Open Access Article
Murtaza Shahab†
a,
Muhammad Awais Jehangir†d,
Muhammad Nomana,
Saad Alshammari*b,
Arafa A. Yagobc and
G. Murtaza
*a
aMaterials Modeling Lab, Department of Physics, Islamia College Peshawar, Pakistan. E-mail: murtaza@icp.edu.pk
bDepartment of Mechanical and Industrial Engineering, College of Engineering, Majmaah University, Al Majmaah 11952, Saudi Arabia
cDepartment of Physical Sciences, Physics Division, College of Science, Jazan University, P. O. Box 114, 45142, Jazan, Saudi Arabia
dInstitute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu, Sichuan 611731, China
First published on 5th November 2025
Double perovskites have garnered significant attention as promising alternatives for sustainable energy solutions owing to their structural versatility and potential for integration into optoelectronic technologies. Halide double perovskites K2TlAsZ6 (Z = F, Cl, Br, and I) were systematically studied using density functional theory to assess their potential for optoelectronic and thermoelectric applications. Calculations were performed using the FP-LAPW + lo method, confirming structural and thermodynamic stability via formation energy, and the Goldschmidt's tolerance factor. The electronic structure calculations using the TB-mBJ + SOC potential revealed direct bandgaps ranging from 3.25 eV (Z = F) to 0.37 eV (Z = I), with significant UV absorption observed in the optical spectra. Thermoelectric performance versus chemical potential, evaluated via the Boltzmann transport theory, showed promising ZT values approaching 1.0 at 1000 K. Additionally, negative Gibbs free energy and increasing entropy with temperature indicate good thermal stability. These results suggest that K2TlAsZ6 compounds are promising materials for next-generation optoelectronic and thermoelectric devices.
Halide perovskites crystallize in the ABX3 form, while double perovskites adopt A2BB′X6, where A is a monovalent cation, B and B′ are metal cations of different oxidation states, and X is a halide anion.5 Lead-free double perovskites (DPs) display distinctive electronic and optical characteristics, making them promising for renewable energy applications.6 For instance, Rb2AgAsX6 (X = Br, I, and Cl) shows tunable band gaps via halide substitution,7 and Cs2ScTlX6 (X = Cl, Br, and I) exhibits mechanical stability in cubic symmetry as confirmed by elastic constants and tolerance factors.8 Cs2LiYCl6, synthesized for thermal neutron detection, demonstrates potential as a scintillator material.9 Computational investigations reveal favorable optoelectronic behavior in Rb2AgBiI6,10 while X2AgBiI6 (X = Cs, K, and Rb) offers high thermoelectric figures of merit with advantageous optical properties.11 Cs2SnI6 has been examined as a lead-free absorber for solar applications,12 and Cs2BiAgX6 (X = Br and Cl) prepared via solid-solution synthesis exhibits direct band-gap behavior with modified transport characteristics.13
Within this class, potassium-based perovskites are emerging as particularly promising due to their stability and favorable electromagnetic, thermoelectric, and optoelectronic properties.14 Potassium doping has enabled the fabrication of hysteresis-free solar cells,15 while high-efficiency, low-hysteresis devices have been reported using potassium-based perovskites.16 Additionally, potassium thiocyanate interlayers have been proposed for eco-friendly indoor and outdoor photovoltaics.17 These findings underscore the potential of K-based halides for large-scale deployment in sustainable energy technologies,18 motivating continued research into novel potassium-based perovskite compositions for advanced solar energy conversion and storage applications.
Munir et al. (2025) have thoroughly examined the structural, electronic, mechanical, optical, and thermoelectric properties of K2TlAsX6 (X = Cl and Br) double perovskites, employing first-principles calculations within the framework of density functional theory (DFT). The electronic band structure was computed using the mBJ exchange potential, both with and without the inclusion of spin orbit coupling (SOC) effects, to account for relativistic interactions. Structural stability was confirmed through negative formation energies, optimized lattice parameters, positive phonon frequencies, and tolerance and octahedral factors within the permissible range for stable perovskite formation. Elastic constants were determined and the associated mechanical parameters, including elastic anisotropy, Pugh's ratio, Poisson's ratio, and Cauchy pressure, were derived to assess the ductility and mechanical anisotropy. Both compounds exhibit direct band gaps, with K2TlAsCl6 showing gaps of 2.22 eV (modified Becke–Johnson, mBJ) and 1.59 eV (mBJ + SOC), while K2TlAsBr6 presents gaps of 1.97 eV and 1.51 eV, respectively. The optical properties reveal strong absorption in the visible and ultraviolet regions, suggesting potential for optoelectronic integration. Moreover, electronic transport analysis indicates high electrical conductivity, favorable thermoelectric figure of merit (ZT) and low lattice thermal conductivity, making these halide perovskites promising candidates for energy-efficient and green optoelectronic applications.19 Although K2TlAsZ6 are designed as lead-free alternatives, it should be noted that these systems contain thallium (Tl) and arsenic (As), both of which are toxic and pose environmental and health risks if improperly handled. Therefore, the term “lead-free” in this context refers specifically to the absence of lead, a well-known hazardous element in conventional perovskites, rather than indicating complete non-toxicity.
The aim of the present work is to assess the structural, optoelectronic, thermoelectric, and thermodynamic properties of K2TlAsZ6 (Z = F, Cl, Br, and I). Our findings are intended to support further theoretical and practical studies of K2TlAsZ6 (Z = F, Cl, Br, and I) halide double perovskites, with an emphasis on their possible applications in thermoelectric and optoelectronic devices.
![]() | (1) |
| ΔHF = E(K)aTlbAsc(F/Cl/Br/I)d − aEKb − bETl − cEAs − dE(F/Cl/Br/I) | (2) |
| ECoh = [aEK + bETl + cEAs + dE(F/Cl/Br/I)] − E(K)aTlbAsc(F/Cl/Br/I)d | (3) |
| Compounds | K2TlAsF6 | K2TlAsCl6 | K2TlAsBr6 | K2TlAsI6 |
|---|---|---|---|---|
| a RefA.19 | ||||
| a0 (Å) | 9.38 | 10.83 | 11.31 | 12.04 |
| 11.00A | 11.53A | |||
| V0 [a.u.]3 | 1394.21 | 2144.88 | 2443.12 | 2945.71 |
| 2586.71A | 2588.00A | |||
| B (GPa) | 43.48 | 26.96 | 23.51 | 19.09 |
| 24.00A | 20.31A | |||
| B' | 5.00 | 5.00 | 5.00 | 5.00 |
| 4.86A | 4.83A | |||
| E0 [Ry.] | −48707.17 | −53021.15 | −78750.39 | −132884.23 |
| −53047.39A | −78789.00A | |||
| τF | 1.0 | 0.96 | 0.95 | 0.93 |
| 0.88A | 0.87A | |||
| ΔHF (eV/atom) | −1.90 | −1.41 | −1.10 | −0.67 |
| −1.92A | −1.61A | |||
| ECohesive | 1.90 | 1.41 | 1.10 | 0.67 |
| ρ (g cm−3) | 3.79 | 2.98 | 3.84 | 4.26 |
![]() | ||
| Fig. 1 (aand b) Tolerance factor (TF), formation (ΔHF) and cohesive energies of K2TlAsZ6 (Z = F, Cl, Br, and I). | ||
m#225 (see Fig. 2a). We utilized PBEsol-GGA to optimize halide DPs (see Fig. 2b–f) to determine the fundamental state lattice parameters, applying Birch Murnaghan's equation of states.27 We determined the fundamental state lattice constant a0 (Å) and bulk modulus B (GPa) by volume optimization, as demonstrated in the following Table 1. The values of a0 show an upsurge from K2TlAsF6 to K2TlAsI6, which can be explained by the increase in atomic radii from F to I. The increase in distance between the atoms is a result of the larger cation size, which subsequently leads to a decrease in the density and strength of the compounds. Therefore, a drop in the bulk modulus B0 is observed by substituting F with Cl, Br and I.28
![]() | ||
| Fig. 2 (a–f) Crystal structure, E–V parabolic curves, and unit cell volume and energy curves with percentage change in its volume for K2TlAsZ6 (Z = F, Cl, Br, and I). | ||
For evaluating the electronic characteristics of the compounds being studied, the TDOS and PDOS, utilizing TB-mBJ, are shown in Fig. 4a–d. The total contribution of each state is shown as TDOS throughout the occupied and unoccupied states, much like the band structure. The interband transitions and hybridization are entirely the result of the occupied state electrons, whose roles in the valence bands are seen in Fig. 4a–d. (F/Cl/Br/I)-5p makes the largest contribution, with As-3p making a smaller one in the semi-core regions of the occupied state, where electrons migrate to the unoccupied state. In the conduction band As-3p states make a larger contribution. The band gap values for K2TlAsZ6 (Z = F, Cl, Br, I) are determined in the range 3.25 eV to 0.37 eV using TB-mBJ methods. The investigated materials exhibit a direct band gap, with both the valence band maximum (VBM) and conduction band minimum (CBM) located at the Γ high-symmetry point in the Brillouin zone. These materials are therefore ideally suited for photovoltaic and optoelectronic applications, including tandem solar cells,30 photodetectors that detect UV light,31 sensors that detect infrared,32 as well as LEDs.33
![]() | (4) |
In this equation, m is the mass, ‘e’ is the electron's charge, and “ω” is the electromagnetic radiation's angular frequency. Conversely, optical transitions are indicated by the dipole matrix element Mck(k) = (uck|e∇|vk).28 The real part of the dielectric function is found using the Kramers–Kronig equation, as explained in the next section:
![]() | (5) |
The real component of the dielectric function is represented by ε1(ω). The resonant frequencies for K2TlAsZ6 (Z = F, Cl, Br, and I) correspond to energy values ranging from 1.76 eV to 3.20 eV, where peak scattering of light occurs (Fig. 5a). The peaks exhibit a rapid decline to their lowest levels after achieving resonance. The zero-frequency limit of the real part of the dielectric function, ε1(0), which corresponds to the electronic contribution to the static dielectric constant, is a critical parameter in the optical spectrum. The calculated ε1(0) values for K2TlAsZ6 (Z = F, Cl, Br, and I) are summarized in Table 2.35,36 An inverse dependence is observed between the band gap energy (Eg) and the static dielectric constant ε1(0). For the investigated K2TlAsZ6 (Z = F, Cl, Br, and I) compounds, ε1(0) exhibits an initial increase from zero frequency, attains a maximum, and then gradually decreases. With increasing photon energy, additional distinct peaks emerge, as depicted in Fig. 5a. These spectral features, or kinks, are primarily located within the photon energy range of 1.76 eV to 3.20 eV. The maximum values of the real part of the dielectric function ε1(ω) are found to be 3.22 at 4.93 eV for K2TlAsF6, 4.80 at 2.87 eV for K2TlAsCl6, 6.42 at 2.35 eV for K2TlAsBr6, and 9.90 at 1.72 eV for K2TlAsI6. The Penn model32 has been employed to estimate ε1(0) and Eg, subject to the criterion for semiconducting materials, i.e., ε1(0) >1.
![]() | (6) |
The imaginary part of the dielectric function, ε2(0), is strongly influenced by the electronic band structure and plays a critical role in determining the optical absorption characteristics of the material. It provides insight into the interband electronic transitions. As the photon energy increases, transitions occur from the valence band, predominantly composed of halogen (F/Cl/Br/I)-5p orbitals, to the conduction band, which mainly consists of As-3p states. This transition consistently gives rise to the first prominent absorption peak observed across all investigated compounds. As illustrated in Fig. 5b, the second peak is also attributed to similar interband transitions. The calculated value of ε2(ω) is determined from Fig. 5b, showing that all photons with energy E (eV) beneath the matching energy gaps (Eg) of K2TlAsZ6 (Z = F, Cl, Br, and I) have a constant value of zero. The threshold points in the spectra of ε2(ω) are 3.80 eV, 1.74 eV, 1.11 eV, and 0.67 eV for K2TlAsF6, K2TlAsCl6, K2TlAsBr6, and K2TlAsI6, respectively, as compared to the highest peak observed at 3.9 eV for K2TlAsCl6 and 3.1 eV for K2TlAsBr6.19 A strong correlation is observed between the electronic band gaps and the corresponding peaks in the optical spectra. The calculated maximum values of the imaginary part of the dielectric function, ε2(ω), for the K2TlAsZ6 (Z = F, Cl, Br, and I) compounds are 2.99 at 5.51 eV for K2TlAsF6, 3.73 at 3.30 eV for K2TlAsCl6, 5.07 at 3.14 eV for K2TlAsBr6, and 8.63 at 2.40 eV for K2TlAsI6. Notably, the most prominent absorption features appear near 3.1 eV and 2.7 eV for K2TlAsCl6 and K2TlAsBr6, respectively, indicating a direct relationship with their optical transition energies.19 According to our accurate observations, both halides have remarkable visible spectrum photon absorption properties, which are essential for the efficiency of opto-electronic systems.
For optical components to be successfully integrated into devices like photonic crystals, waveguides, solar cells, and detection devices, an extensive knowledge of the refractive index is essential. Fig. 5c depicts the variation in the refractive index for K2TlAsZ6 (Z = F, Cl, Br, and I) as a function of the photon energy. The calculated static refractive index values, n (0), for each composition are listed in Table 2 in comparison to the data reported by Junaid Munir et al. (2025).19 As illustrated in Fig. 5c, the refractive indices of the investigated compounds exhibit an increasing trend from zero photon energy (0 eV) up to their respective maximum values of 1.87 (5.04 eV) for K2TlAsF6, 2.25 (3.08 eV) for K2TlAsCl6, 2.62 (2.32 eV) for K2TlAsBr6, and 3.24 (1.78 eV) for K2TlAsI6 as compared to the highest peak observed at 3.38 eV for K2TlAsCl6 and 2.79 eV for K2TlAsBr6, respectively.19 Following the attainment of its peak value, the refractive index for each compound exhibits a decline across specific energy intervals. This behavior suggests a transition in the optical response of the medium from linear to nonlinear, thereby imparting superluminal characteristics to the material.33 The calculated refractive index values for the double perovskites K2TlAsZ6 (Z = F, Cl, Br, and I) are found to be equal to or greater than unity, indicating that these materials exhibit linear optical behavior. As depicted in Fig. 5c (left panel), the group velocity (Vg = c/n) remains positive, with no indication of a transition toward negative energy. A refractive index greater than one implies a decrease in photon velocity upon entering the medium, attributed to the interaction between incident photons and the electronic structure of the material, thereby inducing a temporal delay in photon propagation. In addition, the absorption coefficient k(ω), shown in Fig. 5d, characterizes the material's capacity to attenuate incident light. The spectral profile of k(ω) follows a similar trend to that of the imaginary component of the dielectric function, ε2(ω). The peak values of k(ω) are identified as 0.99 (5.59 eV), 1.06 (4.39 eV), 1.29 (3.33 eV), and 1.72 (2.54 eV) for K2TlAsZ6 with Z = F, Cl, Br, and I, respectively.
Electronic conduction is the cause of the optical conductivity σ(ω). As displayed in Fig. 5e, the first peak values of σ(ω) are 2679.7 [Ω cm]−1 at 9.51 eV, 4578.4 [Ω cm]−1 at 12.88 eV, 4768.5 [Ω cm]−1 at 9.34 eV, and 5351.3 [Ω cm]−1 at 7.66 eV for K2TlAsZ6 (Z = F, Cl, Br, and I), whereas the highest peak ever seen is 5351.3 [Ω cm]−1 at 7.66 eV for K2TlAsI6. In the ultraviolet light spectrum, the σ(ω) is expected to exhibit a significant rise in the 7.6 to 13 eV region, which is essential for optoelectronic applications. Consequently, K2TlAsI6 is the best option for these applications because of its exceptional values.37 Fig. 5f shows the absorption coefficient α(ω) of K2TlAsZ6 (Z = F, Cl, Br, and I). The extent to which light may enter a material before it is fully absorbed is indicated by the symbol α(ω). Each substance has a threshold at which it stops absorbing light. When photons collide with valence-state electrons, photon–electron interactions take place, which, once the threshold value is exceeded, produce very efficient light absorption. The formula shows a significant decrease in energy.38–40
| α = 4πk/λ | (7) |
The materials K2TlAsZ6 (Z = F, Cl, Br, and I) display an upsurge in absorbance from Eg (eV), reaching utmost intensities between 8 to 12 eV. The cutoff value of the absorption edge is established in the following manner: 3.60 eV for K2TlAsF6; 1.82 eV for K2TlAsCl6; 1.30 eV for K2TlAsBr6; and 0.99 eV for K2TlAsI6 as compared to the data at 2.08 eV for K2TlAsCl6 and 1.69 eV for K2TlAsBr6, respectively.19 The apex standards for the α(ω) are 85.26 cm−1 (9.61 eV) for K2TlAsF6; 141.04 cm−1 (12.01 eV) for K2TlAsCl6; 155.01 cm−1 (11.90 eV) for K2TlAsBr6; and 140.59 cm−1 (9.59 eV) for K2TlAsI6, as compared to the highest peak observed at 4.01 eV for K2TlAsCl6 and 1.87 eV for K2TlAsBr6, respectively.19 The double perovskites under analysis show a notable rise in optical absorption in the blue (UV) shift. The in-depth examination of the spectral features confirms that K2TlAsZ6 (Z = F, Cl, Br, and I) has excellent absorption qualities, making it ideal for optoelectronic applications.
The energy emitted per unit area may be expressed as L(ω). Dispersion, thermal impacts, and plasmonic interaction are the principal factors influencing the optical loss L(ω). Fig. 5g exemplifies the reaction of the energy loss function L(ω). The most significant optical loss for K2TlAsZ6 (Z = F, Cl, Br, and I) is documented as 1.98 (10.92 eV), 0.77 (7.79 eV), 0.62 (5.97 eV), and 0.63 (4.77 eV), as compared to the highest peak observed at 7.8 eV for K2TlAsCl6 and 5.8 eV for K2TlAsBr6, respectively.19 Fig. 5g illustrates that the L(ω) reaches its minimum as the reflectivity approaches its maximum, and vice versa. In the purview of visible light absorption, the L(ω) decreases. The findings demonstrate that materials K2TlAsZ6 (Z = F, Cl, Br, and I) are suitable for photonic and solar power applications.
The surface reflectivity of a solid can be assessed through ε(ω), which measures the quantity of incident sunlight that is reflected.41 It may be represented as:
![]() | (8) |
Fig. 5h illustrates the depicted reflectivities R(ω) of DP K2TlAsZ6 (Z = F, Cl, Br, and I). The R(ω) of the DPs vary at diverse light energies (in eV): 24.10% (10.70 eV) for K2TlAsF6, 23.19% (12.23 eV) for K2TlAsCl6, 30.61% (12.09 eV) for K2TlAsBr6, and 39.65% (12.47 eV) for K2TlAsI6, respectively.
![]() | ||
| Fig. 6 (a–e) Thermoelectric properties vs. temperature of K2TlAsZ6 (Z = F, Cl, Br, and I) using the TB-mBJ method. | ||
The generated electric potential in a compound when exposed to a temperature differential is predicted by the Seebeck coefficient (S), a thermoelectric characteristic. Fig. 6(a) for K2TlAsZ6 (Z = F, Cl, Br, and I) DPs displays the computed S versus temperature. All of the temperature ranges under consideration have positive values for S. Consequently, like the majority of DPs, K2TlAsZ6 (Z = F, Cl, Br, and I) DPs are p-type semiconductors.42–44 Moreover, the S-value for DPs falls as the temperature escalates. Table 3 displays, accordingly, the calculated values of S for K2TlAsZ6 (Z = F, Cl, Br, and I). Fig. 6(a) shows that at 300 K, S values for K2TlAsZ6 (Z = F, Cl, Br, and I); at higher temperatures, the S values for the K2TlAsZ6 (Z = F, Cl, Br, and I) compounds decline. There have also been reports of this kind of fluctuation in S for other DPs, when F is replaced at the X-site by Cl, Br or I, respectively.45
| Compounds | K2TlAsF6 | K2TlAsCl6 | K2TlAsBr6 | K2TlAsI6 | ||||
|---|---|---|---|---|---|---|---|---|
| Temperature | 300 K | 1000 K | 300 K | 1000 K | 300 K | 1000 K | 300 K | 1000 K |
| a RefA.19 | ||||||||
| S (μV K−1) | 234.17 | 239.54 | 231.63 | 238.21 | 243.89 | 244.42 | 232.50 | 246.74 |
| 238A | 254A | |||||||
| σ (S m−1) | 38.77 | 213.50 | 24.77 | 131.59 | 16.28 | 98.17 | 15.52 | 87.30 |
| κe (Wm−1 K−1) | 85.93 | 1630.35 | 53.96 | 1000.00 | 38.19 | 780.64 | 34.22 | 756.44 |
| PF | 0.21 | 1.23 | 0.13 | 0.75 | 0.10 | 0.59 | 0.08 | 053 |
| ZT | 0.74 | 0.75 | 0.73 | 0.74 | 0.76 | 0.75 | 0.73 | 0.70 |
| 0.75A | 0.99A | |||||||
A crucial thermoelectric characteristic, σ/τ, measures the amount of charge carriers that may be transported through a material. Fig. 6(b) displays the computed σ/τ that is dependent on temperature for the K2TlAsZ6 (Z = F, Cl, Br, and I) compounds. DPs follow the same trend as the other semiconductors, in that σ/τ grows linearly with temperature. This is because a greater number of charge carriers become accessible for conductivity as temperatures rise. Table 3 presents the σ/τ values calculated for K2TlAsZ6 (Z = F, Cl, Br, and I). The greater band gap is the main reason why K2TlAsF6 has greater σ values compared to other compounds. The relationship between the σ and the carrier concentration (N) may be expressed as σ = Neμ.
The temperature-dependent κe is shown in Fig. 6(c) as having greater values for the K2TlAsF6 compound compared to other compounds. One possible explanation is that K2TlAsF6 has a greater band gap than other compounds. The band gap has also been shown to cause this kind of κe change in other DPs.45,46 On the other hand, similar to other types of semiconductors, the values of κe DPs grow as the temperature rises.13,47 The highest κe values for K2TlAsZ6 (Z = F, Cl, Br, and I) at the temperatures of 300 K and 1000 K are summarized in Table 3.
The power factor (PF) precisely computes the thermoelectric functionality of a material, which is expressed by:
| PF = S2 × σ | (9) |
The relationship among the Seebeck coefficients (S) and electrical conductivity (σ) is described in eqn (9). Fig. 6d shows the PF vs. T (K) data for the K2TlAsZ6 (Z = F, Cl, Br, and I) compounds, depicting the potential use of these compounds in thermoelectric devices with a PF of one or greater than unity.48,49 The PF values increase linearly with the upturn in temperature from 300 K to 1000 K. The decrease in the (S) and surge in the EC (σ) may be due to the variation in the temperature, affecting the PF metrics for the K2TlAsZ6 (Z = F, Cl, Br, and I) DPs. The PF values at different temperatures are given in Table 3.
A material's figure of merit (ZT) is a compilation of its thermoelectric efficiency, which is determined as follows:50
![]() | (10) |
For the ZT values to be high, the denominator (κe) must be low, while S and σ must be high. Fig. 6(e) displays the computed ZT for K2TlAsZ6 (Z = F, Cl, Br, and I). When the temperature is elevated to 1000 K, the ZT standards for K2TlAsZ6 (Z = F, Cl, Br, and I) reach 0.75, 0.74, 0.75, and 0.70, respectively. A prominent aspect of DPs is this rising tendency in ZT with temperature.51 Although some predictable DPs have lower ZT standards, K2TlAsZ6 (Z = F, Cl, Br, and I) DPs have maximum ZT standards.52 Thus, K2TlAsZ6 (Z = F, Cl, Br, and I) are promising materials for thermoelectric power generation.
![]() | ||
| Fig. 7 Thermoelectric properties vs. chemical potential of K2TlAsZ6 (Z = F, Cl, Br, and I) using the TB-mBJ method. | ||
Semiconducting compounds exhibit pronounced sensitivity to temperature gradients; whereby elevated temperatures promote the excitation of charge carriers from the valence band (VB) to the conduction band (CB). This thermal activation results in an increased carrier concentration across the VB and CB gap. The dependence of electrical conductivity (σ) on the chemical potential (μ − EF) is presented in Fig. 7a–d.
As illustrated in Fig. 7a–d, the σ/τ ratio exhibits an increasing trend with temperature at lower chemical potentials. Conversely, at higher chemical potentials, σ/τ decreases as the temperature rises. This behavior is in agreement with the theoretical expectations for intrinsic semiconductors. For the investigated K2TlAsZ6 (Z = F, Cl, Br, and I) double perovskite compounds, the calculated results reveal that the electrical conductivity is more significantly enhanced under negative (μ − EF) conditions than under positive (μ − EF) shifts relative to the Fermi level.56
The dependence of κe/τ on the chemical potential exhibits a trend analogous to that of σ/τ, albeit with a more pronounced temperature dependence. This discrepancy can be attributed to the Wiedemann–Franz law (κe = LσT), which relates the electronic thermal conductivity to the electrical conductivity by means of the Lorentz number (L).57 This trend further reveals that the κe to τ ratio increases with temperature, as clearly depicted in Fig. 7a–d, thereby validating the reliability of the computed results.
To assess the performance of the compounds under study, the power factor (PF) was calculated using the relation (PF = S2σ), where S is the Seebeck coefficient and σ is the electrical conductivity. This parameter reflects the efficiency of a material in converting thermal energy into electrical power. Fig. 7a–d illustrates the variation of the chemical potential with respect to PF for the K2TlAsZ6 (Z = F, Cl, Br, and I) compounds. The results indicate that with increasing temperature from 300 K to 1000 K, the peak positions of PF shift away from the Fermi level and exhibit an increase in magnitude. This trend suggests that elevated temperatures and moderate carrier concentrations contribute to an enhancement in the power factor. Additionally, it is observed that the maximum PF values are located in the p-type region, indicating that these compounds possess significant potential for application as p-type thermoelectric materials. The relationship between ZT and chemical potential is also depicted in Fig. 7a–d. Fig. 7a–d depicts the decrease in the thermoelectric efficiency (Figure of Merit) ZT value for K2TlAsZ6 (Z = F, Cl, Br, and I), respectively. The peak nearly reaches unity. The gradual decrease in the ZT metrics with increasing temperature indicates the peaks approaching unity (1.0) within the lower temperature limits (300 K).55,58
| Compounds | Pressure | Gibbs free energy G* (×107) [kJ mol−1] | Entropy (S) [J mol−1 K−1] | ||||||
|---|---|---|---|---|---|---|---|---|---|
| 300 K | 500 K | 700 K | 1000 K | 300 K | 500 K | 700 K | 1000 K | ||
| K2TlAsF6 | 0 GPa | −6.394 | −6.394 | −6.394 | −6.394 | 346.946 | 482.670 | 578.845 | 689.461 |
| 10 GPa | −6.393 | −6.394 | −6.394 | −6.394 | 259.326 | 385.177 | 473.546 | 571.786 | |
| K2TlAsCl6 | 0 GPa | −6.960 | −6.960 | −6.960 | −6.960 | 410.583 | 548.675 | 646.129 | 758.499 |
| 10 GPa | −6.960 | −6.960 | −6.960 | −6.960 | 287.857 | 414.489 | 502.298 | 599.054 | |
| K2TlAsBr6 | 0 GPa | −1.034 | −1.034 | −1.034 | −1.034 | 468.737 | 607.868 | 705.701 | 818.426 |
| 10 GPa | −1.034 | −1.034 | −1.034 | −1.034 | 333.206 | 461.502 | 549.534 | 646.041 | |
| K2TlAsI6 | 0 GPa | −1.744 | −1.744 | −1.744 | −1.744 | 522.761 | 662.775 | 761.152 | 874.695 |
| 10 GPa | −1.744 | −1.744 | −1.744 | −1.744 | 368.341 | 497.301 | 585.192 | 681.141 | |
The equation for the entropy using the Debye–Slater model is expressed as:
| S = −3nkB ln(1 − e−θD/T) + 4nkBD(θD/T) | (11) |
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05932a.
Footnote |
| † Murtaza Shahab and Muhammad Awais Jehangir made equal contributions. |
| This journal is © The Royal Society of Chemistry 2025 |