Computational methods
DFT calculations were applied to investigate the electronic structure of the Rb2AlInX6 (X = Cl, Br).15,16 To guarantee accurate Eg analyses, the modified Becke–Johnson (mBJ) was used, as shown by the equation below17| |
 | (1) |
It employs the electron density to determine the electronic structure. These are solved using the FP-LAPW technique.18 The wave functions within each muffin-tin sphere are expanded using particular parameters in this technique. The RMT and Kmax products are set to 8 to guarantee accurate representation. A Monkhorst–Pack grid equivalent to 1000 k-points in the full Brillouin zone was employed to ensure accurate sampling. A convergence threshold of 0.00001 Ry for charge density was applied during self-consistent field (SCF) iterations, which ensured total energy convergence within 1 × 10−5 Ry. Energy-volume determinations were used to optimize the structure using the Murnaghan equation:19
| |
 | (2) |
We examined the Eg dependent optical properties using the Kramers–Kronig relations. To calculate the muffin-tin radius (RMT), two conditions had to be met: (i) the MT spheres had to be free of core charge leakage, and (ii) there had to be no overlapping between the spheres. For Rb, Al, In, Cl, and Br, the RMT values were 2.5, 2.15, 2.5, 2.1, and 2.38 bohr to ensure that there was no current loss. The elastic constants were computed separately using the CASTEP module in the Materials Studio package, employing a 4 × 4 × 4 Monkhorst–Pack k-point grid. Additionally, TE These properties were calculated using the BoltzTraP code,20 which employs the rigid band approximation (RBA) and the constant relaxation time approximation (CRTA). In the RBA, the effect of temperature and carrier concentration on the band structure is neglected, which may slightly affect the accuracy of Seebeck coefficient predictions at high temperatures.21 Similarly, the CRTA assumes a constant carrier scattering time (τ) independent of temperature and energy, whereas in halide perovskites, strong electron–phonon coupling can lead to τ variations that influence both conductivity and thermopower.22–24 These approximations, though widely used for qualitative trend analysis, may limit the precision of absolute transport coefficients. Although these approximations simplify the transport description, they remain reasonable and widely used for a first-order estimation of carrier transport.
Results and discussion
Structural features
The structural analysis reveals that Rb2AlInX6 (X = Cl, Br) crystallizes in a cubic structure with space group Fm3m (No. 225). The structure of Rb2AlInX6 (X = Cl, Br) is illustrated in Fig. 1 with the atomic sites of Rb, Al, In, and X at (0.75,0.25,0.25), (0,0,0), (1/2,0,0), and (0.75,0,0), correspondingly. The enhanced lattice constants for Rb2AlInCl6 and Rb2AlInBr6 are determined to be 20.37 and 21.43 bohr, correspondingly. The structures optimization is illustrated in Fig. 2 and relaxation properties are displayed in Table 1 include lattice constants (Å), optimized bulk modulus (B), its derivative (Bp), ground state energy E0 (Ry), and volume.
 |
| | Fig. 1 Crystal structure of Rb2AlInX6 (X = Cl, Br). | |
 |
| | Fig. 2 Optimization in non-magnetic phase of (a) Rb2AlInCl6 and (b) Rb2AlInBr6. | |
Table 1 Calculated values of lattice constant (Å), B (GPa), Bp (GPa), volume (a.u),3 ΔH and ground state energies E0 (Ry) of stable state of cubic Rb2AlInX6 (X = Cl, Br)
| Parameters |
Rb2AlInCl6 |
Rb2AlInBr6 |
Rb2InSbCl6 (ref. 25) |
Rb2InSbBr6 (ref. 25) |
| Lattice constant (Å) |
10.780 |
11.342 |
11.23 |
11.78 |
| B (GPa) |
26.170 |
23.7137 |
23.11 |
19.69 |
| Bp (GPa) |
5.0 |
5.0 |
4.66 |
4.66 |
| V (a.u.)3 |
2113.4876 |
2461.7013 |
— |
— |
| E0 (Ry) |
−29717.339 |
−55458.88 |
— |
— |
| τG |
0.99 |
0.98 |
0.98 |
0.96 |
| ΔH (eV) |
−1.953 |
−3.247 |
−1.62 |
−1.34 |
A stable crystal structure is achieved by minimizing lattice strain and ensuring ideal ionic packing through a well-balanced tolerance factor. The distortion of the metal halide octahedra is also affected by octahedral misfit values, which can improve defect tolerance and adjust electronic band alignment. When constructing improved HDPs, these structural benefits are crucial since they greatly improve the material performance in optoelectronic applications.26,27 By applying the relation given below to calculate the formation enthalpy (ΔH), the thermodynamic integrity of both HDPs is verified.
| | |
ΔH = ERb2AlInX6 − 2ERb − EAl − EIn − 6EX
| (3) |
ΔH needs to be negative to be thermodynamically stable. The calculated ΔH values for Rb2AlInCl6 are −1.953 and −3.247 eV for Rb2AlInBr6, signifying that both are robust and unlikely to degrade under typical conditions. Recent investigations on Cs2InAsX6 (X = Cl, Br) have shown that both are stable and exhibit negative ΔH values of −2.20 and −3.64 eV.28 Furthermore, X2LiSbI6 (X = K, Cs) compounds have been reported to be dynamically stable, exhibiting negative formation energies of −3.88 eV and −3.95 eV, respectively.29
The stability factors like τG, µ, and Δµ are calculated to determine the stability of Rb2AlInCl6 and Rb2AlInBr6.26,27 These are determined as:
| |
 | (4) |
| |
 | (5) |
| |
 | (6) |
These are determined using ionic radii values of Rb, Al, In, and Cl/Br shown as RRb, RAl, and the mean of RCl and RBr. τG range for stable perovskites is 0.8 to 1, with unity values indicating perfect structure. The determined values of this factor are 0.99 and 0.98 for Rb2AlInCl6 and Rb2AlInBr6, confirming their stability. To confirm the validity of these results, a comparison with analogous compounds Cs2XCeI6 (X = Li, Na) from Murtaza et al.30 showed τ values of 0.87 and 0.85 for Cs2LiCeI6 and Cs2NaCeI6. The successful synthesis of several related perovskites including Cs2InBiCl6, Cs2InBiBr6, Cs2InBiI6,31 Cs2ScAgI6,32 Cs2NaLaCl6,33 Cs2YAuBr6,34 K2InBiBr6,35 and Cs2LiCeF6 (ref. 36) further supports their structural stability in agreement with the Goldschmidt tolerance factor model. The Δµ values are computed as 0.13 and 0.12 for Cl and Br perovskites, showing their stability as both are closer to the null value.37 Moreover, the stability of these perovskites is also validated from µ values, which are calculated as 0.43 and 0.40 and exist within the standard stability range of 0.4 to 0.9 as shown in Fig. 3.37,38
 |
| | Fig. 3 Graphical representation of stability Parameters of Rb2AlInX6 (X = Cl, Br). | |
Electronic properties
To comprehend the electronic and optical characteristics Rb2AlInCl6 and Rb2AlInBr6, the electronic band structures (BS) have been examined in conjunction with their structural characteristics. By combining information about crystal structure and electronic properties, BS provides information about the conduction and basic electronic nature of halides.39 mBJ potentials were used to calculate the BS for Rb2AlInCl6 and Rb2AlInBr6. The indirect Eg values of 2.85 eV for Rb2AlInCl6 and 1.90 eV for Rb2AlInBr6 are calculated. As the atomic radius increases from Cl to Br, this trend shows a drop in Eg, as shown in Fig. 4. To further check the relativistic effects, spin orbit coupling (SOC) was incorporated into the mBJ calculations. The inclusion of SOC slightly reduced the Eg values to 1.82 eV for Rb2AlInCl6 and 1.18 eV for Eg, without altering the indirect nature of the Eg. The overall band dispersion and density of states (DOS) features remained qualitatively similar, confirming that SOC has only a minor influence on the electronic structure of these compounds. The corresponding SOC band structure and DOS plots are provided in the SI (Fig. S1). To assess the accuracy of our Eg calculations, we compared them with experimental Eg of well-known double perovskites. For instance, Cs2BiAgCl6 displayed a Eg of 2.2 eV, while Cs2AgBiBr6 showed a Eg of 1.95 eV.40,41 Our predicted values are in agreement with experimental data, with small discrepancies. This comparison confirms the reliability of our Eg predictions, providing quantitative error bars for the Eg. These values are consistent with those observed in the literature and remain applicable for further exploration of related thermoelectric properties.
 |
| | Fig. 4 Band structure of (a) Rb2AlInCl6 and (b) Rb2AlInBr6. | |
The total density of state (TDOS) and partial density of states (PDOS) for Rb2AlInCl6 and Rb2AlInBr6 reveal important insights into their electronic structures. In the valence band (VB) region, the In-5p orbitals show a notable contribution, indicating their strong bonding interactions with the halide atoms. For Rb2AlInBr6, the Br-4p orbitals dominate the VB and display strong hybridization with the In-5p orbitals, highlighting robust bonding. Similarly, in Rb2AlInCl6, the Cl-3p orbitals show a critical role in the VB, with prominent hybridization peaks aligning with the In-5p orbitals. Rb-5s states show negligible contribution, consistent with their non-bonding nature. Al-3p states contribute weakly near the Fermi level, indicating a minor influence on the electronic structure. In the conduction band, contributions from the In-5p orbitals dominate, along with significant involvement from Br4p in Rb2AlInBr6 and Cl-3p in Rb2AlInCl6, indicating their crucial role in optical excitation and conductivity.
The TDOS and PDOS profile of Rb2AlInCl6 and Rb2AlInBr6 is shown in Fig. 5 and 6. The changing of Br with Cl leads to a narrower valence band is observed for Rb2AlInCl6 due to the smaller ionic radius of chlorine, which strengthens bonding and pushes the valence states to slightly deeper energies, whereas the larger ionic radius of bromine in Rb2AlInBr6 leads to a broader valence band with enhanced hybridization effects. Both compounds exhibit semiconducting behaviour, with the CB dominated by In-5p and halide p-orbitals, underlining their potential for optoelectronic uses.
 |
| | Fig. 5 Density of states (TDOS) profile of (a) Rb2AlInCl6 and (b) Rb2AlInBr6. | |
 |
| | Fig. 6 Partial density of states (PDOS) profile of (a) Rb2AlInCl6 and (b) Rb2AlInBr6. | |
Optical properties
A primary consideration in determining whether solar cells are feasible for energy generation is their efficiency. The material optical characteristics, which control how it interacts with incoming electromagnetic (EM) radiation, are directly related to it.42,43 The dielectric function ε(ω) is an intricate function that controls the relationship between a photovoltaic material and incoming EM light. It is important in this context.44–46 The real ε1(ω) and imaginary ε2(ω) components of the ε(ω) can be used to obtain additional optical characteristics that further influence the material's solar power capacity. These characteristics include the absorption coefficient α(ω), refractive index n(ω), reflectivity R(ω), extinction coefficient k(ω), optical conductivity σ(ω), and energy loss function L(ω). Scattering information is provided by ε1(ω), while absorption attributes are provided by ε2(ω).47,48 The ε1(ω) describes the degree of photon scattering and the transmission speed, which is dependent on the largest light dispersion. The static ε1(0) values for Rb2AlInCl6 and Rb2AlInBr6 are 3.15 and 3.91, respectively (Fig. 7(a)). It's interesting to note that this validation of Penn's model shows that Eg and ε1(0) have the opposite connection49 as shown in eqn (7):| |
 | (7) |
 |
| | Fig. 7 Optical parameters (a) real ε1(ω) component, (b) imaginary ε2(ω) component, (c) absorption coefficient α(ω) and (d) optical conductivity σ(ω) for Rb2AlInX6 (X = Cl, Br). | |
A higher ε1(0) indicates a stronger interaction with the electromagnetic field and enhanced polarization under an external electric field, suggesting improved photon–electron coupling efficiency. For photodetectors and solar cells uses, a larger ε1(0) value is associated with increased dielectric screening and better charge separation.50,51 Therefore, the relatively higher ε1(0) of Rb2AlInBr6 implies stronger light–matter interaction in the visible region, while the smaller ε1(0) of Rb2AlInCl6 points to faster photon transmission and suitability for UV and high frequency optoelectronic applications.
Before displaying peaks, the ε2(ω) spectra for Rb2AlInCl6 and Rb2AlInBr6 first show a fluctuating pattern with first peaks at 3.55 eV and 2.70 eV. Some researchers calculated static ε1(0) for the similar compounds. ε1(0) values of 7.00, 3.80, 2.5, and 3.2, respectively were calculated for Cs2ScInBr6, Cs2ScCuCl6, and Cs2ScCuF6,52 Cs2AuInCl6,53 whereas Cs2LiMoX6 (X = Cl, I)54 are 4.84 and 3.2037. Na2AuInCl6, Na2AuInBr6, and Na2AuInI6, it is calculated as 2.15, 2.34, and 3.75, respectively.55 The ε2(ω) component represents the optical absorption process, revealing how efficiently the material can absorb incident photons and promote interband electronic transitions. The ε2(ω) exhibits peaks at 8.63 eV for Rb2AlInCl6 and 7.64 eV for Rb2AlInBr6, producing identical results (Fig. 7(b)). The ε1(ω) trend across related perovskites indicates that Cs2InSbCl6 attains the highest value at 0.88 eV, with Cs2InSbBr6 and Cs2InSbI6 exhibiting increasing responses at 1.43 eV (5.9) and 2.87 eV (8.7), respectively, reflecting the influence of halide substitution on optical behavior.56 The sharp peaks in ε2(ω) confirm direct allowed transitions in both compounds, essential for visible and UV optoelectronics. However, the lower transition energy (2.70 eV) in Rb2AlInBr6 falls directly within the visible-light region, confirming its potential as a photoactive and emissive material for visible light photodetectors and LEDs.57,58 In contrast, Rb2AlInCl6 shows its main transitions at higher energies (3.55 eV and above), suggesting it could serve effectively in UV photodetectors, transparent window layers, or protective coatings in optoelectronic devices.59 The following equation can be used to compute it:60
| |
 | (8) |
The high α(ω) in the visible and near UV regions61 demonstrates strong optical activity and efficient photon utilization.62 Such high absorption is a key requirement for photoactive absorber layers in solar cells and photodetectors, as it enables efficient electron–hole generation even in thin films.63,64 Both Rb2AlInCl6 and Rb2AlInBr6 show considerable values throughout a wide energy range when the α(ω). Both Rb2AlInCl6 and Rb2AlInBr6 exhibit significant optical transitions at 8.68 eV and 7.90 eV, respectively, in Fig. 7(c) corresponding to the deep UV region. In addition, the Rb2AlInBr6 compound shows additional weaker absorption features at lower energies, indicating a modest extension of optical activity into the visible region. This suggests that while both systems are primarily UV absorbers, the Rb2AlInBr6 possesses relatively enhanced visible light response, making it more promising for optoelectronic or photocatalytic applications. The movement of photoelectrically produced photons inside the substance is analysed by the σ(ω) and calculated as:
| |
 | (9) |
The disruption of connections can be explained by the presence of strong EM radiation. σ(ω) exhibits free carriers produced upon capturing electromagnetic radiation, and that σ(ω) and α(ω) are intimately associated. A maximum σ(ω) reflects efficient carrier transport and reduced recombination losses, both vital for high performance optoelectronic devices.65,66 The σ(ω) peaks for Rb2AlInBr6 and Rb2AlInCl6 reach 5286 and 5534 Ω−1 cm−1 at 7.64 eV and 8.59 eV, respectively, indicating strong light induced carrier generation as shown in Fig. 7(d). The superior σ(ω) in Rb2AlInBr6 at lower photon energies suggests its carriers can be effectively excited under visible illumination, while Rb2AlInCl6, responding mainly to higher energy photons, is suitable for UV or high energy optoelectronics.67 Together, their strong σ(ω) and α(ω) responses confirm that both compounds can function as efficient charge transport and photon conversion layers in multi spectral optoelectronic systems.68 While our DFT-based calculations provide a reliable estimate of the fundamental Eg, they do not account for excitonic effects, which are known to be significant in double perovskites with large exciton binding energies on the order of a few hundred m eV.69 These effects may result in a lower optical gap compared to the DFT calculated fundamental gap.70 However, our calculations still provide valuable insight into the material's electronic structure, and the predicted Eg remain applicable for general analysis of the material's properties. The combination of suitable Eg, high absorption, and strong conductivity indicates that Rb2AlInBr6 is more efficient for visible light optoelectronic devices such as solar absorbers, LEDs, and photodetectors, while Rb2AlInCl6, with its wider gap, can play a complementary role in UV optoelectronics and as a transparent or electron blocking layer in heterojunction structures. Further investigations incorporating excitonic effects, such as many body perturbation theory (e.g., GW approximation), would enhance the accuracy of optical absorption predictions.71,72 k(ω) is another optical characteristic that is strongly associated with α(ω). The degree of damping of input photons k(0) in the alloys studied is k(ω), which is brought on by both dispersion and captivation. Importantly, because they are connected by Kramers–Kronig relations, it is similar to the ε2(ω).73,74 The most significant values for Rb2AlInCl6and Rb2AlInBr6 emerge at 8.70 eV and 7.76 eV in the fluctuating pattern of the k(ω) spectrum (Fig. 8(a)). The overall lower extinction coefficient of Rb2AlInBr6 in the visible range minimizes optical damping and photon loss, which is beneficial for light emitting and absorbing devices, whereas the higher k(ω) in Rb2AlInCl6 contributes to its efficiency in UV photon absorption and filtering. A key indicator of the proportion of EM radiation reflecting at a particular energy is the R(ω), which is displayed in Fig. 8(b).75,76 The R(ω) values were notable, first and foremost; for ω = 0, they were 0.08 and 0.10 for Rb2AlInCl6 and Rb2AlInBr6. As energy rises, R(ω) exhibits a fluctuating trend. R(ω) is calculated using the provided relation, and the greatest reflectance for Rb2AlInCl6 is shown at 8.68 eV and 7.81 eV for Rb2AlInBr6. R(ω) can be computed using the relation given below:
| |
 | (10) |
 |
| | Fig. 8 Optical features of Rb2AlInX6 (X = Cl, Br). (a) Extinction coefficient k(ω), (b) reflectivity R(ω), (c) refractive index n(ω) and (d) energy loss function L(ω). | |
Moderate reflection at low energies and increasing reflection at higher photon energies indicate balanced optical behavior for both compounds. Rb2AlInBr6, with higher reflectivity in the visible region, is promising for light emitting and laser applications, while Rb2AlInCl6, exhibiting lower reflection and higher transmission, can be applied in anti-reflective or UV transparent coatings.77 The n(ω) varies as does ε1(ω), an important factor for evaluating substance transparency.78 As for Rb2AlInCl6 and Rb2AlInBr6, their respective static n(0) values are 1.78 and 1.97. According to Fig. 8(c), there are other noteworthy n(ω) peaks for Rb2AlInCl6 at 3.34 eV and for Rb2AlInBr6 at 2.53 eV. The HDP analysis revealed nearly identical results. The n(0) values of Cs2NaMoCl6 and Rb2NaMoCl6 were found to be 1.71 and 1.69,79 while Rb2YAuI6 and Cs2YAuI6 demonstrated slightly higher values of 2.01 and 2.03, consistent with their heavier halide composition.80,81 The slightly higher refractive index of Rb2AlInBr6 in the visible range ensures stronger photon confinement and efficient light–matter interaction, ideal for LEDs and photovoltaic absorbers.82,83 Conversely, the lower n(ω) of Rb2AlInCl6 improves transparency and is suitable for UV photonics or as a top layer coating in tandem solar architectures.84 The L(ω) shows the drop in photon energy as it passes through the material. The L(ω) peaks for Rb2AlInCl6 and Rb2AlInBr6 are seen at 9.32 eV and 8.56 eV (Fig. 8(d)). A greater L(ω) value corresponds to more pronounced plasmonic resonance and stronger collective electron oscillations. The lower energy loss of Rb2AlInBr6 indicates less internal damping and greater photon utilization in the visible region, making it favourable for LEDs and photodetectors.85 Meanwhile, Rb2AlInCl6, with a higher L(ω) and wider Eg, is suitable for UV sensing and protective optical devices. Therefore, it can be concluded that Rb2AlInBr6, with its higher absorption coefficient, greater optical conductivity, and favourable refractive behaviour, is the most promising compound for visible light optoelectronic applications, while Rb2AlInCl6 can serve complementary roles in UV detection, high frequency photonics, and as a transparent coating layer.86,87
Thermoelectric features
HDPs exhibit promising TE properties because of their exceptional electronic structures, favourable carrier transport mechanisms, and tunable Eg. Their TE performance is primarily governed by a delicate balance between electrical conductivity (σ/τ), Seebeck coefficient (S), and thermal conductivity (k/τ).88,89 Using the BoltzTraP algorithm,20 the efficacy of Rb2AlInCl6 and Rb2AlInBr6 as TE materials is assessed. Fig. 9 presents the temperature dependent factors governing the thermoelectric performance of Rb2AlInCl6 and Rb2AlInBr6. The electronic components of k/τ are identified by the BoltzTraP code by removing the contributions from holes. A predetermined relaxation period (τ) of 10−14 seconds for TE characteristics is assumed in the computations. Results demonstrate a significant power factor (PF), low k/τ, and high σ/τ, which suggest strong TE efficiency. The quantity and motion of carriers are influenced by the σ/τ. Since higher T offers them the energy they need to shift, conductivity and temperature are directly correlated.42,90 The σ/τ increases nearly linearly with temperature for both compounds, with Rb2AlInCl6 consistently higher than Rb2AlInBr6 across the entire range. At 800 K, σ/τ reaches 2.3 × 1019 Ω−1 m s−1 for Rb2AlInCl6 and 1.5 × 1019 Ω−1 m s−1 Rb2AlInBr6 as depicted in Fig. 9(a). In comparison to similar halide perovskites, this series of compounds demonstrates competitive room-temperature electric conductivity, with recorded values of 0.29 × 1018, 0.21 × 1018, 0.25 × 1018, and 0.24 × 1018 Ω−1 m s−1 for K2TlBiCl6, K2TlBiBr6, Rb2TlBiCl6, and Rb2TlBiBr6, respectively.91
 |
| | Fig. 9 Thermoelectric factors (a) electrical conductivity (σ/τ), (b) thermal conduction (k/τ) (c) Seebeck coefficient (S) and (d) power factor of Rb2AlInX6 (X = Cl, Br). | |
The k/τ value quantifies a material's ability to conduct heat.92,93 k/τ grows constantly with temperature. The k/τ values increase with temperature, with Rb2AlInCl6 showing higher values than Rb2AlInBr6. At 800 K, k/τ reaches 1.05 × 1015 W m−1 K−1 s−1 for Rb2AlInCl6 and 7.5 × 1014 W m−1 K−1 s−1 Rb2AlInBr6 as depicted in Fig. 9(b). Compared to similar materials, K2AlInF6, K2AlInCl6, and K2AlInBr6, when measured at 800 K, not only exhibit Seebeck coefficients of 150, 160, and 135 µV K−1, respectively, but also a significant rise in k/τ to between 4.0–4.75 × 1014 W m−1 K−1 s−1.94
S plays a key role in determining thermoelectric efficiency.95–97 The primary carriers are holes, as shown by a positive reaction in the S spectrum.98 This verifies the p-type nature of Rb2AlInCl6 and Rb2AlInBr6. For Rb2AlInCl6 and Rb2AlInBr6, S decreases across the whole S spectrum (Fig. 9(c)). For Rb2AlInCl6, the maximum value of S is 209 µV K−1 at 200 K, suggesting the TE potential of both HDPs in various uses at lower temperatures. The PF has a major effect on TE performance and can be computed as PF = S2σ/τ.99 For Rb2AlInCl6 and Rb2AlInBr6, the lowest PF at 200 K is 1.35 × 1011 W K−2 m−1 s−1. As the T rises, both HDP materials show a steadily growing trend in the PF plot (Fig. 9(d)). The most notable PF values for Rb2AlInCl6 and Rb2AlInBr6 at 800 K are 1.0 × 1012 W K−2 m−1 s−1 and 6.01 × 1011 W K−2 m−1 s−1. Thermoelectric analysis of related perovskites shows that Cs2ScTiCl6 achieves a PF of 2.6 × 10−2 W m−1 K−2 s−1 at 800 K, closely followed by Cs2YTiCl6 with 2.3 × 10−2 W m−1 K−2 s−1.100 The higher PF values obtained for Rb2InSbCl6 (4.5 × 1010 W m−1 K−2 s−1) and Rb2InSbBr6 (2.5 × 1010 W m−1 K−2 s−1) further confirm their potential as promising thermoelectric materials.100 The outstanding TE characteristics of Rb2AlInCl6 and Rb2AlInBr6 suggest that these compounds would be enormous options for TE generators and cooler devices.
Conclusion
In this study, the structural, electronic, optical, mechanical, and thermoelectric properties of Rb2AlInX6 (X = Cl, Br) halide double perovskites were comprehensively investigated using density functional theory. The electronic and optical properties were computed using the WIEN2k code with the mBJ potential for accurate estimation of Eg, while the elastic constants were calculated using the CASTEP module to validate mechanical stability. Both compounds crystallize in the stable cubic Fm3m phase and exhibit indirect semiconducting Eg of 2.85 eV for Rb2AlInCl6 and 1.90 eV for Rb2AlInBr6. TDOS and PDOS analyses reveal strong hybridization between In-5p and halogen-p orbitals, confirming a mixed covalent-ionic bonding nature responsible for their semiconducting behaviour. The calculated elastic constants satisfy the Born mechanical stability criteria and indicate ductile mechanical behaviour, suggesting high structural integrity under external stress. Optical analysis reveals pronounced absorption in the UV-visible region, strong dielectric response, and high optical conductivity, indicating efficient light–matter interaction suitable for optoelectronic applications. Thermoelectric calculations performed using the BoltzTraP code reveal enhanced power factors and moderate thermal conductivities, resulting in ZT values of 0.78 for Rb2AlInCl6 and 0.70 for Rb2AlInBr6 at 200 K. Collectively, the results demonstrate that Rb2AlInCl6 and Rb2AlInBr6 are mechanically and thermodynamically stable lead-free halide double perovskites, exhibiting excellent prospects for integration into high performance optoelectronic and thermoelectric energy harvesting devices.
Author contributions
S. M. K. A. Naqvi and R. A. Khera conceived and designed the study; S. Abdalla and K. Akhtar carried out the computational modelling and figure preparation; A. Ali and N. Y. Elamin contributed to data analysis and interpretation; S. M. K. A. Naqvi, F. Abbas and A. M. Khan participated in writing and editing the manuscript.
Conflicts of interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Data availability
All data provided and/or analysed during this study were included as figures and tables in this article.
Supplementary Information: additional results, including detailed mechanical properties and band structure/DOS graphs with spin–orbit coupling (SOC) effects. See DOI: https://doi.org/10.1039/d5ra06712j.
Acknowledgements
This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU), grant number IMSIU-DDRSP2503.
References
- Y. Zhao, A. Taheri, M. Karakus, A. Deng and L. Guo, Minerals, 2021, 11, 1107 CrossRef CAS.
- X. Zhang, Y. Tang, F. Zhang and C. Lee, Adv. Energy Mater., 2016, 6(11), 1502588 CrossRef.
- Q. Peng, A. Farrukh, M. Sajid, K. Shafqat, K. Muhammad, A. A. Awadh Bahajjaj, M. Nazar and J. Rehman, J. Phys. Chem. Solids, 2025, 196, 112332 CrossRef CAS.
- S. Ahmad, J. Feng, M. Zakria, S. Hatim Shah, A. Alam, S. Shakeel, D. Muhammad and I. Ullah, Mater. Sci. Eng. B, 2024, 309, 117641 CrossRef CAS.
- S. Choudhary, S. Tomar, D. Kumar, S. Kumar and A. S. Verma, East Eur. J. Phys., 2021, 2021, 74–80 Search PubMed.
- A. Mera, T. Zelai, S. A. Rouf, N. A. Kattan and Q. Mahmood, J. Mater. Res. Technol., 2023, 24, 5588–5597 CrossRef CAS.
- A. Ayyaz, G. Murtaza, A. Azazi, A. Usman, A. A. A. El-Moula, A. K. Alqorashi, F. U. R. Ahmad and M. Touqir, Opt. Quantum Electron., 2024, 56, 1204 CrossRef CAS.
- P. Aymen Nawaz, G. M. Mustafa, S. Sagar Iqbal, N. A. Noor, T. Shahzad Ahmad, A. Mahmood and R. Neffati, Sol. Energy, 2022, 231, 586–592 CrossRef CAS.
- M. Tariq, M. A. Ali, A. Laref and G. Murtaza, Solid State Commun., 2020, 314–315, 113929 CrossRef CAS.
- G. Volonakis, A. A. Haghighirad, R. L. Milot, W. H. Sio, M. R. Filip, B. Wenger, M. B. Johnston, L. M. Herz, H. J. Snaith and F. Giustino, J. Phys. Chem. Lett., 2017, 8, 772–778 CrossRef CAS PubMed.
- E. T. McClure, M. R. Ball, W. Windl and P. M. Woodward, Chem. Mater., 2016, 28, 1348–1354 CrossRef CAS.
- O. Zayed, T. M. Al-Daraghmeh, Nasarullah, S. V. Menon, F. Mahmood, S. Ray, A. Sinha and A. Gautam, J. Inorg. Organomet. Polym., 2025, 1–16 Search PubMed.
- Nasarullah, M. Nazar, S. M. K. Abbas Naqvi, G. A. A. M. Al-Hazmi and Y. M. Alawaideh, J. Phys. Chem. Solids, 2025, 201, 112637 CrossRef CAS.
- K. Riaz, N. Drissi, S. Abdalla, S. Belhachi, R. Bousbih, M. S. Soliman, Nasarullah, T. H. A. Hasanin, M. Nazar and N. A. Abdulhussein, J. Inorg. Organomet. Polym., 2025, 35, 5363–5376 CrossRef CAS.
- Á. Nagy, Phys. Rep., 1998, 298, 1–79 CrossRef.
- N. Argaman and G. Makov, Am. J. Phys., 2000, 68, 69–79 CrossRef CAS.
- F. Tran and P. Blaha, Phys. Rev. Lett., 2009, 102, 226401 CrossRef PubMed.
- P. Blaha, K. Schwarz, P. Sorantin and S. B. Trickey, Comput. Phys. Commun., 1990, 59, 399–415 CrossRef CAS.
- F. Birch, J. Appl. Phys., 1938, 9, 279–288 CrossRef.
- G. K. H. Madsen and D. J. Singh, Comput. Phys. Commun., 2006, 175, 67–71 CrossRef CAS.
- M.-S. Lee and S. D. Mahanti, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 165149 CrossRef.
- Y. Hattori, T. Konoike, S. Uji, Y. Tokumoto, K. Edagawa and T. Terashima, Appl. Phys. Lett., 2024, 125, 083102 CrossRef CAS.
- R. Claes, G. Brunin, M. Giantomassi, G.-M. Rignanese and G. Hautier, Phys. Rev. B, 2022, 106, 094302 CrossRef CAS.
- B. Zhang, J. Klarbring, F. Ji, S. I. Simak, I. A. Abrikosov, F. Gao, G. Y. Rudko, W. M. Chen and I. A. Buyanova, J. Phys. Chem. C, 2023, 127, 1908–1916 CrossRef CAS PubMed.
- D. Behera, B. Mohammed, S. Taieb, B. Mokhtar, S. Al-Qaisi and S. K. Mukherjee, Eur. Phys. J. Plus, 2023, 138, 520 CrossRef CAS.
- A. E. Maughan, A. M. Ganose, M. A. Almaker, D. O. Scanlon and J. R. Neilson, Chem. Mater., 2018, 30, 3909–3919 CrossRef CAS.
- S. Ghosh, H. Shankar and P. Kar, Mater. Adv., 2022, 3, 3742–3765 RSC.
- M. Z. Kazim, N. Raza, S. A. Aldaghfag, A. Dahshan, K. Ahmad, M. Yasar, M. Ishfaq and M. Yaseen, J. Phys. Chem. Solids, 2024, 189, 111954 CrossRef CAS.
- M. Akhtar, F. K. Alanazi, M. Sajid, S. Abdalla, S. Belhachi, Nasarullah, M. Nazar, S. Murtaza, S. M. K. A. Naqvi, A. K. Alqorashi, Y. M. Alawaideh, M. Faizan and M. S. Ahmad, J. Inorg. Organomet. Polym., 2025, 35, 4810–4824 CrossRef CAS.
- H. Murtaza, Q. Ain, J. Munir, A. S. Aldwayyan, H. M. Ghaithan, A. A. Ali Ahmed and S. M. H. Qaid, Mater. Sci. Semicond. Process., 2024, 181, 108645 CrossRef CAS.
- F. Aslam, H. Ullah and M. Hassan, Mater. Sci. Eng. B, 2021, 274, 115456 CrossRef CAS.
- T. Chargui, F. Lmai and K. Rahmani, Sol. Energy, 2024, 283, 113016 CrossRef CAS.
- S. Shakeel, P. Song, S. H. Shah, Z. Zada, T. Huang, A. Laref, N. Hakimi and M. Faizan, Mater. Chem. Phys., 2024, 324, 129683 CrossRef CAS.
- A. Ayyaz, G. Murtaza, M. Naeem, A. Usman, S. M. Ramay, M. Irfan and H. Irfan, J. Phys. Chem. Solids, 2024, 188, 111936 CrossRef CAS.
- D. Behera and S. K. Mukherjee, JETP Lett., 2022, 116, 537–546 CrossRef CAS.
- N. Rahman, A. Rauf, M. Husain, N. Sfina, V. Tirth, M. Sohail, R. Khan, A. Azzouz-Rached, G. Murtaza, A. A. Khan, S. A. Khattak and A. Khan, RSC Adv., 2023, 13, 15457–15466 RSC.
- M. Ishfaq, M. Yaseen, S. Shukrullah and S. Noreen, Mater. Chem. Phys., 2024, 313, 128728 CrossRef CAS.
- A. El-marghany, K. Muhammad, M. Sajid, M. Nazar, M. Kashif Masood, Nasarullah, Y. M. Alawaideh and J. Rehman, J. Phys. Chem. Solids, 2025, 198, 112477 CrossRef CAS.
- J. Sarkar, A. Talukdar, P. Debnath and S. Chatterjee, J. Comput. Electron., 2023, 22, 1075–1088 CrossRef CAS.
- G. Volonakis, M. R. Filip, A. A. Haghighirad, N. Sakai, B. Wenger, H. J. Snaith and F. Giustino, J. Phys. Chem. Lett., 2016, 7, 1254–1259 CrossRef CAS PubMed.
- A. H. Slavney, T. Hu, A. M. Lindenberg and H. I. Karunadasa, J. Am. Chem. Soc., 2016, 138, 2138–2141 CrossRef CAS PubMed.
- A. Ayyaz, G. Murtaza, A. Azazi, A. Usman, A. A. Abd El-Moula, A. Khadr Alqorashi, F. Ur R. Ahmad and M. Touqir, Opt. Quantum Electron., 2024, 56, 1204 CrossRef CAS.
- H. Ambreen, S. Saleem, S. A. Aldaghfag, S. Noreen, M. Zahid, Hafsa, M. Ishfaq and M. Yaseen, Phys. B, 2023, 667, 415156 CrossRef CAS.
- Y. Al-Douri, M. Ameri, A. Bouhemadou and K. M. Batoo, Phys. Status Solidi B, 2019, 256(11), 1900131 CrossRef CAS.
- J. Annie Abraham, D. Behera, K. Kumari, A. Srivastava, R. Sharma and S. Kumar Mukherjee, Chem. Phys. Lett., 2022, 806, 139987 CrossRef CAS.
- H. Althib, T. H. Flemban, A. I. Aljameel, A. Mera, M. G. B. Ashiq, Q. Mahmood, B. Ul Haq and S. A. Rouf, Bull. Mater. Sci., 2021, 44, 2–7 CrossRef.
- M. A. Ali, M. Musa Saad H.-E., A. M. Tighezza, S. Khattak, S. Al-Qaisi and M. Faizan, J. Inorg. Organomet. Polym. Mater., 2024, 34, 1609–1619 CrossRef CAS.
- K. Assiouan, A. Marjaoui, J. EL Khamkhami, M. Zanouni, H. Ziani, A. Bouchrit and A. Achahbar, J. Phys. Chem. Solids, 2024, 188, 111890 CrossRef CAS.
- D. Penn, Phys. Rev., 1962, 128, 2093 CrossRef CAS.
- R. Su, Z. Xu, J. Wu, D. Luo, Q. Hu, W. Yang, X. Yang, R. Zhang, H. Yu, T. P. Russell, Q. Gong, W. Zhang and R. Zhu, Nat. Commun., 2021, 12, 2479 CrossRef CAS PubMed.
- B. Bernardo, D. Cheyns, B. Verreet, R. D. Schaller, B. P. Rand and N. C. Giebink, Nat. Commun., 2014, 5, 3245 CrossRef CAS PubMed.
- G. Ayub, N. Rahman, M. Husain, M. Sohail, R. Khan, N. Sfina, M. Elhadi, A. Azzouz-Rached and A. Alotaibi, J. Phys. Chem. Solids, 2024, 188, 111942 CrossRef CAS.
- S. Mahmud, M. A. U. Z. Atik, M. N. Mostakim, Md. Tarekuzzaman and Md. Z. Hasan, Comput. Condens. Matter, 2024, 40, e00950 CrossRef.
- Nasarullah, R. Bousbih, M. Sajid, T. Sattar, A. S. Alshomrany, S. Alharthi, M. A. Amin, Y. A. El-Badry, M. Shaban, A. Ahmad and M. Nazar, Phys. B, 2024, 687, 416070 CrossRef CAS.
- H. D. Alkhaldi, J. Inorg. Organomet. Polym., 2025, 35, 3260–3274 CrossRef CAS.
- S. Tariq, L. El Hachemi Omari, V. Tuyikeze and M. Abid, Mater. Today: Proc., 2023 DOI:10.1016/j.matpr.2023.11.149.
- I. E. Castelli, K. S. Thygesen and K. W. Jacobsen, J. Mater. Chem. A, 2015, 3, 12343–12349 RSC.
- M. Shirayama, H. Kadowaki, T. Miyadera, T. Sugita, M. Tamakoshi, M. Kato, T. Fujiseki, D. Murata, S. Hara, T. N. Murakami, S. Fujimoto, M. Chikamatsu and H. Fujiwara, Phys. Rev. Appl., 2016, 5, 014012 CrossRef.
- W. Meng, X. Wang, Z. Xiao, J. Wang, D. Mitzi and Y. Yan, J. Phys. Chem. Lett., 2017, 8, 2999–3007 CrossRef CAS PubMed.
- M. Asghar, H. S. Waheed, U. Abbas, H. Ullah, M. J. I. Khan, S. M. Wabaidur, A. Ali and Y.-H. Shin, Phys. B, 2024, 682, 415916 CrossRef CAS.
- S. A. Khandy, I. Islam, K. Kaur, A. M. Ali and A. F. A. El-Rehim, Mater. Chem. Phys., 2023, 297, 127293 CrossRef CAS.
- M. Y. H. Khan, S. S. Hasan, Md. Z. Rahman, Md. Rasheduzzaman and Md. Z. Hasan, RSC Adv., 2025, 15, 34643–34668 RSC.
- H. Kim, H. Kwak, I. Jung, M. S. Kim, J. Kim, H. J. Park and K.-T. Lee, Opt. Express, 2021, 29, 35366 CrossRef CAS PubMed.
- F.-T. Zahra, M. M. Hasan, Md. B. Hossen and Md. R. Islam, Heliyon, 2024, 10, e33096 CrossRef CAS PubMed.
- N. Algethami, A. Jehan, N. Rahman, S. Naz Khan, W. Mohammed Almalki, M. Husain, V. Tirth, H. A. Althobaiti, Y. M. Alawaideh, K. M. Abualnaja and G. Alosaimi, Results Phys., 2025, 70, 108163 CrossRef.
- S. Nabi, A. W. Anwar, Z. Wazir, S. S. Hayat, M. Ahmad, M. Tayyab, K. Nabi, M. Shamoil, A. A. Khan and B. S. Khan, Eur. Phys. J. B, 2022, 95, 55 CrossRef CAS.
- A. Shukla, V. K. Sharma, S. K. Gupta and A. S. Verma, Mater. Res. Express, 2020, 6, 126323 CrossRef.
- R.-I. Biega, M. R. Filip, L. Leppert and J. B. Neaton, J. Phys. Chem. Lett., 2021, 12, 2057–2063 CrossRef CAS PubMed.
- F. Zhang, W. Gao, G. J. Cruz, Y. Sun, P. Zhang and J. Zhao, Phys. Rev. B, 2023, 107, 235119 CrossRef CAS.
- F. M. Dezashibi, A. Doroudi and H. Noshad, Optik, 2023, 295, 171475 CrossRef CAS.
- F. Sagredo, S. E. Gant, G. Ohad, J. B. Haber, M. R. Filip, L. Kronik and J. B. Neaton, Phys. Rev. Materials, 2024, 8, 105401 CrossRef CAS.
- V. B. Bobrov, S. A. Trigger, G. J. F. van Heijst and P. P. J. M. Schram, Europhys. Lett., 2010, 90, 10003 CrossRef.
- I. Moreels, G. Allan, B. De Geyter, L. Wirtz, C. Delerue and Z. Hens, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 235319 CrossRef.
- W. Jiang and X. Chen, AIP Adv., 2022, 12, 065106 CrossRef CAS.
- A. Bibi, I. Lee, Y. Nah, O. Allam, H. Kim, L. N. Quan, J. Tang, A. Walsh, S. S. Jang, E. H. Sargent and D. H. Kim, Mater. Today, 2021, 49, 123–144 CrossRef CAS.
- K. V. Baryshnikova, M. I. Petrov, V. E. Babicheva and P. A. Belov, Sci. Rep., 2016, 6, 22136 CrossRef CAS PubMed.
- Nasarullah, M. Z. Choudary, S. A. Aldaghfag, Misbah, M. Yaseen, M. Nazar and R. Neffati, Phys. Scr., 2022, 97, 105705 CrossRef CAS.
- A. El-marghany, K. Muhammad, M. Sajid, M. Nazar, M. K. Masood, Nasarullah, Y. M. Alawaideh and J. Rehman, J. Phys. Chem. Solids, 2025, 198, 112477 CrossRef CAS.
- A. Khan, M. Saeed, A. R. Chaudhry, M. A. Jehangir, M. Ibrar and G. Murtaza, Solid State Commun., 2024, 394, 115698 CrossRef CAS.
- A. Nazir, A. Dixit, E. A. Khera, M. Manzoor, R. Sharma and A. J. A. Moayad, Mater. Adv., 2024, 5, 4262–4275 RSC.
- T. D. Flaim, Y. Wang and R. Mercado, Advances in Optical Thin Films, 2004, vol. 423 Search PubMed.
- J. B. Khurgin, ACS Photonics, 2022, 9, 743–751 CrossRef CAS.
- J.-M. Renoirt, C. Zhang, M. Debliquy, M.-G. Olivier, P. Mégret and C. Caucheteur, Opt. Express, 2013, 21, 29073 CrossRef PubMed.
- S. V. Boriskina, T. A. Cooper, L. Zeng, G. Ni, J. K. Tong, Y. Tsurimaki, Y. Huang, L. Meroueh, G. Mahan and G. Chen, Adv. Opt. Photon, 2017, 9, 775 CrossRef.
- D.-Y. Li, H.-Y. Kang, Y.-H. Liu, J. Zhang, C.-Y. Yue, D. Yan and X.-W. Lei, Chem. Sci., 2024, 15, 953–963 RSC.
- B. Zhou and D. Yan, Matter, 2024, 7, 1950–1976 CrossRef.
- F. Aslam, B. Sabir and M. Hassan, Appl. Phys. A:Mater. Sci. Process., 2021, 127, 1–12 CrossRef.
- H. A. Alburaih, J. Alloys Compd., 2021, 876, 159806 CrossRef CAS.
- S. A. Aldaghfag, A. Aziz, A. Younas, M. Yaseen, A. Murtaza and H. H. Hegazy, J. Solid State Chem., 2022, 312, 123179 CrossRef CAS.
- K. Rangar, A. Soni and J. Sahariya, Multiscale and Multidiscip. Model. Exp. and Des., 2025, 8, 451 CrossRef.
- A. Gilani, Nasarullah, S. A. Aldaghfag and M. Yaseen, J. Phys. Chem. Solids, 2023, 175, 111208 CrossRef CAS.
- K. A. Alrashidi, A. Dixit, A. Nazir, E. A. Khera, S. Mohammad, M. Manzoor, R. Khan and R. Sharma, J. Inorg. Organomet. Polym., 2025, 35, 1764–1778 CrossRef CAS.
- S. Saidi, S. Belhachi, S. Abdalla, J. Y. Al-Humaidi, M. W. Iqbal, M. S. M. Al-Saleem, M. M. Rahman, M. Sillanpää, A. Kumar and S. Singh, J. Comput. Chem., 2025, 46, e70222 CrossRef CAS PubMed.
- E. Haque and M. A. Hossain, J. Alloys Compd., 2018, 748, 63–72 CrossRef CAS.
- Y. Wang, Y.-J. Hu, B. Bocklund, S.-L. Shang, B.-C. Zhou, Z.-K. Liu and L.-Q. Chen, Phys. Rev. B, 2018, 98, 224101 CrossRef CAS.
- J. Martin, L. Wang, L. Chen and G. S. Nolas, Phys. Rev. B:Condens. Matter Mater. Phys., 2009, 79, 115311 CrossRef.
- J. De Boor and E. Müller, Rev. Sci. Instrum., 2013, 6, 84 Search PubMed.
- T. S. Ahmad, N. Ehsan, M. Liaqat, S. M. S. Gilani, A. ul Haq, A. M. Quraishi, D. Abduvalieva, V. Tirth, A. Algahtani, R. M. Mohammed, N. M. A. Hadia, A. Mohammed Alsuhaibani, M. S. Refat and A. Zaman, Results Phys., 2024, 63, 107885 CrossRef.
- S. A. Aldaghfag, S. Saleem, Nasarullah, M. Arshad and M. Yaseen, Phys. B, 2024, 677, 415700 CrossRef CAS.
- D. Behera, B. Mohammed, S. Taieb, B. Mokhtar, S. Al-Qaisi and S. K. Mukherjee, Eur. Phys. J. Plus, 2023, 138, 520 CrossRef CAS.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.