Jipin
Peter
a,
Tanu
Choudhary
a and
Raju K.
Biswas
*ab
aDepartment of Physics, Faculty of Natural Sciences, M S Ramaiah University of Applied Sciences, Bengaluru 560058, India. E-mail: rkb@nerist.ac.in
bDepartment of Physics, North Eastern Regional Institute of Science and Technology, Nirjuli, Arunachal Pradesh-791109, India
First published on 4th July 2025
Heterostructures have recently emerged as suitable candidates for energy applications owing to their intriguing high electronic properties and low thermal conductivity. Traditionally, experimentally observable chiral phonons, protected by the crystal hexagonal symmetry in two-dimensional honeycomb lattices and originating primarily at the Brillouin zone corner, play a critical role in achieving ultra-low lattice thermal conductivity (κl). It is mainly attributed to the three-fold rotational symmetry-driven circularly polarized chiral phonons that restrict certain scattering processes and reduce thermal resistance. Herein, we identify an innovative mechanism that not only preserves phonon chirality but also minimises lattice thermal conductivity by satisfying certain phonon selection rules and enhancing anharmonic scattering, particularly in the long wavelength limit. This study encompasses first-principles calculations and solutions to the Boltzmann transport equation, along with various thermal transport theories to investigate the structural, thermal and electronic transport properties of group-IV-based dichalcogenides in hexagonal phase MSe2 (M = Mo and W) and their heterostructures (HS), i.e. MoSe2/WSe2 and MoSeTe/WSeTe. Interestingly, phenomena such as membrane effect and hybridization of acoustics and low-lying optical phonons combined with phonon branching and phonon bunching are identified to execute an imperative role in suppressing thermal transport, resulting in a low κl value of 2.83 W m−1 K−1 in MoSe2/WSe2 and an ultra-low κl of 0.5 W m−1 K−1 in MoSeTe/WSeTe HS. The disparity in computing carrier mobility using deformation potential theory (DPT) is addressed herein, and to rectify the inconsistency, we additionally incorporate the Fröhlich interaction, which accounts for longitudinal optical phonons in estimating the carrier mobility accurately. Overall, the present work exploits the intrinsic properties of chiral phonons through breaking the crystal inversion symmetry, providing important insights into the rational design of low-dimensional thermoelectric materials.
Recently, the emergence of chiral phonons in the WSe2 monolayer has been identified experimentally through the intervalley scattering of holes during the indirect infrared absorption using transient infrared spectroscopy.1 Moreover, the breaking of spatial inversion symmetry in MoSe2/WSe2 vdW heterostructures induces long-lived indirect interlayer excitons with opposite circular polarizations, as observed under the chiral optical pumping.11,12 In recent times, Zhang et al. have carried out first-principles calculations to show the potential application of chiral phonons in the optical transition of MoS2/WS2 heterostructures.3 Moreover, Pandey et al. shows that chiral phonons in the 1D chain of Ba3N follow new symmetry-based scattering rules that limit three-phonon interactions in the system.13 However, weak chain coupling breaks these scattering rules in the bulk structure, resulting in a lower κl value.13 Therefore, tailoring the mechanism of chiral phonons specifically by weakening phonon chirality can enhance the three-phonon scattering events. The inherent phonon–phonon interactions arising from such anharmonic systems can undergo the lowest-order anharmonic processes, namely AAA (acoustic–acoustic–acoustic), AAO (acoustic–acoustic–optical), AOO (acoustic–optical–optical) and OOO (optical–optical–optical), which includes the decay of a phonon into two others or the coalescence of two phonons into a third as reported by Ravichandran et al.14 Recently, the MoS2/MoSe2 bilayer heterostructure has shown an intrinsically low lattice thermal conductivity of 25.39 W m−1 K−1 compared to the values of 84 W m−1 K−1 and 59 W m−1 K−1 of their monolayer counterparts MoS2 and MoSe2, respectively, highlighting the effective significance of heterostructures in lowering the lattice thermal conductivity and their potential application in thermoelectric devices.15
Considering the experimental synthesis and technical effectiveness, we used the first-principles-based density functional theory (DFT) calculations to effectively model the MoSe2/WSe2 HS by choosing single-layer MoSe2 and WSe2 as parent materials. Additionally, we investigated the variation in thermoelectric properties occurring due to the substitution of Te atoms in place of Se, at one side of each layer in the HS. Furthermore, we explored the underlying mechanisms of phonon modes, which enable overcoming the limitation imposed by phonon chirality in minimising κl and maximising ZT in both HS compared to their monolayer counterparts. Our results reveal that the MoSe2/WSe2 HS exhibits phonon branching, leading to the AOO scattering process, whereas the bunching of acoustic and LLO phonon modes observed in the MoSeTe/WSeTe HS results in the AAA process. Interestingly, the MoSe2/WSe2 HS shows chirality due to the breaking of spatial inversion symmetry compared to their homo-structures; however, doping of Te induces distinctive phonon features such as phonon bunching and mass-induced phonon softening, enhancing the three-phonon scattering process resulting in an ultra-low κl value of 0.5 W m−1 K−1 at room temperature and a maximum ZT value of 3.37 at 700 K.
In this study, for electronic transport calculations such as density of states (DOS) analysis for both heterostructures, we performed a non-self-consistent field (NSCF) calculation using an increased k-grid of 24 × 24 × 4. Moreover, we incorporated the LOBSTER code23 as a post-processing tool along with the QE package for evaluating the crystal orbital Hamilton population (COHP)24 analysis and computed the integrated crystal orbital bond index (ICOBI).25 The COHP analysis revealed the interaction between two atomic orbitals of neighbouring atoms of a material in terms of their Hamiltonian matrix element.24 The multiplication of this matrix element with the corresponding density of states gives a quantitative measure of the bonding strength arising from bonding, non-bonding, or antibonding interactions.24 Additionally, the ICOBI values serve as an effective tool for understanding the chemical bonding hierarchy by detecting covalent and ionic bonds in solid-state materials.25 Furthermore, we evaluated the electrical transport coefficient (i.e., Seebeck coefficient), by solving the semi-classical Boltzmann transport theory within the constant scattering time approximation (CSTA), as implemented in the BoltzTraP code.26
As shown in Fig. 1a, the MoSe2 monolayer is vertically stacked on top of the WSe2 monolayer via interlayer van der Waals interactions, constituting the MoSe2/WSe2 HS. Composing single-layer MoSe2 on WSe2 along the normal direction breaks the mirror symmetry at the ±K point, which gives rise to non-degenerate phonon modes with definite polarisation. Based on the relative position of the two monolayers, different stacking configurations are considered. Previous studies have demonstrated that the low-energy configuration of such heterostructures is achieved by stacking the metal atom of one layer directly above the chalcogen atoms of the other layer, namely the AA′ stacking.34,35 In our present study, the Mo and Se atoms of the MoSe2 layer were exactly aligned over the Se and W atoms of the WSe2 layer, respectively, to form the MoSe2/WSe2 heterostructure with the AA′ configuration, as shown in Fig. 1a.35 This heterostructure also lacked the inversion symmetry similar to its parent materials (MoSe2 and WSe2 monolayers), arising due to the AA′ configuration.35 After geometry optimisation, the MoSe2/WSe2 HS maintains the hexagonal symmetry similar to its parent materials, with lattice parameters of a = b = 3.32 Å and an interlayer distance (d) of 6.49 Å (the vertical distance from Mo to W atom), which closely aligns with the previously reported theoretical value (d = 6.48 Å).36 The bond lengths between Mo–Se (∼2.54 Å) and W–Se (∼2.54 Å) atoms in the heterostructure are identical and are closely similar to those in the individual MoSe2 (∼2.54 Å) and WSe2 (∼2.55 Å) monolayers,27 respectively, as presented in Table S1 of the ESI.† The side and top views of the MoSe2/WSe2 HS are illustrated in Fig. 1a and b, respectively.
Furthermore, we modeled the 2H phase of the MoSeTe/WSeTe HS by stacking one layer of Janus MoSeTe on top of the WSeTe monolayer, which preserved the hexagonal symmetry having AA′ stacking analogous to the MoSe2/WSe2 heterostructure. In our present study, we designed different configurational spaces with random sampling of Se and Te atoms to explore different atomic arrangements and calculate the DFT energies for those configurations. Our analysis reveals that the structure shown in Fig. 1c represents the most dynamically stable configuration under the harmonic approximation, which is opted for further investigation. The optimized MoSeTe/WSeTe heterostructure has a lattice parameter of a = b = 3.46 Å and an interlayer distance of 6.33 Å. This decrement in the interlayer distance of Te doped system is attributed to the enhanced van der Waals attraction between the sub-layers in MoSeTe/WSeTe (0.39 eV) compared to the MoSe2/WSe2 (0.35 eV) heterostructure. Additionally, the bond length of the Mo–Se atom is the same as the bond length of W–Se (∼2.56 Å), while the bond lengths of Mo–Te and W–Te are also found to be identical (∼2.73 Å), and the corresponding values are presented in Table S1 of the ESI.† The increase in lattice parameter and bond lengths in MoSeTe/WSeTe compared to the MoSe2/WSe2 HS is attributed to the larger atomic size of Te than that of the Se atom. The schematic representations of the side and tilted top views of the MoSeTe/WSeTe HS are depicted in Fig. 1c and d, respectively. Moreover, to assess the stability and bonding nature of these heterostructures, we evaluated the binding energy, Born criteria of elastic stability37 (Table S2, ESI†), and performed the crystal orbital Hamilton population (COHP) (Fig. S1, ESI†) and Bader charge analyses (Table S3, ESI†), as discussed in ESI† (Note S1).
Since phonon modes play a crucial role in determining the lattice thermal conductivity of semiconductors, we additionally conducted a detailed analysis of these phonon dispersion curves, shown in Fig. 2c and d, in comparison to their monolayer counterparts. For MoSe2 and WSe2 monolayers, there are three atoms in the unit cell, yielding a total of 9 vibrational modes in the phonon spectra. In the case of MoSe2 monolayer, the phonon modes are arranged in three sets, where the first group are the three highly dispersive acoustic branches within a frequency range of ∼0–152 cm−1, the second group are the three mid-range optical phonon branches confined within the frequency width of ∼160–233 cm−1 and third are the three higher energy phonon bands within ∼275–344 cm−1, as evident in Fig. 2a. However, in the WSe2 monolayer, the presence of heavier W atoms shifts the three aforementioned higher energy phonon modes toward the mid-range optical region (∼238–298 cm−1), thus resulting in the softening of these phonon modes. Interestingly, the phonon spectra of the WSe2 monolayer exhibit bunching phonon modes (non-dispersive or branches coming close to one another), specifically observed for the acoustic branches, as shown in Fig. 2b, attributed to the presence of heavier atoms. The bunching of acoustic phonon branches eliminates the dispersive nature of acoustic modes and constrains them to a narrow frequency width of ∼0–120 cm−1, which is 25% lower than that of the MoSe2 monolayer.
Furthermore, our study reveals that the modelling of HS from their monolayers introduces new distinctive features in the phonon spectra. A notable characteristic in the phonon dispersion of heterostructures compared to its monolayers is the appearance of three low-lying optical (LLO)41 phonons (as evident in Fig. 2c and d), including the LO mode, with frequencies as low as 36.98 cm−1 for MoSe2/WSe2 and 35.60 cm−1 for MoSeTe/WSeTe heterostructures. Recently, Xiaokun Gu et al. have also reported the occurrence of LLO phonon branches in the MoS2 bilayer, with the flexural ZO mode reaching a minimum of ∼60 cm−1, which have direct evidence in reducing the lattice thermal conductivity by 12% compared to its single layer.42 The emergence of such LLO phonons in heterostructures can be attributed to the interlayer van der Waals interactions and the proximity effect.33,43 The existence and impact of proximity effect in heterostructures demonstrate that the presence of one layer induces effects in another layer, which is absent in the case of monolayer. It has been previously investigated and shown to have significant influence on properties such as superconductivity and magnetism.43 In our present study, the emergence of LLO phonons, including the LO mode whose frequency is found to be ∼76% lower than that of its parent materials at the high-symmetry Γ point (160 cm−1 and 167 cm−1 for MoSe2 and WSe2 monolayers, respectively, as shown in Fig. 2a and b), can be attributed to the proximity of the MoSe2 sub-layer to the WSe2 sub-layer. Furthermore, we plotted the eigen vectors of low-lying flexural ZO mode for both heterostructures, as shown in Fig. S2a and b (ESI†), where it is clearly visible that the aforementioned proximity effect limits the vibration of the WSe2 sub-layer (similar to the WSeTe sub-layer in MoSeTe/WSeTe), thereby reducing the phonon frequency in both heterostructures compared to their parent materials.
Additionally, these LLO phonons in the MoSe2/WSe2 HS exhibit coupling with the neighbouring acoustic branches along the high symmetry points, as shown in the phonon band structure (Fig. 2c). Particularly, the LLO phonon branches, especially the LO mode in MoSe2/WSe2, exhibit significant dispersiveness (phonon branching) reaching a maximum frequency of 163 cm−1. The presence of such dispersive phonon modes, specifically low-energy phonons, resembles monolayer MoSe2 phonon band dispersion, attributed to the reduced frequency gap between the LLO and higher energy optical phonons, leading to an enhanced coupling between the optical branches compared to the single-layer MoSe2 and WSe2 structures. It is well known that the pronounced lattice thermal conductivity observed in MoSe2 and WSe2 monolayers is primarily due to the presence of a large frequency gap in the phonon dispersion curve.38 However, the presence of dispersive LLO optical modes in the MoSe2/WSe2 HS reduces the frequency gap, contributing to an increased phonon scattering rate, which can simultaneously result in a low lattice thermal conductivity relative to its parent materials. Moreover, the occurrence of LLO modes in the low frequency range induces the acoustic branches of MoSe2/WSe2 HS to exhibit strong phonon softening (Fig. 2c) compared to its parent materials, as evident from the phonon dispersion curve shown in Fig. 2a and b, which further suppresses the heat carrying phonons in the system.
However, in the case of MoSeTe/WSeTe HS, the substitution of the heavier chalcogen atom Te in place of lighter Se reduces the dispersiveness of the phonon branches, effectively eliminating the phonon branching phenomenon, as evident in Fig. 2d. Moreover, the acoustic and LLO branches of the MoSeTe/WSeTe HS exhibit the phonon bunching phenomena similar to those observed in the phonon spectra of the WSe2 monolayer (Fig. 2b). Additionally, the incorporation of Te atoms lowers the frequency (phonon softening) of all phonon modes, including the acoustic and LLO branches introducing distinct energy separation among the phonon groups. One such separation occurs between the LLO (∼30–110 cm−1) and mid-range optical phonons (136–199 cm−1), and another between the mid-range and high-energy optical phonons (216–311 cm−1), attributed to the substitution of heavier Te atoms. The incorporation of heavier Te atoms follows the general diatomic linear chain model, where the zone edge frequency of the acoustic branch is ωac ∝ M−1/2 (M is the mass of heavier element), which causes the relative shift of acoustic spectral frequencies to a lower value in MoSeTe/WSeTe than that in the MoSe2/WSe2 HS, which is consistent with the results obtained by Thomas et al.41 Moreover, the frequency gap between the LLO and mid-range optical modes of the MoSeTe/WSeTe HS primarily arises from the minimal dispersiveness of the LLO branches compared to the MoSe2/WSe2 HS. Notably, the bunching nature of acoustic and LLO phonons in the MoSeTe/WSeTe HS results in strong coupling between the acoustic and LLO phonon branches, confining them within a narrower frequency width of ∼108 cm−1, which is lower than that of the MoSe2/WSe2 HS (∼161 cm−1). This enhanced coupling between the acoustic phonons themselves and with LLO phonon branches may enhance the phonon scattering mechanisms, which limits the lattice thermal conductivity in the system. Furthermore, in the MoSeTe/WSeTe HS, the coupling of phonon branches varies with the q direction and is found to be stiffer at the high symmetry K point (as evident in Fig. 2d), attributed to the occurrence of the phonon softening phenomena. Henceforth, the strong coupling between acoustic and LLO phonons, combined with pronounced acoustic and optical phonon softening, will significantly limit the group velocity of heat carrying acoustic phonons and enhance the anharmonic three-phonon scattering process in the MoSeTe/WSeTe HS, resulting in a low lattice thermal conductivity in the structure.
To provide a more comprehensive understanding of the atomic contribution, we also plotted the phonon density of states (phonon DOS) as a function of frequency, as illustrated in Fig. S3 of the ESI.† In the case of MoSe2 monolayer, low-frequency phonon modes are mainly contributed by the vibration of both Mo and Se atoms, whereas higher frequencies are dominated by the vibration arising from the Mo atom. However, in the case of WSe2 monolayer, the low-frequency range is dominated by the vibration of heavier W atom and higher frequencies are dominated by the vibration from the Se atoms. For the MoSe2/WSe2 HS, it is clearly visible that the major proportion of atomic vibrations in the low-frequency region encompassing both the acoustic and LLO branches are primarily arising from the heavier W atom. However, the higher energy optical branches are due to the vibration of lighter Mo and Se atoms. Furthermore, in the case MoSeTe/WSeTe HS, the major contribution to the acoustic region is from the heavier Te atoms along with vibrations arising from the W atom. The high-energy optical modes are due to the rattling atoms Mo and Se, which are found to have a large atomic vibration as observed in the high energy range in the phonon spectra.
![]() | ||
Fig. 3 Lattice vibrations of phonon modes at the high-symmetry K point for the MoSe2/WSe2 HS and the corresponding frequencies of each mode from 1 to 18. Circular polarisation or phonon chirality is represented by ellipses49 with red arrows. Forces along the out-of-plane direction (z axis) are represented by purple arrows. The size of each ellipse and the length of each purple arrow are proportional to the amplitudes of the circular and linear vibrations, respectively. The direction of each red arrow denotes the corresponding direction of chirality and the phonon modes are arranged in the increasing order of energy. |
Furthermore, the conservation of such quantized phonon PAM contributes an additional selection rule apart from the well-known energy and momentum conservation rules. Hence, in our present study, the introduction of three-fold rotation operator to the phonon wave function ψph yields the expression
where lph gives the phonon PAM, which has a discrete value of 0, +1, or −1 at the ±K points with a three-fold rotational symmetry.3 Herein, all vibrating sub-layers or sub-lattices contribute to ψph and their vibrations give the same phase change under three-fold rotation, i.e., the same lph. Henceforth, we can compute lph by any vibrating sub-lattice. In contrast, the sub-lattices that remain static do not have any contribution to ψph and thus cannot be used to obtain lph. Interestingly, the term ψph has both a nonlocal contribution from the Bloch phase factor eiq·r and a local contribution from sub-lattice relative vibrations.3 Henceforth, usually the phonon PAM is obtained as a sum of corresponding nonlocal orbital PAM (lo) and local spin PAM (ls) of the vibrating sub-lattice.3 Moreover, lo depends on atomic in-plane sites and ls is determined by sub-lattice vibration. The right-handed circular vibration, left-handed circular vibration, and linear vibration along the z-axis are all eigen states of
with ls = 1, −1, and 0, respectively.3 In order to show the conservation of phonon PAM, we also computed the aforementioned right-handed circular vibration, left-handed circular vibration, and linear vibration along the z-axis for both the heterostructures, as evident in Fig. 3 and 4. Here, the circular vibrations are shown in ellipse and the direction of rotation (right and left) is indicated by the red arrow. A recent study has suggested that the occurrence of such phonon chirality limits the three-phonon interaction occurring in the system, attributed to a higher lattice thermal conductivity (κl).13
![]() | ||
Fig. 4 Lattice vibrations of phonon modes at the high-symmetry K point for the MoSeTe/WSeTe heterostructure and the corresponding frequencies for each mode from 1 to 18. |
In the MoSe2/WSe2 and MoSeTe/WSeTe HS, the time reversal symmetry at the Γ point is preserved, resulting in the vanishing of phonon polarisation (nearly equal to zero).3 Hence, we extend our analysis to the K point, where the phonon has definite circular polarization arising from the non-degenerate chiral phonon modes due to the breaking of time reversal symmetry.3 To provide a comprehensive understanding of how phonon chirality affects the κl value of MoSe2/WSe2 and MoSeTe/WSeTe heterostructures, we calculated the phonon eigen vectors and analysed the specific vibrational motion of both layers, as illustrated in Fig. 3 and 4. The phonon chirality on individual atoms is represented by an ellipse, with the amplitude of vibrations indicated by the radius of the ellipse. The different chirality of phonons representing the right- and left-hand phonon rotations is depicted by arrows within the circles. The phonon eigen vectors for each mode at the high-symmetry K point in the MoSe2/WSe2 heterostructure closely resemble those of the previously reported MoS2/WS2 bilayer heterostructure.3 Moreover, the strengths of phonon chirality determined from the phonon PAM (lph) by incorporating the terms lo and ls for both heterostructures are presented in Tables 1 and 2. In the case of MoSe2/WSe2 HS, a majority of the phonon modes possess a definite circular polarisation indicated by lph = ±1, as evident from Table 1. Interestingly, these chiral phonons observed in the MoSe2/WSe2 heterostructure differ from those in the MoSeTe/WSeTe heterostructure, probably due to the substitution of Te atoms. As shown in Fig. 3, it is evident that in the case of MoSe2/WSe2 HS, either the MoSe2 or the WSe2 layer remains stiff with no circular polarisation, while the net chirality arises from other layer, or vice versa, attributed to the lack of spatial inversion symmetry compared to their homo-structures. In contrast, the incorporation of Te atoms in the MoSeTe/WSeTe HS induces opposite circular polarisation in both layers, which will cancel each other and result in a zero net phonon chirality. Furthermore, comparing the phonon PAM values (shown in Tables 1 and 2), it is clear that in the case of MoSe2/WSe2 and MoSeTe/WSeTe HS, the phonon chirality remains persistent. Recent investigation by Tribhuwan Pandey et al. has shown that the symmetry-driven chirality in single-chain Ba3N restricts the phonon–phonon interactions, resulting in high lattice thermal conductivity.13 However, the weakening of this chirality by breaking the symmetry in its bulk counterpart enhances the phonon scattering, leading to a threefold reduction in κl compared to the single-chain structure.13 In contrast, for MoSe2/WSe2 and MoSeTe/WSeTe HS, despite the presence of phonon chirality, the effective tuning of phonon features such as emergence of LLO phonon modes, phonon branching and phonon bunching (as discussed earlier) induced by heterostructure formation may significantly enhance the phonon–phonon scattering, probably resulting in low κl.
Modes | Frequency (cm−1) | l oMo,Se1,Se2 | l oW,Se3,Se4 | l sMo | l sSe1 | l sSe2 | l sW | l sSe3 | l sSe4 | l ph |
---|---|---|---|---|---|---|---|---|---|---|
1 | 83 | −1 | 1 | — | — | — | 1 | 0 | 0 | −1 |
2 | 113.9 | −1 | 1 | — | — | — | 0 | −1 | −1 | 1 |
3 | 127.8 | −1 | 1 | — | — | — | −1 | 1 | 1 | 0 |
4 | 129.7 | −1 | 1 | 0 | 1 | 1 | — | — | — | −1 |
5 | 132.4 | −1 | 1 | −1 | 0 | 0 | — | — | — | 1 |
6 | 157.5 | −1 | 1 | 1 | −1 | −1 | — | — | — | 0 |
7 | 189.9 | −1 | 1 | — | 0 | 0 | — | — | — | 0 |
8 | 195.3 | −1 | 1 | — | −1 | −1 | — | — | — | 0 |
9 | 195.4 | −1 | 1 | — | 1 | 1 | — | 0 | 0 | −1 |
10 | 200.3 | −1 | 1 | — | 1 | 1 | — | 0 | 0 | −1 |
11 | 204.7 | −1 | 1 | — | — | — | — | 0 | 0 | 0 |
12 | 205.9 | −1 | 1 | — | — | — | — | 1 | 1 | 0 |
13 | 240.5 | −1 | 1 | — | — | — | −1 | 1 | 1 | 0 |
14 | 240.9 | −1 | 1 | — | — | — | 0 | −1 | −1 | 1 |
15 | 250.2 | −1 | 1 | — | — | — | −1 | 0 | 0 | 1 |
16 | 278.1 | −1 | 1 | −1 | 0 | 0 | — | — | — | 1 |
17 | 280.9 | −1 | 1 | 0 | 1 | 1 | — | — | — | −1 |
18 | 300.7 | −1 | 1 | 1 | −1 | −1 | — | — | — | 0 |
Modes | Frequency (cm−1) | l oMo,Se,Te | l oW,Se,Te | l sMo | l sSe | l sTe | l sW | l sSe | l sTe | l ph |
---|---|---|---|---|---|---|---|---|---|---|
1 | 47.4 | −1 | 1 | 0 | — | 1 | — | — | −1 | 0 |
2 | 48.8 | −1 | 1 | — | — | 1 | 0 | — | −1 | 0 |
3 | 52.9 | −1 | 1 | — | — | — | −1 | 0 | 0 | 1 |
4 | 68.2 | −1 | 1 | 1 | 0 | 0 | — | — | — | −1 |
5 | 69 | −1 | 1 | −1 | — | — | 0 | 0 | 1 | 0 |
6 | 72.3 | −1 | 1 | 0 | 0 | −1 | — | — | — | 1 |
7 | 146.8 | −1 | 1 | — | — | — | 1 | 0 | 0 | −1 |
8 | 153.1 | −1 | 1 | −1 | 0 | — | — | — | — | 0 |
9 | 159.9 | −1 | 1 | — | 1 | — | −1 | 1 | 1 | 0 |
10 | 165.9 | −1 | 1 | — | — | — | — | −1 | −1 | 0 |
11 | 166.5 | −1 | 1 | — | 1 | 1 | — | — | — | 0 |
12 | 174.1 | −1 | 1 | 0 | −1 | — | — | — | — | 0 |
13 | 217.1 | −1 | 1 | — | — | — | 0 | 1 | — | 0 |
14 | 220.9 | −1 | 1 | — | — | — | −1 | 1 | — | 0 |
15 | 226.4 | −1 | 1 | — | — | — | 1 | 0 | — | 0 |
16 | 245.7 | −1 | 1 | −1 | 0 | — | — | — | — | 0 |
17 | 254.4 | −1 | 1 | 0 | — | — | — | — | — | 0 |
18 | 254.4 | −1 | 1 | 1 | −1 | — | — | — | — | 0 |
![]() | (1) |
![]() | ||
Fig. 5 (a) and (b) Phonon scattering rate, (c) and (d) mode-dependent Grüneisen parameters and (e) and (f) group velocity for MoSe2/WSe2 and MoSeTe/WSeTe heterostructures, respectively, at 300 K. |
The phonon frequency range shown in Fig. 5 is consistent with the dispersion curves as in Fig. 2. It is noteworthy that the scattering rate of the MoSe2/WSe2 HS shown in Fig. 5a is much higher than that of the two parent materials (0.1 ps−1 and 1 ps−1 for MoSe238,45 and WSe245 monolayer, respectively), within the whole frequency range. Previous studies have shown that the presence of high anharmonic scattering rates is one of the fundamental reasons for low lattice thermal conductivity in 1T phase monolayers of HfSe2, SnSe2 and ZrS2.40,46 Additionally, the appearance of LLO phonons is the causative factor for the enhanced anharmonic scattering process observed in the heterostructures compared to their parent materials. In the case of MoSe2/WSe2 HS, it is notable that the scattering rate of LLO phonons (∼12.5 ps−1 at a frequency range of 150 cm−1) is more pronounced than acoustic and optical phonon modes, attributed to the phonon branching phenomena. Such dispersive LLO phonon branches get scattered by both acoustic and high-energy optical phonons due to the reduced frequency gap between them, resulting in a higher scattering rate of LLO modes. Hence, these highly dispersive phonon branches dominate three-phonon scattering, which primarily arises from the AOO process47 involving low-energy acoustic and optical modes within the frequency range of 150 cm−1. Moreover, the large optical bandwidth of the MoSe2/WSe2 HS further strengthens the AOO process to dominate the scattering rates even at low frequencies.
Furthermore, in the MoSeTe/WSeTe HS, the anharmonic scattering rate is higher across the entire frequency range, as evident from the phonon eigen modes shown in Fig. 4. In the MoSe2/WSe2 HS, the presence of chirality resulting from the breaking of spatial inversion symmetry compared to their homo-structure counterparts constrains the anharmonic interactions and reduces the phonon scattering rate in the system. Moreover, the doping of Te induces distinctive phonon phenomena (phonon bunching) in conjunction with mass anisotropy, which will substantially enhance the three-phonon scattering events that can consequently reduce the lattice thermal conductivity in MoSeTe/WSeTe compared to the MoSe2/WSe2 HS. The doping of such heavier Te atom induces substantial phonon softening, particularly affecting the acoustic and LLO phonon modes. This mass-induced phonon softening constraints the phonon frequency towards a low-energy region (discussed earlier in phonon spectra using a diatomic linear chain model), facilitating enhanced phonon–phonon interactions and, thereby, increasing the phonon scattering rate. Additionally, in the low frequency range below 35 cm−1, the MoSeTe/WSeTe HS exhibits a pronounced phonon scattering rate of ∼105 ps−1, attributed to the enhanced coupling of acoustic phonons themselves and with LLO modes, resulting from the bunching behaviour of phonon branches and strong acoustic phonon softening, due to the incorporation of Te atoms. Our results show that the bunching nature (non-dispersive) of acoustic phonon branches, combined with strong acoustic phonon softening within a very low frequency range of 75 cm−1, induces phonon–phonon interactions, primarily dominated by the AAA process.47 The anharmonic scattering values observed in the MoSeTe/WSeTe HS are higher than those previously reported for the MoS2–MoSe2 lateral superlattice, which shows a promising lattice thermal conductivity of ∼14.6 W m−1 K−1 among the 2H phase TMDs.38
To further understand the increased anharmonic scattering rate occurring in MoSeTe/WSeTe compared to the MoSe2/WSe2 heterostructure (as discussed above), we performed a microscopic analysis on weighted phase space (W3ph), a robust factor that describes the allowed phonon–phonon scattering limited by fundamental momentum and energy conservation conditions. Fig. S4 (ESI†) illustrates the weighted phase space as a function of frequency, depicting the available transition probabilities for phonon absorption (λ + λ′ → λ′′) and emission (λ → λ′ + λ′′) processes in both heterostructures. The transition probabilities for phonon–phonon scattering reach maximum values of 0.001 ps4 rad−4 (at 20 cm−1) and 0.08 ps4 rad−4 (at 10 cm−1) for MoSe2/WSe2 and MoSeTe/WSeTe HS, respectively. Interestingly, the W3ph plot for the MoSeTe/WSeTe HS clearly highlights the presence of significantly higher transition probabilities or scattering processes depicted as more dense points, particularly in the low-frequency region below 75 cm−1, attributed to the bunching of acoustic and LLO phonon branches. This bunching nature of phonon modes greatly increases the available phase space for three-phonon scattering processes, particularly the AAA scattering channels. The proximity of phonon branches constrained within a narrow frequency range enables a higher probability of scattering events in the structure, leading to a substantial increase in the anharmonic scattering rate. The presence of significant scattering channels combined with high transition probability values observed in the MoSeTe/WSeTe HS further substantiates the computed high anharmonic scattering rate in the system compared to the MoSe2/WSe2 HS. Moreover, the dominant scattering channels available in the MoSeTe/WSeTe HS can substantially suppress the heat carrying phonons, further leading to a low lattice thermal conductivity.
Additionally, the anharmonic scattering rate of the investigated heterostructures shows prominent scattering events in the acoustic and LLO phonon frequency ranges, indicated by highly dense points for both HS, as evident in Fig. 5a and b, compared to the previously reported WSe2 and WSe2/WTe2 bulk structures48 as shown in Fig. S5 of the ESI.† These intensified scattering events observed in MoSe2/WSe2 and MoSeTe/WSeTe heterostructures suggest enhanced phonon–phonon scattering, which suppresses the heat transport and probably contributes to a significantly low κl value.
In order to explain the curious behaviour of lattice thermal conductivity in the studied heterostructures, we evaluated the strength of anharmonic interactions occurring in the system by computing the mode-specific Grüneisen parameter (γ) expressed as follows:49
![]() | (2) |
To uncover the thermal transport mechanisms in both heterostructures, we also computed the phonon group velocity of each mode expressed as33 where ω, k and q denote the phonon frequency, vibrational mode index and wave vector, respectively. Fig. 5e and f illustrate the frequency-dependent phonon group velocities at 300 K for both MoSe2/WSe2 and MoSeTe/WSeTe heterostructures, respectively. Notably, the MoSe2/WSe2 HS possesses the highest phonon group velocity across the entire frequency region compared to the MoSeTe/WSeTe HS, in which a maximum group velocity of 8 km s−1 is observed for the LA mode, while the TA and quadratic ZA modes reach 4.1 km s−1 and 2.5 km s−1, respectively. Moreover, the LLO modes display high group velocities (7.8 km s−1), reaching as high as the longitudinal acoustic mode, arising due to their dispersive nature (as shown in Fig. 2c), allowing the LLO phonons to travel longer distances without scattering. The low group velocities of higher energy optical phonons indicate the dominating nature of acoustic and LLO phonon contribution to the lattice thermal conductivity. However, Te doping suppresses the phonon group velocities of all vibrational modes in the low-energy acoustic region due to increased scattering, visible in Fig. 5f, which results from the strong coupling nature of the acoustic branches themselves as well as with the LLO phonon branches and the enhanced phonon softening in the MoSeTe/WSeTe HS. Furthermore, the incorporation of Te in the MoSeTe/WSeTe HS reduces the frequency of LLO modes (especially, the LO mode as discussed earlier), arising due to phonon softening, which limits the group velocity of the LA branch to a minimum of ∼6 km s−1. Such LLO phonon softening also enhances the coupling between these branches, leading to a low group velocity (5.7 km s−1) for these modes. Consequently, the suppression of group velocity via Te substitution may significantly lower the κl value.
The temperature-dependent lattice thermal conductivity (κl) was calculated by solving the phonon Boltzmann transport equation (pBTE) in conjunction with the interatomic force constants (IFCs), as implemented in the ShengBTE package.22 The lattice thermal conductivity is given as follows:22
![]() | (3) |
Fig. 6a and b illustrate the lattice thermal conductivity as a function of temperature, calculated using both relaxation time approximation (RTA) and iterative methods. Our results reveal a very low lattice thermal conductivity for the MoSe2/WSe2 HS, with a minimum of ∼2.83 W m−1 K−1 using the iterative method and ∼2.42 W m−1 K−1 by RTA, whereas the MoSeTe/WSeTe HS shows an ultra-low value of 0.50 W m−1 K−1 using the iterative method and 0.142 W m−1 K−1 by RTA at room temperature. There is discrepancy in the κl values obtained from both methods, as RTA neglects the term Δλ, because the approximation considers the excitation of each individual phonon mode ignoring the phonon distribution.46 Such assumption works very well only for materials where normal scattering dominates the Umklapp scattering, which is typically applicable for high-thermal conductivity materials.46 In contrast, the iterative method, which accounts for the three-phonon scattering, solves the pBTE exactly by including both the Umklapp and normal scattering processes and provides a more accurately converged κl value, which is applicable for both high- and low-lattice thermal conductivity materials.46 At high temperatures, the mismatch of RTA values in predicting κl is less than 3%, as the Umklapp scattering becomes more significant. The computed lattice thermal conductivity values for both heterostructures are significantly lower than that of single-layer MoSe2 (∼43 W m−1 K−1)36,38,52 and WSe2 (∼52 W m−1 K−1)36,52 at room temperature. A previous study has shown that the creation of the MoS2–MoSe2 lateral superlattice increases the anharmonic scattering in the system, resulting in a promising κl value of 14.6 W m−1 K−1 at 300 K, which is nearly 1/5 and 1/4 of the two parent materials.38 Similarly, the modelling of the MoSe2/WSe2 HS has reduced the lattice thermal conductivity by ∼98% and ∼94% compared to the MoSe2 and WSe2 monolayer counterparts, respectively. The strategy of stacking MoSe2 over the WSe2 monolayer reduces the acoustic–optical phonon band gap (key factor for higher κl in the parent materials as discussed earlier), significantly enhances the three-phonon scattering dominated by the AOO process and consequently helps to achieve a low κl value of ∼1.24 W m−1 K−1 at 700 K (as shown in Fig. 6a), similar to those in the 1T phase of the HfSe2 and SnSe2 monolayers.40,46 Moreover, the substitution of Te in the MoSeTe/WSeTe HS amplifies the three-phonon scattering process along with strong anharmonic interactions arising due to the mass anisotropy and distinct phonon features such as phonon softening and phonon bunching in the system, leading to an ultra-low κl value of 0.5 W m−1 K−1 at room temperature. The low lattice thermal conductivity achieved by incorporating heavy Te atoms follows the traditional Keyes theory,53 which describes inverse relationship between κl and heavier elements. The presence of strongly coupled acoustic and LLO phonons in the low-frequency region, along with the high Grüneisen parameter, further increases the scattering rate dominated by the AAA process in the MoSeTe/WSeTe HS, consequently resulting in a very low κl value reaching a minimum of 0.21 W m−1 K−1 at 700 K, which is ∼83% lower than that of the MoSe2/WSe2 HS.
Furthermore, to provide a comprehensive understanding of the acoustic phonon contribution to κl in both heterostructures, we computed the lattice thermal conductivity using the Slack model, that assumes that the lattice thermal resistance arises only from the intrinsic phonon–phonon interaction, and is expressed as follows:54,55
![]() | (4) |
Furthermore, to understand the specific phonon contribution towards thermal conductivity, we plotted the cumulative lattice thermal conductivity as a function frequency for both the heterostructures, as shown in Fig. 6c and d. In the case of MoSe2/WSe2 HS, κl is mainly contributed by phonons within the frequency range of <150 cm−1, contributing 91.5% to the total lattice thermal conductivity. However, in the case of MoSeTe/WSeTe HS, the major contribution (88.2%) is constrained within a very low frequency range of <75 cm−1, suggesting the overall significance of low-energy phonon contribution in lattice thermal conductivity.
![]() | (5) |
The calculated m*, C2D and E1 values for electrons and holes in MoSe2/WSe2 and MoSeTe/WSeTe heterostructures are provided in Table S6 (ESI†). The variations in curvature observed in the band structures of MoSe2/WSe2 and MoSeTe/WSeTe heterostructures result in distinct effective masses for charge carriers. In the case of MoSe2/WSe2 HS, the m* values of electrons and holes were found to be 0.22m0 and 0.27m0, respectively, where m0 represents the bare electron mass. In the MoSeTe/WSeTe HS, the effective mass of electrons (∼0.23m0) is similar to that of the MoSe2/WSe2 HS, while the effective mass of holes shows a steady decline of 0.20m0. These results suggest that the substitution of Te instead of the Se atom does not significantly alter the curvature of CB, but induces slightly steeper bands in the valence band, contributing to a lower effective mass of holes in the MoSeTe/WSeTe HS. Table S6 (ESI†) presents the elastic constant values for both heterostructures, which are significantly higher than those reported for parent materials (MoSe2 and WSe2)57 and the previously reported MoS2/WS2 HS (∼259.95 J m−2).58 These higher C2D values, as shown in Table S6 (ESI†), indicate that these heterostructures exhibit enhanced ability to withstand external or deforming forces. Furthermore, the deformation potential constant was calculated under both tensile and compressive strain conditions. Notably, the magnitude of hole–phonon coupling strength consistently surpasses the electron–phonon coupling strength in both heterostructures, indicating higher mobility for electrons than that for holes. In the case of MoSeTe/WSeTe HS, the substitution of Te in place of the Se atom slightly reduces the E1 value for electrons (∼2.58 eV), suggesting a weaker acoustic phonon–electron coupling in MoSeTe/WSeTe than that in the MoSe2/WSe2 HS (∼2.71 eV). However, the Te substitution significantly increases the deformation potential constant (∼7.36 eV) for holes, suggesting a higher probability of scattering events to occur in MoSeTe/WSeTe than that in the MoSe2/WSe2 HS, which could potentially result in a lower hole mobility.
Fig. 7a and b illustrate the acoustic phonon limited mobility as a function of temperature for both heterostructures, and the corresponding values at room temperature are given in Table S6 (ESI†). The computed mobility values indicate that substituting Te for the Se atom results in a significant increase in the mobility of both electrons and holes in MoSeTe/WSeTe compared to the MoSe2/WSe2 heterostructure. Despite similar effective masses of electron in both the heterostructures, the pronounced increase in the electron mobility of MoSeTe/WSeTe (∼93664.03 cm2 V−1 s−1) compared to the MoSe2/WSe2 HS (∼49
157.20 cm2 V−1 s−1) is attributed to elevated C2D and E1. The computed μLA value for both the heterostructures is higher than that of their constituent monolayers57 and previously reported Janus 2D materials.57 Moreover, to understand the underlying mechanism of higher mobility in heterostructures, we computed the relaxation time for both the HS using the equation derived as,59
and the calculated relaxation time values for both MoSe2/WSe2 and MoSeTe/WSeTe heterostructures at 300 K are presented in Table S6 (ESI†). The observed relaxation time for both electrons (∼12.25 ps) and holes (∼1.73 ps) in MoSeTe/WSeTe are significantly higher than those of the MoSe2/WSe2 HS, indicating the reduced scattering events in the system, which allows the carriers to travel a longer distance without any hinderance, resulting in high carrier mobility in the MoSeTe/WSeTe HS.
In our study, the obtained DPT mobilities (Fig. 7a and b) are not quite promising, as only the scattering of charge carriers by LA phonons is considered and the contribution from optical phonons is completely neglected.57 In such scenario, considering the scattering arising from longitudinal optical (LO) phonons is inevitable to precisely determine the total mobility of the materials. The LO mode-resolved phonon mobility is given as follows:57
![]() | (6) |
Furthermore, the precise total mobility (μtotal) limited by both LA and LO phonons was computed using Matthiessen's rule derived as,57 and Fig. 7e and f represent the total mobility as a function of temperature. The results for MoSe2/WSe2 and MoSeTe/WSeTe heterostructures indicate that μtotal is primarily limited by the LO phonon scattering, consistent with previous studies on 2D TMDs.57 The computed μtotal values of HS is very close to the LO phonon limited mobility, henceforth, the omission of LO phonon scattering is the pivotal factor in the overestimation of DPT mobility by nearly an order of magnitude two in MoSe2/WSe2 and MoSeTe/WSeTe heterostructures, aligning with previously reported study.60 Interestingly, although the MoSeTe/WSeTe HS exhibits higher μLA, the incorporation of scattering from LO phonons results in a lower μtotal value for electrons (∼221.61 cm2 V−1 s−1) and holes (∼91.41 cm2 V−1 s−1) in MoSeTe/WSeTe than that of the MoSe2/WSe2 HS. This indicates that the elevated LO phonon scattering impedes the charge carrier transport in the MoSeTe/WSeTe HS.
In our present study, we employed the versatile software tool BoltzTraP26 to compute the temperature-dependent Seebeck coefficient (S). The Seebeck coefficient was obtained by solving the semi-classical Boltzmann transport equation within the constant scattering time approximation (CSTA), as implemented in the BoltzTraP code.26 In this theory, the Seebeck coefficient is expressed as follows:26
![]() | (7) |
The temperature-dependent Seebeck coefficients for p- and n-type MoSe2/WSe2 and MoSeTe/WSeTe heterostructures are shown in Fig. 8a and b, and the corresponding values are presented in Table S8 of the ESI.† Our findings reveal that the n-type Seebeck coefficient of the MoSe2/WSe2 HS (∼1316.53 μV K−1) is significantly higher than that of the MoSeTe/WSeTe heterostructure (∼671.56 μV K−1) at 300 K, attributed to the presence of heavy bands that result in an elevated carrier effective mass in the MoSe2/WSe2 HS. A similar pattern is observed for the p-type S values, where the MoSe2/WSe2 HS (∼1302.52 μV K−1) exhibits a value notably higher than that of the MoSeTe/WSeTe HS (∼612.74 μV K−1). The computed S values for the MoSe2/WSe2 HS are nearly twice as high as that of the MoSeTe/WSeTe structure, indicating a significantly higher thermopower. The obtained S values for the heterostructures are higher than those reported for the parent materials (MoSe2 and WSe2 monolayers).32 Notably, the S values for both p- and n-type heterostructures show a steady decline with the increase in temperature, which can be attenuated to the excitation of more carriers at higher temperatures. This trend of declining Seebeck values with the increase in temperature is consistent with the previously reported data.5
Furthermore, we evaluated the electrical conductivity,59σ = neμ, which is a transport coefficient property directly related to the mobility of charge carriers, to accurately compute the thermoelectric figure of merit, ZT. Here σ, n, e and μ represent the electrical conductivity, carrier concentration, charge of electron, and mobility of charge carriers, respectively. Fig. 8c and d illustrate the calculated electrical conductivity as a function of temperature for the p-type and n-type in the MoSe2/WSe2 and MoSeTe/WSeTe heterostructures. The optimisation of carrier concentration in conjunction with the Seebeck coefficient value plays a pivotal role in the fine-tuning of electrical conductivity to achieve more precise and accurate ZT values. In our study, we carefully optimised the carrier concentration in the range of 1017–1018 cm−3 and 1018–1019 cm−3 for MoSe2/WSe2 and MoSeTe/WSeTe heterostructures, respectively, aligning it with the previously reported data.32 The calculated electrical conductivity values show an increasing trend with temperature, as evident from Fig. 8c and d. The electrical conductivity values in the MoSeTe/WSeTe HS are significantly higher for both p-type (∼49.38 S cm−1) and n-type (∼87.97 S cm−1) than those of the MoSe2/WSe2 heterostructure at 700 K. The pronounced electrical conductivity observed for the MoSeTe/WSeTe HS over the entire temperature range is due to the enhanced carrier concentration, arising due to the low Seebeck coefficient in the system compared to the MoSe2/WSe2 HS.
Furthermore, we also computed the thermoelectric efficiency of both heterostructures using the following expression:61
![]() | (8) |
Footnote |
† Electronic supplementary information (ESI) available: The data used in and generated by this study. See DOI: https://doi.org/10.1039/d5cp02035b |
This journal is © the Owner Societies 2025 |