Open Access Article
Muhammad Yousufa,
Ali Razaa,
Farooq Alia,
Hamid Ullah
*a,
Young-Han Shin
b and
Essam A. Al-Ammarc
aDepartment of Physics, RIPHAH International University, Lahore Campus, Pakistan. E-mail: hamid.uou@gmail.com
bMultiscale Materials Modeling Laboratory, Department of Physics, University of Ulsan, Ulsan 44610, Republic of Korea
cDepartment of Electrical Engineering, College of Engineering, King Saud University, P.O. Box 800, Riyadh 12372, Saudi Arabia
First published on 1st December 2025
We employed first-principles DFT calculations to investigate the structural, optoelectronic, and thermoelectric properties of halide-based MoSX (X = Cl, Br, I) materials. We observed that these materials exhibit an energetically stable nature due to lower formation energies. MoSX (X = Cl, Br, I) possess a semiconducting behavior with bandgaps ranging from 1.24 eV to 1.38 eV. Notably, MoSCl and MoSBr show a direct bandgap, which is advantageous for optoelectronic devices such as light-emitting diodes and solar cells. Moreover, the calculated figure of merit (ZT) values highlight the suitability of MoSX (X = Cl, Br, I) for thermoelectric applications. These findings establish a theoretical foundation for experimentalists to pursue applications in solar energy generation and thermal energy management.
This limitation has spurred interest in this new class of materials2,7,11,12 with exceptional properties, such as broken mirror symmetry, inducing a vertical dipole moment and piezoelectricity, which enhance charge separation and interfacial reactivity.13 Recent advances in synthesis, including the pioneering work by Lu et al.13 on MoSSe via selective epitaxy, have demonstrated the feasibility of creating such asymmetric architectures.11,14–16 Computational studies further predict that tailoring the Janus composition can modulate electronic and optical properties, making them ideal for energy conversion.2,12,16,17
Halide-based Janus materials, particularly MoSX (X = Cl, Br, I), represent an underexplored frontier. Replacing one chalcogen layer with a halogen atom introduces significant electronegativity contrast, potentially amplifying built-in electric fields and enabling novel bandgap engineering.2,13,18 For instance, Waheed et al.,3 Asghar et al.,16 and manymore7,11,12,17 computationally demonstrated that MoSX exhibits enhanced structural stability and anisotropic carrier mobility compared to symmetric TMDs.13 The tunability of halides (Cl, Br, I) also allows systematic control over electronic properties: lighter halides like Cl widen bandgaps, while heavier counterparts (I) reduce them, aligning with requirements for visible-light photocatalysis.19–22 Moreover, the intrinsic polarization in Janus MoSX promotes efficient electron–hole separation, a critical factor for hydrogen evolution reactions (HER) and solar fuel generation.10,18,23
Despite these prospects, comprehensive computational insights into the physicochemical properties of MoSX systems remain sparse. Previous studies have focused on chalcogen-based Janus materials, leaving halide variants relatively unexplored. This knowledge gap hinders the rational design of MoSX for targeted applications. For example, the role of halogen electronegativity in modulating interfacial charge transfer—key for photocatalytic water splitting—requires deeper scrutiny.3,11,19
In this study, we employ density functional theory (DFT) simulations to unravel the structural stability, electronic structure, optical absorption, and thermoelectric properties of MoSX (X = Cl, Br, I). By analyzing the interplay between halogen choice and material properties, we aim to establish design principles for optimizing these materials in green energy applications. Our findings reveal that MoSX exhibits a bandgap ranging 1.24–1.38 eV, strong absorption, and suitable ZT (∼0.75) make them ideal for solar cells and thermoelectric applications. Ultimately, this computational framework provides a roadmap for synthesizing and integrating MoSX into next-generation energy technologies.
In order to have reliable electronic properties, we used the Tran–Blaha modified Becke–Johnson (TB-mBJ) potential27 to increase the accuracy. TB-mBJ is a well-established semi-local exchange-correlation functional capable of predicting bandgaps closer to experimental values than standard GGA.27,28 The WIEN2k framework divides the crystal lattice into muffin-tin spheres which represent a harmonic expansion of the electron density, and an interstitial region which is described by plane waves. The basis set size was controlled by the parameter RMT × Kmax = 8, ensuring energy convergence. The Fourier expansion of the charge density and potential was truncated at Gmax = 16 Ry−1, and the energy convergence was set to 1 × 10−6 Ry. A k-mesh of 21 × 21 × 10 was adopted for Brillouin zone integration, balancing the computational efficiency and precision. For the optical properties calculations, we followed the same techniques as in ref 2 and 29. For thermoelectric properties, including electrical conductivity and Seebeck coefficient, the semi-classical Boltzmann transport equations were solved under the constant relaxation time approximation using the BoltzTrap package.30 This approach leverages the DFT-calculated band structure and carrier dispersion relations to evaluate transport coefficients as functions of temperature and chemical potential.
3m (# 216), which is confirmed by the Jain et al..31 The atomic positions within the unit cell are defined by the Wyckoff positions at 16e sites for Mo (0.35, 0.15, 0.85), S (0.12, 0.38, 0.62), and X (0.12, 0.38, 0.12), as illustrated in Fig. 1(a). Prior to computing the physical properties, structural optimization was performed to determine the ground state energies at different volumes, as illustrated in Fig. 1(b). From Fig. 1(b), one can observed that MoSBr possesses lower optimization energy in comparison to MoSCl and MoSI. The obtained lattice parameters for MoSX (X = Cl, Br, I) follows the trend and agreement with previously reported theoretical work.2
![]() | ||
| Fig. 1 The optimized (a) structure and (b) optimization curve of the bulk MoSX (X = Cl,Br,I). The Mo, S, and X (Cl, Br, I) are represent with color purple, yellow, and green. | ||
Our computed results indicate a monotonic increase in the lattice parameter (summarized in Table 1) from Cl to I, consistent with the increasing ionic radii of the halogens, following the relation
. Conversely, the bulk modulus (B0) exhibits a decreasing trend along the same series, adhering to the inverse proportionality
.29,32 Furthermore, one can see from Table 1, that bond length between Mo and S (denoted by d1) remain unchanged, while it changes significantly between Mo and X (shown by d2) atoms, which is due to difference in electronegativity.
| Compound | a0 (Å) | d1 Mo-S(Å) | d2 Mo-X(Å) | Ef (eV) | Eg (eV) |
|---|---|---|---|---|---|
| MoSCl | 9.70 | 2.32 | 2.62 | −1.28 | 1.38 |
| MoSBr | 9.96 | 2.32 | 2.76 | −0.97 | 1.31 |
| MoSI | 10.36 | 2.33 | 2.92 | −0.93 | 1.24 |
The structural properties of these materials directly influence their electronic behavior. Variations in the lattice parameter, induced by anion substitution (Cl → Br → I)29,32–34 or cation modification,35,36 lead to tunable physical properties. Specifically, the increase in ionic radius from Cl to I results in lattice expansion and a concomitant reduction in the band gap. This trend arises from the decreased binding energy of valence electrons, which require less energy for excitation into the conduction band.
The thermodynamic stability of the investigated double perovskites is confirmed by the negative formation energies (Ef = [EMoSX − (EMo + ES + EX)], where EMoSX, EMo, ES, and EX representing the energies of the MoSX unit cell, the total energies of the Mo, S, and Cl atoms in their bulk form), as listed in Table 1. Notably, the magnitude of Ef decreases progressively as the halide ion (X) varies from Cl to I.29 This trend can be attributed to the increasing lattice constant, which enhances the effective interionic separation within the crystal lattice, thereby reducing the Ef. We have compared our results with the already available literature on bulk SnXY(X ≠ Y = S, Se)2 and found a good agreement.
For the photovoltaic efficiency, the ideal band gap range is 0.8–2.2 eV.29,39 The computed band gaps of MoSX (X = Cl, Br, I) ranges from 1.24–1.38 eV fall within this range, The materials with band gap in this range exhibit properties comparable to lead-based perovskites while offering the advantage of non-toxicity, making them superior alternatives for photovoltaic applications.40
The electronic bands were further examined by computing the total/projected (T/PDOS). The TDOS and PDOS for MoSX (X = Cl, Br, I) were evaluated within an energy range of −4 eV to 4 eV, as illustrated in Fig. 3. The TDOS plots demonstrate that going from Cl to I induces a more pronounced downward shift in the conduction states compared to the upward displacement of the valence states. This asymmetric energy state displacement results in band gap narrowing and tunability.
![]() | ||
| Fig. 3 The calculated DOS and contributions of orbitals of different atoms of MoSBr, MoSCl, and MoSI. | ||
The valence band maximum (VBM) is primarily formed by the hybridization of Mo-d and S-/X-p states, whereas the conduction band minimum (CBM) arises from the interaction between S-p and X-p states. This hybridization generates an effective attractive interaction, leading to a downward shift in the conduction band states (dominated by I Mo-d and S-/X-p). Consequently, the X (Cl, Br, I) atoms ionic radius critically governs the hybridization-driven band gap suppression, offering a practical pathway for tailoring optoelectronic materials with tunable spectral responses.
The stability of the MoSX (X = Cl, Br, I) is confirmed by the Debye temperature (θD). From tensor matrix of Charpin,41 the value of θD could be computed in terms of average sound velocity, as given in the relation below ;42–44
From the transverse parts of the Navior equations vt and the longitudinal vl, the average sound velocity can be found out by the following relation:
The computed value of has been expressed in Table 2. The calculated s θD and sound velocities provide valuable information on the lattice dynamics and thermal transport properties of MoSX (X = Cl, Br, I). The θD is directly related to the highest vibrational frequency of the lattice and reflects the stiffness of the crystal. A higher θD indicates stronger interatomic bonding and lower lattice anharmonicity, which are favorable for enhanced thermal conductivity. Among the studied compounds, MoSI exhibits the highest θD (642.4 K), indicating stronger bonding and a stiffer lattice, whereas MoSCl shows the lowest θD (453.9 K), suggesting relatively softer lattice vibrations compared to MoSBr and MoSI. This trend correlates with the increase in atomic mass from Cl to I, leading to enhanced phonon frequencies and thus higher θD in the iodide compound.
The calculated Debye temperatures and acoustic sound velocities are summarized in Table 2. These parameters provide important insight into the lattice dynamics and thermal transport behavior of the MoSX (X = Cl, Br, I) double perovskites. The Debye temperature (θD) reflects the lattice stiffness and strength of interatomic bonding.
Similarly, the longitudinal (vl) and transverse (vt) sound velocities, along with the average sound velocity (vm), increase in the order MoSCl < MoSBr < MoSI, which again confirms the higher lattice rigidity of MoSI. The higher sound velocities in MoSI imply faster phonon transport and potentially higher lattice thermal conductivity, while MoSCl is expected to show comparatively lower phonon transport efficiency due to its lower sound velocity values.
These observations are also consistent with the bandgap trend (MoSCl = 1.38 eV > MoSBr = 1.31 eV > MoSI = 1.24 eV). The compound with the largest bandgap (MoSCl) shows weaker lattice stiffness, while the one with the smallest bandgap (MoSI) exhibits stronger bonding and more compact lattice dynamics. Overall, the results highlight how the halide substitution significantly influences lattice vibration behavior and thermal transport properties, which is crucial in evaluating their suitability for thermal management and optoelectronic applications.
Beyond the zero-frequency limit, ε1(ω) increases, reaching maximum values at energies of 2.12 eV, 2.15 eV, and 1.96 eV for MoSX (X = Cl, Br, I), respectively. After attaining these maxima, ε1(ω) declines with increasing incident energy. The real dielectric function exhibits distinct peaks as a function of incident energy, arising from transitions between discrete energy levels in the valence and conduction bands.7,11 Notably, ε1(ω) becomes negative for all three double perovskites within the energy range 6.0–7.5 eV, indicating a metallic character in this regime. The imaginary part of the dielectric function, ε2(ω), quantifies photon absorption within the 0–8.0 eV range. This absorption results from electronic excitations from occupied valence-band states to unoccupied conduction-band states, as illustrated in Fig. 4(b). The ε2(ω) spectra for all three perovskites exhibit a linear increase beyond distinct absorption edges, which correspond to their respective electronic band gaps (see Fig. 4). The first prominent absorption peaks for MoSX (X = Cl, Br, I) occur at 3.51 eV, 2.98 eV, and 2.39 eV, respectively. A systematic red shift in these peaks is observed as the halogen anion varies from Cl to I, consistent with the trend in the calculated band gaps.
Additionally, the complex refractive index, n(ω), was computed for MoSX (X = Cl, Br, I), as shown in Fig. 4(c). Comparative analysis both the ε1(ω) and n(ω) exhibit analogous energy-dependent behavior. We determined the n(0) from the low-energy extrapolation, yields values of 3.01, 3.02, and 3.16 for MoSX (X = Cl, Br, I), respectively. These values satisfy the fundamental relation ε1(0) = n2(0),46 validating the consistency between dielectric and optical responses. Beyond the static limit, n(ω) reaches maximum values of 3.83, 3.85, and 3.88 at energies of 2.12 eV, 2.15 eV, and 1.96 eV for the respective compounds, followed by a monotonic decrease with increasing energy. Notably, n(ω) falls below unity at in the higher energies for the Cl, Br, and I, respectively. In higher energy regime, the materials exhibit superluminal propagation characteristics,47 permitting photon transmission without significant phase retardation. Fig. 4(d) presents the extinction coefficient k(ω), which mirrors the behavior of the imaginary dielectric function ε2(ω), confirming its role in characterizing optical absorption. Prominent absorption peaks occur in the energies range 2–7 eV for the respective compounds, in excellent agreement with the ε2(ω) spectra.
The absorption coefficient α(ω) quantifies electronic transitions from the valence to conduction bands induced by incident photons. The overall absorption intensity increases with photon energy, reaching prominent peaks at in the given energies range (shown in Fig. 4(e). The tunable absorption edges (spanning ultraviolet to visible energies) coupled with strong UV absorption suggest promising applications of these double perovskites in optoelectronic devices.
The optical reflectivity R(ω), which quantifies surface interactions, is presented in Fig. 4(f). The reflectivity peaks at higher energies correlate with the onset of metallic behavior, as evidenced by negative values of the real dielectric function ε1(ω). The tunable optical properties, particularly in the visible-to-UV range, suggest promising applications in optoelectronic devices and energy-harvesting technologies.
The σ/τ arises from the movement of free charge carriers (electrons or holes) and exhibits a temperature-dependent increase due to enhanced carrier kinetic energy (Fig. 5(a)). The σ/τ increases with the increasing temperature. One can observed from the Fig. 5(a), that the σ/τ decreases when we go from Cl to I. This decrease may be due to the increasing molecular size, results in weaker overlap between their p-orbitals, leading to less efficient electron delocalization.
The Seebeck coefficient (S), a key parameter reflecting the thermoelectric voltage induced by a thermal gradient, demonstrates high sensitivity to temperature (Fig. 5 (b)). At 100 K, the S values for at 200 K for MoSCl (180 µV K−1), MoSBr (148 µV K−1), and MoSI (224 µV K−1) decreases to 130 µV K−1, 125 µV K−1, and 150 µV K−1 at 800 K. The sustained high Seebeck coefficients suggest strong thermoelectric response even at elevated temperatures. The Seebeck coefficient (S) varies with the anion substitution (Cl → Br → I), correlating with the reduction in band gap (Eg). As the band gap narrows from Cl to I, the charge carrier density increases, requiring less energy for carrier excitation, thereby enhancing electrical conductivity.
Thermal conductivity in these materials is dominated by electronic contributions (Ke/τ), as phonon effects are negligible. As depicted in Fig. 5(c), Ke/τ increase from 0.23–031 W K−1 m−1 s−1 (200 K) to 0.54–1.05 W K−1 m−1 s−1(800 K) for MoSX (X= Cl, Br, I). The low thermal conductivity is desirable for efficient thermoelectric performance.
Additionally, we have calculated the heat capacity (Cν), representing the thermal energy storage capacity, rises with temperature (Fig. 5(d)). The MoSCl, and the MoSBr take the lead to reach the Dulong-petit asymptotes at T ∼ 400 K and the MoSI reaches to asymptotes value with relatively higher temperature (T ∼ 600 K). Moreover, the Cν increase from 4.4–7.8 J mol−1 K−1 (200 K) to 9.95–16.5 J mol−1 K−1 (800 K) for MoSX (X= Cl, Br, I). This trend indicates superior thermal stability for Cs2InBiI6, suggesting its potential for high-temperature thermoelectric applications.
The power factor (PF = S2σ/τ), a critical metric for device applications, is derived from the Seebeck coefficient and electrical conductivity. Fig. 5(e) and 6(a) illustrates a significant enhancement in PF across the temperature range: MoSCl (0.33–1.1 × 1011 W K−2 m−1 s−1), MoSBr (0.19–0.58 × 1011 W K−2 m−1 s−1), and MoSI (0.186–1.37 × 1011 W K−2 m−1 s−1). This trend underscores the potential of the MoSX (X = Cl, Br, I) for high-temperature thermoelectric applications, provided phase stability is maintained. Notably, the PF exhibits higher magnitudes on the p-type region due to the enhanced S in this regime. The figure of merit (ZT) is a decisive metric for evaluating the efficiency of thermoelectric materials in thermal energy conversion devices.49,50
![]() | ||
| Fig. 6 The calculated (a) power-factor, and (b) figure of merit with restive chemical potential for MoSX (X = Cl, Br, I). | ||
The dependence of ZT on chemical potential (µ) illustrated in Fig. 5(f) and 6(b). At optimal carrier concentration the ZT values approach unity for n-/p-type regimes. At 300 K, the ZT values for MoSCl, MoSBr, and MoSI are computed to be 0.69, 0.62, and 0.76, respectively, demonstrating their potential for thermoelectric applications. Our calculated values of ZT at 300 K are higher than the available experimental study on ReSTe (0.4).51 Thus, the investigated MoSX (X = Cl, Br, I) compounds exhibit promising thermoelectric performance, making them viable candidates for energy conversion applications.
| This journal is © The Royal Society of Chemistry 2025 |