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

Exploring photocatalytic, thermoelectric, and SLME efficiency of photovoltaic absorber layers for halide double antiperovskites K6NaAsX2 (X = Cl, Br, I): a first-principles approach

M. Usman Saeeda, Irslan Saeeda, Shahan Alia, Ahmed Ali Khana, M. Zia Ur Rehmana, Shamiala Pervaiza, Pervaiz Ahmadb, Hosam O. Elansaryc, Sohail Mumtaz*d and Y. Saeed*a
aDepartment of Physics, Abbottabad University of Science and Technology, Abbottabad, KPK, Pakistan. E-mail: saeedy@alumni.kaust.edu.sa; yasirsaeedphy@aust.edu.pk; Tel: +(92)-3454041865
bDepartment of Physics, College of Science and Humanities, Prince Sattam Bin Abdulaziz University, P. O. Box 173, Al-Kharj 11942, Saudi Arabia
cPrince Sultan Institute for Environmental, Water and Desert Research, King Saud University, Riyadh, Saudi Arabia
dDepartment of Chemical and Biological Engineering, Gachon University, 1342 Seongnamdaero, Sujeong-gu, Seongnam-si 13120, Republic of Korea. E-mail: sohail.ahmed2015@gmail.com; Tel: +8201065596861

Received 12th August 2025 , Accepted 24th November 2025

First published on 3rd December 2025


Abstract

Density functional theory (DFT) was used in this study to examine the structural, electronic, optical, mechanical, thermoelectric, photovoltaic, thermodynamic, and photocatalytic characteristics of double antiperovskite (DAP) compounds K6NaAsX2 (where X = Cl, Br, I). In order to optimise lattice parameters and obtain lower bandgaps, first-principles calculations were performed in WIEN2k using the FP-LAPW + LO method with Wu-Cohen GGA (WC–GGA) and the Tran–Blaha modified Becke–Johnson (TB–mBJ) potential. The bandgaps of 1.48 eV, 1.34 eV, and 1.16 eV were found for K6NaAsCl2, K6NaAsBr2, and K6NaAsI2, respectively, according to band structure investigations using TB–mBJ + SOC. The orbital contributions close to the Fermi level were revealed by the density of states. The optical characteristics, including reflectivity, absorption, extinction coefficient, and refractive index, were computed, while elastic stability creteria confirmed mechanical stability. Thermodynamic properties, including heat capacities, entropy, enthalpy, and Gibbs free energy, were also assessed. Spectroscopic limited maximum efficiency (SLME) analysis revealed promising solar cell efficiency, while photocatalytic results indicated strong oxidizing power suitable for water splitting. Overall, the reduced bandgaps and multifunctional behavior indicate these DAPs as promising candidates for eco-friendly optoelectronic and energy applications.


1. Introduction

Increasing industrialization and using fossil fuels excessively have led to severe environmental pollution and an international energy crisis. In order to combat this, scientists are paying more attention to renewable energy sources, especially solar energy, owing to its environmental advantages and sustainability.1–3 Because of their excellent light-absorbing properties and potential to increase power conversion efficiency, perovskite materials are becoming increasingly popular for use in solar energy applications. As a result, they have become a major focus of ongoing research efforts.4–6

Perovskites, ABX3, and the crystal structure of CaTiO3 exist in several structural variants, such as antiperovskites (X3BA) and double perovskites (A2B′B″X6). These configurations offer distinct advantages, including ease of synthesis and structural stability.7–9 A recent advancement in this field is emerging double antiperovskites with the formula X6B″B′A2, which have shown significant potential for diverse technological applications.10 Hybrid halide perovskites such as CH3NH3PbX3 have been extensively studied for high-efficiency solar cells. However, their moisture sensitivity and lead toxicity remain major drawbacks. To overcome these issues, researchers have turned to stable, lead-free alternatives such as halide double perovskites Cs2AgBiX6 (X = Cl, Br, I). Hybrid halide perovskites such as CH3NH3PbX3 have been extensively studied for high-efficiency solar cells. However, their moisture sensitivity and lead toxicity remain major drawbacks. To overcome these issues, researchers have turned to stable, lead-free alternatives such as halide double perovskites Cs2AgBiX6 (X = Cl, Br, I).11–15

Recently, Rani et al. studied new double antiperovskites, namely Na6SOCl2, Na6SOBr2, Na6SOI2, K6SOCl2, K6SOBr2, and K6SOI2, with bandgaps of 4.34 eV, 3.71 eV, 3.33 eV, 3.99 eV, 3.38 eV, and 2.90 eV, respectively, and showed that they are good candidates for thermoelectric applications.16 Yu et al. predicted that Na6SOI2 is particularly promising for low-temperature solid-state sodium-ion battery applications.17 Mebrouki et al. explained that the effect of temperature on the elastic constants C11, C12, and C44 of BaVO3 becomes less significant as the temperature increases.18 Djebari et al. determined parameters such as formation energy, tolerance factor, and octahedral factor to confirm the structural integrity of A3+B4+(O2N)7− and A2+B5+(O2N)7−. Based on their calculated bandgaps, these compounds show potential for applications in ferroelectric and photovoltaic devices.19 The thermal expansion coefficient has been reported to be adversely affected by changes in pressure and temperature.20 Given that double antiperovskites have only recently been proposed, experimental and theoretical studies remain limited, highlighting opportunities for further investigation. Here, we are examining the double antiperovskites K6NaAsX2 (where X = Cl, Br, I) in terms of their structural, electrical, optical, mechanical, thermoelectric, photovoltaic, thermodynamic, and photocatalytic characteristics with a focus on materials exhibiting bandgaps comparable to that of MAPI.21 We also compare our results with the previously studied double perovskite material K2NaAsCl6.14

2. Computational details

All simulations in this study are conducted using density functional theory (DFT), specifically the full-potential linearized augmented plane wave plus local orbitals (FP-LAPW + LO) method, which is renowned for its high accuracy in determining material properties. These computations were carried out using the WIEN2k code.22 The Wu–Cohen generalized gradient approximation (WC–GGA) was employed to account for exchange–correlation effects, which are a crucial component of DFT.23 The modified Becke–Johnson (mBJ) potential was applied on top of WC–GGA to improve the accuracy of bandgap predictions, addressing a well-recognized shortcoming of traditional DFT.24,25 Lead halide perovskites are among the semiconductors for which this combination has previously been validated and shown to be in strong agreement with experimental data.26–31 The majority of the calculations in this study were performed using a 13 × 13 × 13 k-point mesh. Thermoelectric and optical properties were evaluated more precisely with a denser 24 × 24 × 24 mesh. The BoltzTraP code was employed to calculate the thermoelectric properties of K6NaAsX2 (X = Cl, Br, I) over a temperature range of 300–900 K. The technical parameters included a kinetic energy cutoff of −7.0 Rydbergs to distinguish valence from core electrons. A plane-wave expansion cutoff of Gmax = 20 and an angular momentum cutoff of lmax = 12 were used. Structural optimizations were performed until the total energy converged within 10−5 Ry,32 ensuring highly accurate results.

3. Results and discussions

3.1 Structural properties

The first step in understanding a material's physical properties is a thorough examination of its structural features. This study focuses on optimizing the structural parameters of K6NaAsX2 (X = Cl, Br, I) by minimizing the total energy to obtain the ground-state configuration. Fig. 1(a) shows the crystal structure of K6NaAsX2, with initial lattice parameter taken from K2NaAsCl6 (10.31 Å) as a starting point. The energy–volume relationship, shown in Fig. 1(b), used for structural optimization. For fitting the nonlin- ear energy vs. volume curve, initial lattice parameters of Cs2AgBiX6 (X = Cl, Br) were also considered as reference points. The Birch–Murnaghan approach for the equation of state was used to derive the important structural parameters from this volume optimization, including the bulk modulus (B in GPa) and the lattice constant (a in Å).33 The compounds' optimised lattice constants fall between 12.65 Å and 12.72 Å while the bulk moduli vary between 3.16 GPa and 4.19 GPa, as summarized in Table 1. Previously studied double perovskite compound K2NaAsCl6 has lattice parameter 10.31 Å.14 In our study the lattice parameter of compound increases by making antiperovskite of K2NaAsCl6 compound.
image file: d5ra05938k-f1.tif
Fig. 1 For cubic double antiperovskites, (a) crystal structure and (b) the energy vs. volume curve, K6NaAsX2 (X = Cl, Br, I).
Table 1 Optimized lattice parameters a (Å), bulk modulus B (GPa) and Bandgap Eg (eV) of double antiperovskites Cs6AgBiX2 (X = Cl, Br, I) with SCF, SOC and hybrid, goldschmidt tolerance factor (τG), octahedral factor µ, formation energy Ef (eV), effective mass of electron image file: d5ra05938k-t7.tif and effective mass of hole image file: d5ra05938k-t8.tif
Materials a B EGGA EGGA+SOC Hybrid τG µ Ef

image file: d5ra05938k-t9.tif

image file: d5ra05938k-t10.tif

K6NaAsCl2 12.65 3.16 0.62 0.58 1.48 0.87 0.42 −2.06 0.32 0.40
K6NaAsBr2 12.66 4.06 0.61 0.57 1.34 0.91 0.39 −1.94 0.12 0.25
K6NaAsI2 12.72 4.19 0.60 0.54 1.16 0.85 0.34 −1.79 0.03 0.26
K2NaAsCl6 10.31 [14] 3.52 [14]


For K6NaAsX2 (X = Cl, Br, I), the effective masses of electrons image file: d5ra05938k-t1.tif and holes image file: d5ra05938k-t2.tif are shown in Table 1. Since all of the values are less than silicon's electron effective mass image file: d5ra05938k-t3.tif, these compounds appear to be ideal for high-performance device applications.34

The thermal stability of these double perovskites is confirmed by Table 1, which lists the calculated formation energies of the elements and compounds.35 The Goldschmidt tolerance factor (τG) and the octahedral factor (µ) were also used to evaluate the stability of the double antiperovskite materials.36

The ionic radii of K, Na, As, Cl, Br, and I are 1.38 Å, 0.95 Å, 0.58 Å, 1.81 Å, 1.96 Å, and 2.20 Å, respectively. For perovskites and antiperovskites, the crystal structure stability ranges are 0.415 < µ < 0.895 and 0.183 < τG < 1.107.36 Table 1 shows that the calculated octahedral factor (µ) and Goldschmidt tolerance factor (τG) fall within these ranges, confirming the structural stability of all the compounds.

3.2 Electronic band structure

The electronic band structure is a crucial aspect of understanding a material's behavior, as it provides key insights into its metallic or semiconducting nature. It directly affects essential properties such as the optical response and electrical resistivity, which are critical for device design. The band structures of K6NaAsX2 (X = Cl, Br, I) were calculated to investigate their electronic characteristics. Initially, a self-consistent field (SCF) method was used to estimate the bandgaps. Spin–orbit coupling (SOC) and the modified Becke–Johnson (mBJ) potential were then applied to improve agreement with experimental trends and refine the bandgap values, particularly for compounds with gaps below 1 eV. As a validation step, we also recalculated the bandgap of K2NaAsCl2, previously reported to be 3.52 eV.14 Our calculated bandgap values for K6NaAsX2 (X = Cl, Br, I) using the mBJ + SOC (hybrid) approach are 1.48 eV, 1.34 eV, and 1.16 eV, respectively (Fig. 3(a–c)). For comparison, the previously reported double perovskite K2NaAsCl6 has a bandgap of 3.52 eV,14 indicating that our study provides improved bandgap predictions for antiperovskites. Without SOC, the bandgaps are 0.62 eV, 0.61 eV, and 0.60 eV, while with SOC they decrease slightly to 0.58 eV, 0.57 eV, and 0.54 eV, respectively. Furthermore, our goal was to obtain bandgap values close to the experimental observations for MAPI, which exhibits a direct bandgap at a high-symmetry point, a feature also reproduced in our studied compounds.32 Our analysis indicates that the FP-LAPW + lo method, when used alone, significantly underestimates the bandgap values of double antiperovskites. Therefore, our goal is to obtain a bandgap close to 1.55 eV while using lead-free materials, making these compounds promising candidates to replace MAPI for room-temperature applications.37

The computed materials exhibit a direct bandgap both with and without SOC, as shown in Fig. 2(a–c). A slight decrease in bandgap values was observed for K6NaAsX2 (X = Cl, Br, I) upon SOC application. A bandgap close to 1.55 eV is ideal for efficient semiconductor photovoltaic devices, as it allows effective electron excitation with minimal heat loss. Although direct bandgaps are observed in several recently investigated double antiperovskites, their values often fall outside this ideal range. Conversely, compounds with suitable bandgaps are typically indirect, which limits their usefulness in solar applications. Pressure-induced bandgap tuning could be a practical strategy to bring the bandgap closer to the ideal 1.55 eV.


image file: d5ra05938k-f2.tif
Fig. 2 The band structures of (a) K6NaAsCl2 (b) K6NaAsBr2, and (c) K6NaAsI2 without SOC (black) and with SOC (red).

image file: d5ra05938k-f3.tif
Fig. 3 The band structures of (a) K6NaAsCl2 (b) K6NaAsBr2, and (c) K6NaAsI2 with hybrid.

3.3 Density of states (DOS)

We investigated density of states (DOS) of K6NaAsX2 (X = Cl, Br, I), focusing on the atomic-level contributions. From Fig. 4(a–c) analysis reveals that potassium (K) and arsenic (As) atoms predominantly contributed to the Valence Band Maxima (VBM) in all three compounds, while the Conduction Band Minima (CBM) arises from a combination of K, Na, and halogen (Cl, Br, or I) states. Notably, K atoms play a significant role in determining the bandgap. A more detailed examination of the PDOS provides further insight, and Fig. 4 illustrates the distinct orbital contributions of X= Cl, Br in the VBM and CBM regions.
image file: d5ra05938k-f4.tif
Fig. 4 Density of States(DOS) of (a) K6NaAsCl2, (b) K6NaAsBr2 (c) K6NaAsI2.

The K-p orbital predominates in the CBM, whereas the As-p orbital makes a substantial contribution to the VBM, as shown in Fig. 4(a). Contributions from other atoms are relatively minor. The Cl-p orbital also exhibits a notable presence in the CBM, while d-orbital contributions to the VBM are limited. Within the VBM, Cl's p-orbital exerts a stronger influence than those of Na and As, and K's d-orbital extends throughout the valence region down to −5 eV. In contrast, Br's d-orbital contributes negligibly to both the VBM and CBM. As shown in Fig. 4(b), a similar pattern emerges: the CBM is primarily influenced by the d-orbital of K and the p-orbital of Br, while the VBM is mainly governed by the p-orbital of K. In Fig. 4(c), the CBM is largely shaped by the p-orbitals of both I and K, whereas the VBM shows a dominant contribution from the d-orbital of K. These recurring orbital characteristics underscore the significant roles of halogen p-orbitals and potassium orbitals in shaping the electronic structure of K6NaAsX2 (X = Cl, Br, I).

3.4 Optical properties

The advancement of optical technologies depend on the optical properties of materials. Perovskites have become one of the most optimistic materials for high-efficiency optoelectronic applications.37,38 A material's interaction with an external electromagnetic field is described by its complex dielectric function, which is (ω) = 1(ω) + i∈2(ω).39 In this expression, 1(ω) symbolizes the real component, which has to do with the material's dispersion, while 2(ω) is the imaginary part, which accounts for absorption.

1(ω), represents the dielectric constant and indicates a material becomes polarized in electric field due to induced dipole formation. The imaginary part, 2(ω), reflects the material's ability to absorb and attenuate electromagnetic radiation. These dielectric properties are closely linked to the material's electronic structure, particularly through the momentum transition matrix elements and the joint density of states (DOS). Furthermore, 1(ω) can be obtained from 2(ω) using the Kramers–Kronig relations.40 When 1(ω) > 0, light can propagate through the material; when 1(ω) < 0, the electromagnetic wave is attenuated and cannot propagate.41,42

The frequency-dependent real and imaginary components of the dielectric function provide important information about the optical characteristics of a material, such as its refractive index, absorption coefficient, reflectivity, and optical conductivity. The refractive index, n(ω), and the extinction coefficient, k(ω), are the two main factors that define the optical response.43,44

image file: d5ra05938k-t4.tif
where n(ω) governing phase velocity and k(ω) give information about absorption of material.

The compounds K6NaAsX2 (X = Cl, Br, I) exhibit high static dielectric constants, 1(ω), of 6.94, 7.43, and 8.34, respectively. These values are substantially higher than those of other potential solar cell materials, including MAPbI3 (5.4), FAPbI3 (5.7), Cs2LiTlBr6 (1.6), Cs2NaTlBr6 (1.7), and Cs2AgCrI6 (5.7).40,45–49 Materials with a high dielectric constant are beneficial for solar cell applications, as their enhanced polarization and reduced exciton binding energies facilitate more efficient charge separation and transport.50,51 For the chloride, bromide, and iodide-based compounds, the real part of the dielectric function, 1(ω), shows maximum polarization values of 2.01, 1.83, and 1.94 inside the visible range at photon energies of 2.50 eV, 2.45 eV, and 2.28 eV, respectively. At photon energies of 1.78 eV, 1.73 eV, and 1.59 eV, respectively, the imaginary part of the dielectric function, 2(ω), representing optical absorption, peaks at 6.71, 6.81, and 7.02 for the chloride, bromide, and iodide compounds. These values are marginally outside the visible range. The first peak at 1.5 eV originates from direct transitions at the band edge, where halogen p-derived valence states transition to the conduction-band minimum dominated by Na-site cation s orbitals. The second peak around 2.6 eV is mainly attributed to transitions from deeper halogen p states and K-site hybridized orbitals in the valence band to conduction states. The higher-energy feature at 4.8 eV is associated with transitions from lower valence states to high-lying conduction bands with strong As d character, consistent with the large joint density of states in this region52,53 These results, shown in Fig. 5, demonstrate the materials' linear optical response. We also calculated key optical properties, including the optical conductivity, σ(ω), reflectivity, R(ω), extinction coefficient, k(ω), absorption coefficient, α(ω), and refractive index, n(ω), to further evaluate their optoelectronic behavior, as illustrated in Fig. 6. The static dielectric constants, (0), for K6NaAsCl2, K6NaAsBr2, and K6NaAsI2 are 2.97, 2.99, and 3.20, respectively,35 aligned well with our observations.


image file: d5ra05938k-f5.tif
Fig. 5 Optical properties of K6NaAsX2 (X = Cl, Br, I).

image file: d5ra05938k-f6.tif
Fig. 6 Optical properties (a) refracrive index, (n, k), (b) optical conductivity, σ, (c) absorption coefficient, α and (d) reflectivity, R of K6NaAsX2 (X = Cl, Br, I).

As shown in Fig. 6(a), the static refractive indices, n(ω), for K6NaAsCl2, K6NaAsBr2, and K6NaAsI2 are 3.11, 3.13, and 3.14, respectively. The maximum values of n(ω) and the extinction coefficient, k(ω), occur at photon energies of 1.64 eV and 2.01 eV for K6NaAsCl2, 1.51 eV and 1.83 eV for K6NaAsBr2, and 1.29 eV and 1.67 eV for K6NaAsI2. As shown in Fig. 6(b), the optical conductivity, σ(ω), for K6NaAsCl2 reaches a maximum value of 2.17 × 104 (Ω cm)−1 at 2.74 eV, while K6NaAsBr2 and K6NaAsI2 peak at 1.91 × 104 (Ω cm)−1 at 2.66 eV and 1.87 × 104 (Ω cm)−1 at 2.54 eV, respectively. This indicates that K6NaAsCl2 exhibits the highest optical conductivity within the visible light range (1.87–2.17 eV), which is beneficial for photovoltaic applications. The absorption coefficient has a direct correlation with solar energy conversion efficiency, which quantifies how far specific light frequencies penetrate the material before being absorbed. As shown in Fig. 6(c), the absorption coefficient, α(ω), exhibits peak values of 3.2 × 105 cm−1 at 2.86 eV (433.51 nm) for K6NaAsCl2, 2.85 × 105 cm−1 at 2.73 eV for K6NaAsBr2, and 2.60 × 105 cm−1 at 1.84 eV for K6NaAsI2. These results are comparable to MAPI, which exhibits absorption in the range of 104–105 cm−1, indicating strong absorption in the visible region, particularly for K6NaAsCl2. As illustrated in Fig. 6(d), the maximum reflectivity, R(ω), for K6NaAsCl2, K6NaAsBr2, and K6NaAsI2 is 0.32%, 0.31%, and 0.30% at approximately 0.21 eV, 0.19 eV, and 1.86 eV, respectively. These low reflectivity values suggest efficient light absorption, which is beneficial for photovoltaic performance.

3.5 Photocatalytic properties

Direct bandgap semiconductor materials can effectively harness solar energy for hydrogen production through water splitting.54,55 This process offers a promising route for generating clean, renewable energy via photocatalytic water splitting.56,57 In this process, photo-generated holes oxidize water molecules, while the excited electrons drive the corresponding reduction reaction.58

A material must have a valence band maximum (VBM) below the O2 oxidation potential (1.23 eV) and a conduction band minimum (CBM) above the H+ reduction potential (0 eV) in order to be appropriate for photocatalytic water splitting. This alignment ensures that photo-generated holes can drive water oxidation, while excited electrons facilitate proton reduction. To assess this criterion, Mulliken electronegativity is used to estimate the locations of the valence and conduction band edges in the studied materials (Table 2).59

Table 2 Photocatalytic parameters of K6NaAsX2 (X = Cl, Br, I) at pH = 0
Compounds χ (eV) Eg (eV) ECB (eV) ECB (eV)
K6NaAsCl2 1.09 1.48 2.66 1.18
K6NaAsBr2 1.08 1.34 2.74 1.40
K6NaAsI2 1.07 1.16 2.85 1.69


Fig. 7 illustrates that, in photocatalytic water splitting, the redox potentials are referenced against the standard hydrogen electrode (SHE) and then converted to the vacuum energy scale, which is commonly used in solid-state physics. These values indicate the energy levels at which electrons must be supplied (for reduction) or removed (for oxidation) for the redox reactions to proceed. To enable water splitting under sunlight, a material's CBM must be 0 eV, allowing electrons to reduce H+, while its VBM must be 1.23 eV to accept electrons from water and drive oxygen evolution.60 This reference allows the CBM and VBM of the material to be positioned relative to the water redox potentials.61 The conduction band (CB) and valence band (VB) potentials of a photocatalyst shift with pH, typically by about 0.059 eV per pH unit as calculated by using Nernst equation. At lower pH = 0, the CB and VB potentials are more positive, making the holes more oxidative but potentially reducing the efficiency of electron-driven reductions. When pH increases from 0 to 7 the band edges increase. At pH = 7, the band edges are less positive, which may favor reduction reactions but could weaken the oxidative power of holes for certain reactions.62


image file: d5ra05938k-f7.tif
Fig. 7 Photo-catalytic properties of (a) K6NaAsCl2 (b) K6NaAsBr2, and (c) K6NaAsI2.

From the figure, it is evident that all studied materials possess suitable VBM positions for water oxidation, but their CBM positions lie below the required level for water reduction, indicating that they are ineffective for complete water splitting via photocatalysis. The band-edge alignment of the three double antiperovskite compounds, K6NaAsX2 (X = Cl, Br, I), show good response for oxidation relative to the water-splitting redox potentials across different pH values.63,64

3.6 Thermoelectric properties

In metallic compounds, the high density of free electrons tends to cancel out potential differences, leading to low Seebeck coefficient values. Since electrons in metals transport both charge and heat, thermal energy can flow easily through the material. This high thermal conductivity is undesirable for thermoelectric applications, as it diminishes the temperature gradient required for efficient power generation. Overall, the combination of a low Seebeck coefficient(S) and high electronic thermal conductivity (κe) makes metals inefficient for thermoelectric (TE) applications. In insulators, the low carrier concentration allows even a small temperature difference to generate a large voltage, resulting in a high Seebeck coefficient (hundreds to thousands of µV K−1). Since there are very few free electrons, heat is not efficiently conducted electronically, which is beneficial for thermoelectric efficiency. However, the lack of mobile charge carriers limits power generation. Optimal thermoelectric materials achieve a balance between electrical conductivity and thermal insulation. Materials with an electronic carrier concentration on the order of 1019 cm−3 are considered ideal for efficient thermoelectric performance.65

Fig. 8(a) illustrates the temperature dependence of the Seebeck coefficient, S (µV K−1). As shown in Table 3, the peak value of S at 300 K is observed for K6NaAsI2, reaching 250.19 µV K−1, while K6NaAsBr2 exhibits the lowest value at 204.53 µV K−1. The Seebeck coefficient of K6NaAsCl2 lies between the other two materials, at 240.13 µV K−1 at 300 K, which is significantly higher than that of the previously reported double perovskite K2NaAsCl6 (199 µV K−1 at 300 K).14 Across all three compounds, the decrease in Seebeck coefficient with increasing temperature while remaining positive. At 900 K, K6NaAsCl2 exhibits the lowest Seebeck coefficient among the studied materials.


image file: d5ra05938k-f8.tif
Fig. 8 Thermeoelctric properties (a) seebeck coefficient, S, (b) electronic conductivity, σ/τ , (c) electronic thermal conductivity, κe/τ, (d) power factor, S2σ/τ, (e) figure of merit, ZT, and (f) lattice thermal conductivity, κL, with temperature variation from 300 K to 900 K for K6NaAsX2 (X = Cl, Br, I).
Table 3 Thermoelectric propertes, S (µV K−1), σ/τ(1018 × 1 Ω−1 ms−1), κe/τ(1014 × W mK−1 s−1), κL(W mK−1), PF(1011 W mK−2 s−1) and ZT (1 K−1) (with calcuated κL at 300 K for K6NaAsX2 (X = Br, Cl)
Compounds S σ/τ κe/τ κL PF ZTL
K6NaAsCl2 240.13 4.65 0.59 3.34 1.99 0.77
K6NaAsBr2 204.53 2.88 0.74 2.03 2.24 0.71
K6NaAsI2 250.19 2.70 0.57 2.60 1.98 0.78
K2NaAsCl6 199[14] 4.31 [14] 0.36 [14] 1.72 [14] 0.11 [14]


The electrical conductivity's temperature dependency, represented as σ/τ (1 Ω−1 ms−1), is displayed in Fig. 8(b). At room temperature (300 K), the highest σ/τ value is observed for K6NaAsCl2, reaching 4.65 × 1018 (1 Ω−1 ms−1), while the lowest is recorded for K6NaAsI2 at 2.70 × 1018 (1 Ω−1 ms−1). For the previously reported double perovskite K2NaAsCl6, this value reaches 4.31 × 1018 (1 Ω−1 ms−1).14 For all compounds, electrical conductivity increases with rising temperature. At 900 K, K6NaAsBr2 exhibits the highest σ/τ value of 11.82 × 1018 (1 Ω−1 ms−1), whereas K6NaAsCl2 shows a lower value of 10.14 × 1018 (1 Ω−1 ms−1). First-principles electron–phonon coupling would provide a more quantitative description, which we identify as a direction for future work.

Fig. 8(c) demonstrates how the electronic contribution to thermal conductivity varies with temperature, denoted as κe/τ (in units of 1014 W mK−1 s−1). The contribution of charge carriers, such as electrons or holes, to the overall thermal conductivity is represented by this number. The Wiedemann–Franz law (κ = L × σ × T), where L is the Lorentz number, σ is the electrical conductivity, and T is the absolute temperature, establishes a direct relationship between the electrical conductivity (σ) and the electronic thermal conductivity (κe). Since both κe/τ and σ/τ are proportional to each other (sharing the same relaxation time, τ), they exhibit similar trends with temperature. Consequently, the κe/τ curve closely mirrors the behavior observed for σ/τ.

At 300 K, the lowest κe/τ is estimated for K6NaAsI2 at 0.57 × 1014 W mK−1 s−1, whereas K6NaAsBr2 exhibits the highest value of 0.74 × 1014 W mK−1 s−1 at the same temperature. For the previously reported double perovskite K2NaAsCl6, this value is 0.36 × 1014 W mK−1 s−1,14 which increases to 0.59 × 1014 W mK−1 s−1 in the antiperovskite K6NaAsCl2. A similar trend is consistently observed across the entire temperature range, remaining stable even at elevated temperatures up to 900 K.

Fig. 8(d) illustrates the temperature dependence of the electronic power factor (PF), expressed in units of W mK−2 s−1. The power factor is calculated using the relation: PF = S2 × σ/τ; at room temperature (300 K), the maximum PF is predicted for K6NaAsBr2, reaching 2.40 × 1011 W mK−2 s−1, whereas the lowest value of 1.98 × 1011 W mK−2 s−1 is observed for K6NaAsI2. The PF for K6NaAsCl2 is 1.99 × 1011 W mK−2 s−1, which is higher than that of the previously reported double perovskite K2NaAsCl6 (1.99 × 1011 W mK−2 s−1).14 A similar trend is maintained at higher temperatures.

For most solid materials, thermal conductivity generally decreases with increasing temperature. Slack's equation offers a thorough understanding of heat transfer in solids and is frequently used to determine the temperature-dependent lattice thermal conductivity.66,67

The lattice thermal conductivity, κL (W mK−1), calculated using Slack's method, is shown in Fig. 8(f). Increased phonon scattering, thermal expansion, and weakening of interatomic bonds are the causes of the observed reduction in κL with increasing temperature. At 300 K, K6NaAsCl2 exhibits the highest lattice thermal conductivity of 3.34 W mK−1, whereas K6NaAsBr2 shows the lowest value of 2.03 W mK−1. The lower κL in K6NaAsBr2 contributes to its highest thermoelectric figure of merit (ZT) at 450 K (see Fig. 8(e)). Increased phonon–phonon scattering, which hinders the passage of thermal energy and lowers lattice thermal conductivity, results from lattice vibrations (phonons) being more energetic and exhibiting increasing frequency dependence as temperature rises. Interatomic lengths are further changed by thermal expansion, which has an impact on thermal transport and phonon dispersion.68,69 Additionally, elevated temperatures can weaken interatomic bonding, decreasing phonon group velocities and further lowering κL. Understanding these factors provides a comprehensive explanation for the temperature-dependent decline in lattice thermal conductivity.

A dimensionless metric known as the figure of merit, or ZT, is used to quantify a thermoelectric material's performance or efficiency, which is its capacity to transform heat into electricity. ZT is inversely proportional to total thermal conductivity (κ = κe + κL) and directly proportional to power factor (PF). The goal is to decrease thermal conductivity and maximize power factor (high Seebeck coefficient and electrical conductivity) in order to raise ZT. (ZT = S2σ/κe + κL). As shown in Fig. 4(e), the ZT values at 300 K, calculated using the average total thermal conductivity (κ), which includes both electronic and lattice contributions, range from 0.71 to 0.78. Among the studied compounds, K6NaAsI2 exhibits the highest ZT value of 0.78 at 300 K, primarily due to its comparatively lower lattice thermal conductivity (κL). The ZT value for K6NaAsCl2 is reported as 0.77, which is much greater than that of the adjacent perovskite K2NaAsCl6, which has a value of 0.11.14

3.7 Thermodynamic properties

One of the key thermodynamic parameters frequently analyzed in materials is the heat capacity (CP). Although it is one of the more challenging quantities to interpret physically, it provides significant insight, with over six distinct definitions available.70 Fig. 9(a) shows that the heat capacity at constant pressure (CP) increases rapidly at low temperatures and then begins to plateau. Among the three compounds, K6NaAsI2 (green) exhibits the highest CP, followed by K6NaAsBr2 (red), and then K6NaAsCl2 (black). This trend likely reflects the increasing atomic mass of the halides (Cl < Br < I), which affects vibrational modes and, consequently, the heat capacity. The heat capacity at constant volume (CV) can provide additional information regarding solid-phase transitions, energy band structure, and lattice vibrations.71
image file: d5ra05938k-f9.tif
Fig. 9 Thermodynamic properties (a) heat capacity, Cp, (b) heat capacity, Cv, and (c) energies, with temperature variation from 0 K to 1000 K for K6NaAsX2 (X = Cl, Br, I).

The results for the heat capacity at constant volume (CV) are presented in Fig. 9(b). CV increases quickly below 300 K, while the rate of increase slows down at higher temperatures. In accordance with the Dulong–Petit limit, CV approaches a constant value at 1000 K, reaching a maximum of 247 cal per cell per K. At any given temperature, CV increases in the order of Cl < Br < I, reflecting the influence of the halide atomic mass on lattice vibrations. Understanding a material's thermal behavior is essential, as energy conversion and heat dissipation occur during practical use. Fig. 9(c) shows the temperature-dependent variations of entropy multiplied by temperature (TS), enthalpy (H), and Gibbs free energy (F), represented by the top, middle, and bottom curves, respectively. These quantities provide insight into the thermodynamic stability and energy characteristics of the compounds. The formulas are H = U + pV and F = UTS, where U, p, and V represent the internal energy, pressure, and volume of the system, respectively. The T·S curves become steeper as CV gradually increases from Cl to Br and I at a constant temperature. Additionally, Fig. 9(c) shows that enthalpy (H) rises with temperature, T·S increases more rapidly, and Gibbs free energy (F) decreases, indicating enhanced thermodynamic stability at higher temperatures. Among the studied compounds, K6NaAsI2 (green) consistently exhibits the highest T·S and enthalpy values, along with the most negative free energy, suggesting superior thermal stability. Enthalpy rises as a result of the system's thermal motion being intensified by temperature increases, which also raise internal energy, pressure, and volume. F = UTS < U + pVTS = HTS, F < HTS < 0 (ref. 71) since H < TS and the three materials' enhanced range of H is less than TS.

3.8 SLME analysis

The spectroscopic limited maximum efficiency (SLME) approach was used to assess the photovoltaic efficiency of K6NaAsX2 (X = Cl, Br, I). The SLME method provides a more accurate estimate of the maximum power conversion efficiency of solar cells by extending the traditional Shockley–Queisser model by taking into account the material's thickness and absorption spectrum.72

The global sun spectrum (AM 1.5G), the absorption spectrum calculated from DFT calculations, and the fundamental and direct-allowed band gap (Eg) of the material are the input parameters used to calculate the SLME.73 Fig. 10 presents the calculated SLME (%) for K6NaAsX2 (X = Cl, Br, I) at fixed thicknesses and temperatures. The results indicate that device temperature, band gap, and material thickness have the most significant impact on SLME. As shown in Fig. 10, SLME increases gradually with increasing material thickness and eventually saturates, approaching a constant value. In comparison K6NaAsX2 (X = Cl, Br, I) SLME efficiencies of @6.82%, @7.34% and @7.73% respectively for 0.5 µm thick layer, according to standard SLME analysis. Because of this, K6NaAsI2 is a better photovoltaic material and a compound that may be used in solar cell applications with a layer of material that is 500 nm thick. However, K6NaAsX2 (where X = Cl, Br, I) has maximum efficiency of @31.04%, @31.61%, and @32.41%. The experimental efficiency limit of CH3NH3PbI3 is 25% for thicknesses of 0.4 and 0.5 µm.74 The SLME approach shows that these compounds have a maximum effectiveness of more than 25%, proving the veracity of our findings.


image file: d5ra05938k-f10.tif
Fig. 10 SLME Efficiency (η) vs. Thickness (µm) of K6NaAsX2 (X = Cl, Br, I).

3.9 Elastic properties

Because they show how a material reacts to external loads, effective elastic constants are crucial for evaluating a material's mechanical stability and practicality. Together with the bulk modulus (B) and shear modulus (G), the elastic constants C11, C12, and C44 determine a material's resistance to compression and deformation. When taking into account industrial applications, these factors are very crucial. The computed elastic constants for K6NaAsX2 (X = Cl, Br, I) are summarized in Table 4. For cubic crystals, the Born–Huang stability criteria must be satisfied: C11 > 0, C44 > 0, C11 + 2C12 > 0, C11C12 > 0, and C12 < B < C11.75 All studied materials meet these conditions, confirming their mechanical stability.
Table 4 Elastic properties of K6NaAsX2 (X = Cl, Br, I)
Compounds K6NaAsCl2 K6NaAsBr2 K6NaAsI2
C11 16.79 13.83 17.03
C12 1.38 2.19 1.67
C44 1.57 3.32 2.66
B 6.52 6.55 6.82
GV 4.02 4.18 4.67
GR 2.30 3.94 3.60
G 3.16 4.06 4.14
Y 10.02 10.33 11.40
B/G 2.06 1.61 1.64
C 7.70 5.45 7.68
C −0.18 −0.41 −0.99
ν 0.27 0.23 0.22
A 0.20 0.60 0.34
Az 1.27 1.69 1.41
Au 3.73 0.30 1.47
ξ 0.23 0.36 0.25


The elastic constants of the materials can be further evaluated using the Voigt–Reuss–Hill (VRH) approximation, which offers a reliable approach to estimate macroscopic elastic moduli from elastic constants of the single crystal.76,77

In the Voigt–Reuss–Hill (VRH) approximation, the subscripts V and R refer to the Voigt and Reuss bounds, representing the upper and lower limits of the elastic moduli, respectively. As listed in Table 4, the Voigt shear moduli (GV) for K6NaAsX2 (X = Cl, Br, I) are 4.02 GPa, 4.18 GPa, and 4.67 GPa, respectively. These values reflect the materials' resistance to shear deformation, indicating that they are relatively soft yet mechanically stable.

Poisson's ratio (µ) and Pugh's ratio (B/G) are widely used to evaluate a material's ductility and brittleness. A material is considered ductile if µ > 0.26 or B/G > 1.75, whereas values below these thresholds indicate brittleness.78 According to these criteria, K6NaAsBr2 and K6NaAsI2 are ductile, while K6NaAsCl2 is classified as brittle. Generally, ionic materials have µ values ranging from 0.3 to 0.4, with the theoretical upper limit being 0.5, whereas covalent materials typically exhibit lower µ values around 0.2. For K6NaAsX2 (X = Cl, Br, I), the calculated Poisson's ratios are 0.27, 0.23, and 0.22, respectively, suggesting that these compounds exhibit more covalent than ionic character (Fig. 11).79


image file: d5ra05938k-f11.tif
Fig. 11 Elastic properties (a) elastic modulus, (b) modulus (bulk B, shear G and young Y), (c) ratio's (Pugh's (B/G), Poisson's (ν) and isotropy index (A), and (d) cauchy's pressure values (C′, C″) of K6NaAsX2 (X = Cl, Br, I).

Elastic anisotropy is closely associated with the formation of microcracks under mechanical stress, which can impact the durability and reliability of materials in practical applications. To evaluate this property, the elastic anisotropy index was calculated for K6NaAsX2 (X = Cl, Br, I).80

A material is considered isotropic if the anisotropy index A equals 0 or 1; any deviation from these values indicates elastic anisotropy.80,81 The mechanical properties of K6NaAsX2 (X = Cl, Br, I), summarized in Table 4, were evaluated using standard elastic relations. Among the three compounds, K6NaAsCl2 exhibits lower values of shear modulus (G) and Young's modulus (Y), indicating reduced stiffness and a lower resistance to plastic deformation compared to K6NaAsBr2 and K6NaAsI2.

Cauchy's pressure, defined as the difference between the elastic constants C12C44, provides insight into a material's ductility. Positive values generally indicate ductile behavior, whereas negative values suggest brittleness. As shown in Table 4, the positive Cauchy's pressure values for K6NaAsX2 (X = Cl, Br, I) confirm the ductile nature of these compounds.77

The shear constants, which serve as critical indicators of a material's dynamic stability, exhibit positive values for K6NaAsCl2, K6NaAsBr2, and K6NaAsI2, as reported in Table 4. These results confirm the mechanical stability of the compounds.

Additionally, anisotropic parameters such as the Zener anisotropy index (AZ) and the Universal anisotropy index (AU) were calculated using the second-order elastic constants (SOECs),82,83 as presented in Table 4. For isotropic materials, AZ equals 1, indicating a uniform shear response. Any deviation from unity, as observed in the computed AZ and AU values, confirms the anisotropic nature of K6NaAsX2 (X = Cl, Br, I). Furthermore, the elastic constants were used to compute the Kleinman parameter (ξ), reported in Table 5. The relatively low values of ξ indicate that these materials exhibit high resistance to bond angle distortions and bond bending.77

Table 5 Thermodynamic parameters for K6NaAsX2 (X = Br, Cl)
Parameters K6NaAsCl2 K6NaAsBr2 K6NaAsI2
vt (ms−1) 1547.29 1589.10 1479.74
vl (ms−1) 2852.04 2728.81 2554.71
vm (ms−1) 1685.86 1712.90 1596.52
θD (K) 90.79 98.52 96.93
ωD (THz) 74.91 81.32 80.02
Hv (GPa) 0.002 0.003 0.003
Tm (K) 652.28 634.75 653.67


3.10 Thermal properties

Understanding thermodynamic properties is crucial for improving energy conversion efficiency in solar cells and other photovoltaic devices. To investigate these characteristics, simulations were performed on K6NaAsX2 (X = Cl, Br, I) at high temperatures and constant pressure using the quasi-harmonic Debye approximation, based on their ground-state structural parameters. This approach allowed the evaluation of essential thermodynamic parameters relevant to optoelectronic performance.

The overall efficiency of solar cells depends on key factors such as short-circuit current, open-circuit voltage, and energy conversion efficiency. Thermodynamic quantities, including average sound velocity (vm), Vickers hardness (HV), melting temperature (Tm), Debye temperature (θD), and Debye frequency (ωD), play a critical role in optimizing materials for such applications.84 Given that solar devices often operate under harsh conditions, thermodynamic stability is a key criterion for material selection. As shown in Table 5, increasing the halogen atom's atomic number in K6NaAsX2 (X = Cl, Br, I) leads to a systematic increase in parameters such as transverse sound velocity, reflecting predictable trends in physical behavior due to halogen substitution. Among these parameters, the Debye temperature (θD) is particularly significant, as it reflects the strength of atomic bonding and directly influences elasticity, heat capacity, and melting point. Higher θD values indicate stronger bonding, improved thermal conductivity, and greater thermal stability. The θD values for the studied compounds, derived from average sound velocities, are summarized in Table 5.

This calculation is based on key physical constants, including Planck's constant image file: d5ra05938k-t5.tif, Boltzmann's constant (KB), and Avogadro's number (NA), as well as the total number of atoms in the compound (N) and the average acoustic sound velocity (vm). These parameters are incorporated into the relevant equations to evaluate the thermodynamic properties.77

During solar cell fabrication, materials are often exposed to high temperatures, particularly during processes such as crystalline silicon growth or metal contact formation. Therefore, the melting temperature of a material is a critical parameter. A low melting point may lead to deformation, cracking, or melting, which can negatively impact device performance and reliability. Conversely, an excessively high melting point can complicate processing and increase manufacturing costs. The elastic constant C11 is frequently used to estimate a material's melting temperature,85 providing valuable insight into its thermal stability and suitability for high-temperature solar cell fabrication.77

The melting temperatures of the proposed compounds are listed in Table 5. Among them, K6NaAsI2 exhibits the highest melting point (653.67 K), followed by K6NaAsCl2 (652.28 K) and K6NaAsBr2 (634.75 K), indicating stronger atomic bonding in the iodine-based compound. The Debye frequency, ωD, which represents the maximum vibrational frequency that phonons can attain during heat transfer, is another key parameter for assessing thermal behavior. Additionally, the Vickers hardness (HV) and ωD of these compounds were calculated using established theoretical models.86

These properties provide valuable insights into the mechanical strength and thermal stability of the materials. As shown in Table 5, the Debye temperature, Debye frequency, melting temperature, and hardness all increase with the atomic number of the halogen, reflecting stronger atomic bonding and greater thermal durability across the series.

4. Conclusion

In this study, density functional theory (DFT) was employed to investigate the structural, electronic, optical, elastic, thermoelectric, thermodynamic, photovoltaic, and photocatalytic properties of novel double antiperovskites (DAP) K6NaAsX2 (X = Cl, Br, I) using the FP-LAPW + lo method (WIEN2k). Structural optimizations were performed with WC-GGA and found that the lattice constants are 12.65 Å (K6NaAsCl2), 12.66 Å (K6NaAsBr2) and 12.72 Å (K6NaAsI2) which are greater than previously studied K2NaAsCl6 (10.31 Å). By calculating the bandgap of all materials, we observe that these materials have small values of the bandgap without incorporatin spinorbit coupling (SOC) which are 0.62 eV (K6NaAsCl2), 0.61 eV (K6NaAsBr2) and 0.60 eV (K6NaAsI2). These values of the bandgap are further decreased on applying SOC from 0.62 eV to 0.58 eV for K6NaAsCl2, 0.61 to 0.57 eV for K6NaAsBr2 and 0.60 eV to 0.54 eV for K6NaAsI2. By applying modified Beck–Johnson approximation (mBJ) along with SOC (Hybrid calculation), we obtained direct bandgap of 1.48 eV for K6NaAsCl2 (3.52 eV for K2NaAsCl6), 1.34 eV (K6NaAsBr2) and 1.16 eV (K6NaAsI2). DOS analysis elucidated the orbital contributions close to the Fermi level from which As-p orbital has maximum contribution in valence band than others and K-p orbital has maximum the highest contribution in conduction band in all studied materials. Optical properties, including refractive index, absorption, reflectivity, and optical conductivity, were presented. These results show that K6NaAsCl2 has the highest absorption of 3.2 × 105 cm−1 at 2.86 eV (433.51 nm). Effective mass of electron image file: d5ra05938k-t6.tif for (K6NaAsX2) (X = Cl, Br, I) 0.32, 0.12 and 0.03 respectively which are less than that of Si (1.09) respectively, therefore it is expected that the carrier mobility of these materials is greater than Si which is strongly demanded for optoelectronic devices. Thermoelectric properties show that K6NaAsCl2 has ZT of 0.77 at 300 K and crosses 0.8 at 900 Kwhich is greater than reported double perovskite K2NaAsCl6 (0.11). Calculated elastic constants confirm mechanical stability, while thermodynamic parameters such as CP, CV, enthalpy, Gibbs free energy, and TS were computed. The values for CP and CV for all materials increase abruptly upto 200 K while gradually increases from 200 K to 1000 K. According to standard SLME analysis, the SLME efficiencies of the material under study were 6.82%, 7.34%, and 7.73% for K6NaAsX2 (X = Cl, Br, I) at 0.5 µm thick layer, respectively. The band-edge alignment of the all studied DAP compounds show good response for oxidation relative to the water-splitting redox potentials across pH values from 0 to 7. These DAP demonstrate potential for optoelectronic applications and solar water-splitting due to their optimal bandgaps and strong photocatalytic activity.

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

Data will be made available on request.

Acknowledgements

The authors extend their appreciation to Ongoing Research Funding program - Research Chairs (ORF-RC-2025-5522), King Saud University, Riyadh, Saudi Arabia. The author Y. Saeed would like to thank Higher Education Commission (HEC) of Pakistan for providing grant under NRPU-15844.

References

  1. T. Wu, M. L. Yao and M. Q. Long, First principle calculations of interface interactions and photoelectric properties of perovskite CsPbX3 (X = Cl, Br, I) and penta-graphene van der Waals heterostructures, Acta Phys. Sin., 2021, 70, 056301 CrossRef.
  2. H. K. Ting, L. Ni, S. B. Ma, Y. Z. Ma, L. X. Xiao and Z. J. Chen, Progress in electron-transport materials in application of perovskite solar cells, Acta Phys. Sin., 2015, 64, 038802 CrossRef.
  3. P. Zhao, J. Su, Y. J. Guo, L. Wang, Z. H. Lin, Y. Hao, X. P. Ouyang and J. J. Chang, Cs2TiI6: a potential lead-free all-inorganic perovskite material for ultrahigh-performance photovoltaic cells and alpha-particle detection, Nano Res., 2022, 15, 2697–2705 CrossRef CAS.
  4. Y. Guo, Y. B. Xue and L. Q. Xu, Interfacial interactions and enhanced optoelectronic properties of GaN/perovskite heterostructures: insight from first-principles calculations, J. Mater. Sci., 2021, 56, 11352–11363 CrossRef CAS.
  5. G. M. Sui, G. J. Yan, G. Yang, B. Zhang and Y. Q. Feng, Theoretical investigation on structure and optoelectronic performance of two-dimensional fluorbenzidine perovskites, Acta Phys. Sin., 2022, 71, 208801 CrossRef.
  6. D. Wang, H. M. Zhu, Z. M. Zhou, Z. W. Wang, S. L. Lv, S. P. Pang and G. L. Cui, Effect of solvent on the perovskite thin film morphology and crystallinity, Acta Phys. Sin., 2015, 64, 038403 CrossRef.
  7. Y. B. Fu, H. Wei, L. Wei, H. D. Zhang, X. P. Wang, B. Liu, Y. Y. Zhang, X. S. Lv, J. X. Zhou and H. J. Yu, Origin of the difference in thermal conductivity and anharmonic phonon scattering between LiNbO3 and LiTaO3, CrystEngComm, 2021, 23, 8572–8578 RSC.
  8. Y. N. Xiao, P. F. Sui, Y. C. Zhao, J. Ni, S. Meng and Z. H. Dai, Low lattice thermal conductivity of hydride-based cubic antiperovskites A3HB (A = Li, Na; B = S, Se, Te) with higher-order anharmonicity correction, Int. J. Energy Res., 2022, 46, 13687–13697 CrossRef CAS.
  9. Y. Liu, L. H. Ni, G. Xu, C. L. Song, G. R. Han and Y. Zheng, Phase transition in PbTiO3 under pressure studied by the first-principles method, Physica B, 2008, 403, 3863–3866 CrossRef CAS.
  10. D. Han, C. B. Feng, M. H. Du, T. Zhang, S. Z. Wang, G. Tang, T. Bein and H. Ebert, Design of high-performance lead-free quaternary antiperovskites for photovoltaics via ion type inversion and anion ordering, J. Am. Chem. Soc., 2021, 143, 12369–12379 CrossRef CAS PubMed.
  11. J. Chen, X. Cai, D. Yang, D. Song, J. Wang, J. Jiang, A. Ma, S. Lv, M. Z. Hu and C. Ni, Recent progress in stabilizing hybrid perovskites for solar cell applications, J. Power Sources, 2017, 355, 98–133 CrossRef CAS.
  12. M. J. Fang, C. W. Tsao and Y. J. Hsu, Semiconductor nanoheterostructures for photoconversion applications, J. Phys. D: Appl. Phys., 2020, 53, 143001–143026 CrossRef CAS.
  13. Y. H. Chiu, T. H. Lai, M. Y. Kuo, P. Y. Hsieh and Y. J. Hsu, Photoelectrochemical cells for solar hydrogen production: Challenges and opportunities, APL Mater., 2019, 7, 080901–080911 CrossRef.
  14. N. Erum, R. Sharma, A. Kumar, J. Nanda, R. Ghodhbani and N. Kamolova, Probing the physical properties of novel lead-free double perovskites A2NaAsCl6 (A = K and Rb): A first-principles approach, New J. Chem., 2025, 49, 11010–11021 RSC.
  15. M. Z. Rehman, S. A. M. Abdelmohsen, E. A. Mahmoud, M. U. Saeed, M. Idrees, M. Shafiq, B. Amin and Y. Saeed, First principles study of structural, electronic, elastic and optical properties of Cs2LiTlBr6 and Cs2NaTlBr6, Mater. Sci. Semicond. Process., 2022, 151, 106993 CrossRef.
  16. U. Rani, P. K. Kamlesh, R. Agarwal, J. Kumari and A. S. Verma, Electronic and thermo-physical properties of double antiperovskites X6SOA2 (X = Na, K and A = Cl, Br, I): a non-toxic and efficient energy storage materials, Int. J. Quantum Chem., 2021, 121, 26759 CrossRef.
  17. Y. Yu, Z. Wang and G. S. Shao, Theoretical design of double anti-perovskite Na6SOI2 as a super-fast ion conductor for solid Na+ ion batteries, J. Mater. Chem. A, 2018, 6, 19843–19852 RSC.
  18. M. Mebrouki, T. Ouahrani and Y. O. Ciftci, Unraveling thermal and dynamical properties of the cubic BaVO3 perovskite from first-principles calculation, Int. J. Thermophys., 2016, 37, 71–80 CrossRef.
  19. K. Djebari, A. Dahani, M. Djermouni, K. Dine, A. Cherif, O. Arbouche, A. Zaoui and S. Kacimi, Spontaneous polarization study in A3+B4+(O2N)7− and A2+B5+(O2N)7− perovskite-type oxynitrides: a first principles investigation, Appl. Phys. A, 2022, 128, 398–410 CrossRef CAS.
  20. Q. Gao, R. Y. Ma, L. Li, H. H. Xie, J. B. Deng and X. R. Hu, First-principle study on compensated half metallic double perovskite compound Ba2PrVO6, J. Supercond. Novel Magn., 2017, 30, 545–554 CrossRef CAS.
  21. H. X. Gao, D.-Y. Hu, T.-Y. Tang, Q.-Q. Liang, Q. Dai, S.-S. Wu and Y.-L. Tang, First-principles study on the structural, electronic, elastic, optical and thermodynamic properties of double antiperovskites X6BiSbN2 (X = Mg, Ca, Sr), J. Phys. Chem. Solids, 2004, 187, 111859 CrossRef.
  22. P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz, WIEN2k; an Augmented Plane Wave Plus Local Orbital Program for Calculating Crystal Properties, Vienna University of Technology, Austria, 2001 Search PubMed.
  23. Z. Wu and R. E. Cohen, More accurate generalized gradient approximation for solids, Phys. Rev. B:Condens. Matter Mater. Phys., 2006, 73, 235116 CrossRef.
  24. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  25. E. Engel and S. H. Vosko, Exact exchange-only potentials and the virial relation as microscopic criteria for generalized gradient approximations, Phys. Rev. B:Condens. Matter Mater. Phys., 1993, 47, 13164 CrossRef CAS PubMed.
  26. Z. Ali, I. Ahmed, I. Khan and B. Amin, Electronic structure of cubic perovskite SnTaO3, Intermetallics, 2012, 31, 287 CrossRef CAS.
  27. G. Murtaza, I. Ahmed, B. Amin and M. Zahid, Investigation of structural and optoelectronic properties of BaThO3, Opt. Mater., 2011, 33, 553 CrossRef CAS.
  28. H. Jiang, Band gaps from the Tran-Blaha modified Becke-Johnson approach: A systematic investigation, J. Chem. Phys., 2013, 138, 134115 CrossRef PubMed.
  29. J. A. Camargo-Martinez and R. Baquero, Performance of the modified Becke-Johnson potential for semiconductors, Phys. Rev. B:Condens. Matter Mater. Phys., 2008, 86, 195106 CrossRef.
  30. R. A. Jishi, O. B. Ta and A. A. Sharif, Modeling of lead halide perovskites for photovoltaic applications, J. Phys. Chem. C, 2014, 118, 28344–28349 CrossRef CAS.
  31. R. A. Jishi, Modifed Becke-Johnson exchange potential: improved modelling of lead halides for solar cell applications, AIMS Mater. Sci., 2016, 3, 149–159 CAS.
  32. Y. Saeed, B. Amin, H. Khalil, F. Rehman, H. Ali, M. I. Khan, A. Mahmood and M. Shafiq, Cs2NaGaBr6: a new lead-free and direct bandgap halide double perovskite, RSC Adv., 2020, 10, 17444–17451 RSC.
  33. C. Y. Wang, S. R. Li, S. F. Wang, P. X. Zhao, R. S. Zhuo and B. Y. Yu, Effect of biaxial Pre-proof strain on monolayer MoS2 and its vacancy defect system: a first-principles study, J. Solid State Chem., 2023, 317, 123711 CrossRef CAS.
  34. M. A. Green, Intrinsic concentration, effective densities of states, and effective mass in silicon, J. Appl. Phys., 1990, 67, 2944–2954 CrossRef CAS.
  35. A. A. Emery and C. Wolverton, High-throughput DFT calculations of formation energy, stability and oxygen vacancy formation energy of ABO3 perovskites, Sci. Data, 2017, 4, 170153 CrossRef CAS PubMed.
  36. C. Li, X. Lu, W. Ding, L. Feng, Y. Gao and Z. Guo, Formability of ABX3 (X = F, Cl, Br, I) halide perovskites, Acta Crystallogr., Sect. B:Struct. Sci., 2008, 64, 702–707 CrossRef CAS PubMed.
  37. J. Liang, C. Wang, Y. Wang, Z. Xu, Z. Lu, Y. Ma, H. Zhu, Y. Hu, C. Xiao, X. Yi, G. Zhu, H. Lv, L. Ma, T. Chen, Z. Tie, Z. Jin and J. Liu, All-inorganic perovskite solar cells, J. Am. Chem. Soc., 2016, 138, 15829–15832 CrossRef CAS PubMed.
  38. Z. Abbas, N. Jabeen, A. Hussain, F. Kabir, T. Alshahrani, H. H. Raza, S. Muhammad, S. Azam and I. Gorczyca, Effect of Nb, Ta and V replacements on electronic, optical and elastic properties of NbCu3Se4: a GGA+U study, J. Solid State Chem., 2021, 301, 122338 CrossRef CAS.
  39. P. Mondal, N. A. Shahed, S. Khanom, M. K. Hossain and F. Ahmed, Effect of oxygen defects in tuning the half-metallic and optical nature of Sr2CoXO6 (X = Mo and W): a first-principles study, Mater. Chem. Phys., 2022, 277, 125429 CrossRef CAS.
  40. L. D. Whalley, J. M. Frost, Y. K. Jung and A. Walsh, Perspective: Theory and simulation of hybrid halide perovskites, J. Chem. Phys., 2017, 146, 220901 CrossRef PubMed.
  41. D. Y. Hu, X. H. Zhao, T. Y. Tang, L. Li, L. K. Gao and Y. L. Tang, First-principles calculations to investigate structural, elastic, electronic and optical properties of lead-free perovskite derivatives Cs2SeX6 (X = Cl, Br, I)” Optic, Mater, 2021, 119, 111316 CAS.
  42. L. Y. Gong, P. Zhang, Q. Chen, Z. H. Lou, J. Xu and F. Gao, First principles study of structure and property of Nb5+ doped SrTiO3”, Acta Phys. Sin., 2021, 70, 227101 CrossRef.
  43. M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light, 7th (expanded) edn, Cambridge University Press, Cambridge, 1999 Search PubMed.
  44. M. Fox, Optical Properties of Solids, 2nd edn, Oxford University Press, 2010 Search PubMed.
  45. W. F. Goh and W. E. Pickett, Topological and thermoelectric properties of double antiperovskite pnictides, J. Phys.:Condens. Matter, 2020, 32, 345502 CrossRef CAS PubMed.
  46. C. Y. Wang, S. R. Li, S. F. Wang, P. X. Zhao, R. S. Zhuo and B. Y. Yu, Effect of biaxial[110] strain on monolayer MoS2 and its vacancy defect system: a first-principles study, J. Solid State Chem., 2023, 317, 123711 CrossRef CAS.
  47. Z. Abbas, N. Jabeen, A. Hussain, F. Kabir, T. Alshahrani, H. H. Raza, S. Muhammad, S. Azam and I. Gorczyca, Effect of Nb, Ta and V replacements on electronic, optical and elastic properties of NbCu3Se4: a GGA+U study” J, Solid State Chem., 2021, 301, 122338 CrossRef CAS.
  48. P. Mondal, N. A. Shahed, S. Khanom, M. K. Hossain and F. Ahmed, Effect of oxygen defects in tuning the half-metallic and optical nature of Sr2CoXO6 (X = Mo and W): a first-principles study, Mater. Chem. Phys., 2022, 277, 125429 CrossRef CAS.
  49. S. Berri, Theoretical analysis of the structural, electronic and optical properties of tetragonal Sr2GaSbO6, Chin. J. Phys., 2017, 55, 2476–2483 CrossRef CAS.
  50. M. Maglione and M. A. Subramanian, Dielectric and polarization experiments in high loss dielectrics: A word of caution, Appl. Phys. Lett., 2008, 93, 032902 CrossRef.
  51. L. Jin, F. Li and S. Zhang, Decoding the fingerprint of ferroelectric loops: comprehension of the material properties and structures, J. Am. Ceram. Soc., 2014, 97, 1–27 CrossRef CAS.
  52. M. Barhoumi, I. Said, N. Sfina, N. K. Al-Saleem and T. Ghrib, A DFT study of the electronic and optical properties of four 2D thin films, Mater. Chem. Phys., 2022, 286, 126158 CrossRef CAS.
  53. P. Rani, G. S. Dubey and V. K. Jindal, DFT study of optical properties of pure and doped graphene, Physica E, 2014, 62, 28–35 CrossRef CAS.
  54. K. Maeda and K. Domen, Photocatalytic water splitting: recent progress and future challenges, J. Phys. Chem. Lett., 2021, 1, 2655–2661 CrossRef.
  55. R. M. Navarro Yerga, M. C. Alvarez Galvan, F. Del Valle, J. A. Villoria de la Mano and J. L. Fierro, Water splitting on semiconductor catalysts under visible-light irradiation, ChemSusChem, 2009, 2, 471–485 CrossRef CAS PubMed.
  56. F. E. Osterloh, Inorganic materials as catalysts for photochemical splitting of water, Chem. Mater., 2008, 20, 35–54 CrossRef CAS.
  57. X. Hu, G. Li and J. C. Yu, Design, fabrication and modification of nanostructured semiconductor materials for environmental and energy applications, Langmuir, 2010, 26, 3031–3039 CrossRef CAS PubMed.
  58. A. Kudo, Photocatalysis and solar hydrogen production, Pure Appl. Chem., 2007, 79, 1917–1927 CrossRef CAS.
  59. M. U. Saeed, T. Usman, S. Pervaiz, Z. Ali, Y. M. Alanazi, A.-U.-R. Bacha and Y. Saeed, DFT analysis of alkali scandium sulfides AScS2 (A = K, Rb): Unveiling structural, electronic, and optical properties for enhanced photocatalysis, ACS Appl. Opt. Mater., 2024, 2, 1955–1964 CrossRef CAS.
  60. J. Liu, X. Fu, S. Chen and Y. Zhu, Electronic structure and optical properties of Ag3PO4 photocatalyst calculated by hybrid density functional method, Appl. Phys. Lett., 2011, 99, 191903 CrossRef.
  61. M. Saeed, I. U. Haq, A. S. Saleemi, S. U. Rehman, B. U. Haq, A. R. Chaudhry and I. Khan, First-principles prediction of the ground-state crystal structure of double-perovskite halides Cs2AgCrX6 (X = Cl, Br, and I), J. Phys. Chem. Solids, 2022, 160, 110302 CrossRef CAS.
  62. T. A. Pham, D. Lee, E. Schwegler and G. Galli, Interfacial effects on the band edges of functionalized Si surfaces in liquid water, J. Am. Chem. Soc., 2014, 136, 17071–17077 CrossRef CAS PubMed.
  63. X. Li, J. Yu, J. Low, Y. Fang, J. Xiao and X. Chen, Engineering heterogeneous semiconductors for solar water splitting, J. Mater. Chem. A, 2015, 3, 2485–2534 RSC.
  64. M. U. Saeed, T. Usman, S. M. Ali, S. Pervaiz, H. O. Elansary, I. M. Moussa, M. A. El-Sheikh, A.-U.-R. Bacha and Y. Saeed, Exploring electronic, optical, elastic, and photocatalytic properties in new double perovskites Cs2TlSbCl6 and Cs2TlBiCl6 materials: A GGA+SOC and hybrid functional study, J. Mater. Res., 2024, 39, 3155–3165 CrossRef CAS.
  65. J.-C. Zheng, Recent advances on thermoelectric materials, Front. Phys., 2008, 3, 269 CrossRef.
  66. D. T. Morelli and G. A. Slack, High lattice thermal conductivity solids, in High Thermal Conductivity Materials, Springer, 2006, vol. 1, pp. 37–68 Search PubMed.
  67. G. L. Menaria, U. Rani, P. K. Kamlesh, R. Singh, M. Rani, N. Singh, D. C. Sharma and A. S. Verma, Electro-optic and transport properties with stability parameters of cubic KMgX (X = P, As, Sb, and Bi) half-Heusler materials: Appropriate for green energy applications, Mod. Phys. Lett. B, 2024, 245, 2450283 CrossRef.
  68. S. Y. Yue, T. Xu and B. Liao, Ultralow thermal conductivity in a two-dimensional material due to surface-enhanced resonant bonding, Mater. Today Phys., 2018, 7, 89–95 CrossRef.
  69. G. L. Menaria, U. Rani, P. K. Kamlesh, M. Rani, N. Singh, D. C. Sharma and A. S. Verma, Comprehensive theoretical investigation of NaAlX (X = C, Si, and Ge) half-Heusler compounds: Unveiling the multifaceted properties for advanced applications, Int. J. Mod. Phys. B, 2024, 255, 2550052 Search PubMed.
  70. N. V. Prabhu and K. A. Sharp, Heat capacity in proteins, Annu. Rev. Phys. Chem., 2005, 56, 521 CrossRef CAS PubMed.
  71. Q. Gao, R. Y. Ma, L. Li, H. H. Xie, J. B. Deng and X. R. Hu, First-principle study on ompensated half metallic double perovskite compound Ba2PrVO6, J. Supercond. Novel Magn., 2017, 30, 545–554 CrossRef CAS.
  72. Z. Huang, C. He, X. Qi, H. Yang, W. Liu, X. Wei, X. Peng and J. Zhong, Band structure engineering of monolayer MoS2 on h-BN: first-principles calculations, J. Phys. D: Appl. Phys., 2014, 47, 075301 CrossRef CAS.
  73. R. M. Abraham, J. alvarez-Muniz, C. A. Arguelles, A. Ariga, T. Bostan, M. Bustamante and A. Camming, Tau neutrinos in the next decade form GeV to EeV, J. Phys. G: Nucl. Part. Phys., 2022, 49, 110501 CrossRef CAS.
  74. A. A. Baloch, M. I. Hossain, N. Tabet and F. H. Alharbi, Practical efficiency limit of methylammonium lead iodide perovskite solar cells, J. Phys. Chem., 2018, 9, 426–434 CAS.
  75. J. Li and H. Chen, First-principles investigation of the mechanical and thermodynamic properties of the antiperovskite centrosymmetric platinum-based superconductors APt3P (A = Ca, Sr, La and Ce), Physica B, 2020, 577, 411791 CrossRef CAS.
  76. M. Y. Rudysh, M. Piasecki, G. L. Myronchuk, P. A. Shchepanskyi, V. Y. Stadnyk, O. R. Onufriv and M. G. Brika, AgGaTe2 The thermoelectric and solar cell material: structure, electronic, optical, elastic and vibrational features, Infrared Phys. Technol., 2020, 111, 103476 CrossRef CAS.
  77. B. Asghar, M. U. Saeed, L. Sajid, S. Pervaiz, S. Khan, Z. Ali, A. Khan, Z. Ullah, H. O. Elansary and Y. Saeed, Ab-initio study of electronic, elastic, and optical properties of double antiperovskite, Li6AgBiX2 (X = Cl, Br, I), Chem. Phys. Lett., 2025, 844, 141900 CrossRef.
  78. B. T. Wang, W. X. Zhang and W. D. Li, Mechanics, Lattice dynamics, and chemical bonding in ZrB2 and ZrB2 from first-principles calculations, Sci. Adv. Mater., 2013, 5, 1916–1921 CrossRef CAS.
  79. X. H. Zhao, Y. L. Tang, T. Y. Tang, X. F. Diao, L. K. Gao, Q. Xie, B. Shi, L. Yuan and L. M. Lu, Study on mechanical, electronic and optical properties of Pb-free double halide perovskites In2TiX6 (X = Cl, Br, I) for solar cells based on first-principles, Mater. Today Commun., 2021, 26, 102180 CrossRef CAS.
  80. L. K. Gao, X. H. Zhao, X. F. Diao, T. Y. Tang and Y. L. Tang, First-principles study of photoelectric properties of CsSnBr3 under hydrostatic pressure, Acta Phys. Sin., 2021, 70, 158801 CrossRef.
  81. J. Li and H. Chen, First-principles investigation of the mechanical and thermodynamic properties of the antiperovskite centrosymmetric platinum-based superconductors APt3P (A = Ca, Sr, La and Ce), Physica B, 2020, 577, 411791 CrossRef CAS.
  82. S. I. Ranganathan and M. Ostoja-Starzewski, Universal elastic anisotropy index, Phys. Rev. Lett., 2008, 101, 055504 CrossRef PubMed.
  83. S. F. Pugh, XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals, Philos. Mag., 1954, 45, 823 CAS.
  84. M. A. Rehman, J. U. Rehman and M. B. Tahir, Density functional theory study of structural, electronic, optical, mechanical, and thermodynamic properties of halide double perovskites Cs2AgBiX6 (X = Cl, Br, I) for photovoltaic applications, J. Phys. Chem. Solids, 2023, 181, 111443 CrossRef CAS.
  85. M. Hattab and K. Rahmani, Thermodynamic, optical, and morphological studies of the Cs2AgBiX6 double perovskites (X = Cl, Br, and I): Insights from DFT study, J. Alloys Compd., 2023, 960, 170650 CrossRef.
  86. S. A. Dar, R. Sharma, V. Srivastava and U. K. Sakalle, Investigation on the electronic structure, optical, elastic, mechanical, thermodynamic and thermoelectric properties of wide band gap semiconductor double perovskite Ba2InTaO6, RSC Adv., 2019, 9, 9522–9532 RSC.

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