Juan M.
Marmolejo-Tejada
and
Martín A.
Mosquera
*
Department of Chemistry and Biochemistry, Montana State University, Bozeman, MT 59717, USA. E-mail: martinmosquera@montana.edu
First published on 20th May 2022
Two-dimensional (2D) quantum materials are poised to transform conventional electronics for a wide spectrum of applications that will encompass chemical sciences. For the study of thermal transport in single-layer (1L) or multi-layer transition metal dichalcogenides (TMDs), this work explores the combination of density functional theory (DFT) and algorithmic training for the generation of a moment tensor potential (MTP) that models 1L-MoS2, 1L-WS2 and their alloys, and demonstrates a synergy of theoretical techniques that is anticipated to play an important role in the field. From a high-performance computing perspective, these yield very convenient inter-atomic (or inter-molecular in other contexts) potentials that are useful to predict the response of quantum materials to thermal perturbations, or other driving forces. We show that our trained MTP functions successfully describe vibrational properties of the systems, and their thermal conductivities. The trained potential displays consistent agreement with DFT calculations, as well as the Stillinger–Weber (SW) potential. We also find that the thermal conductivity of the 2D alloys is little affected by sulfur vacancies. This is a behavior that may aid the fine-tuning of material's thermal properties for heat management and energy storage and conversion applications.
Recently, machine-learned interatomic potentials (MLIPs) have been explored for modeling large-scale molecular systems with high accuracy and significantly lower computational cost than first-principles methods.12,13 Although these potentials usually have very low transferability, they have a flexible functional form and can be systematically improved by increasing the training set and number of basis functions, but at the expense of higher computational cost.12,14–17 Moment tensor potentials (MTPs) describe atomic local environments throughout rotationally covariant tensors (formally, the moment of inertia)12 and have been shown to provide high accuracy and performance, when compared to competing MLIP models in the literature,17 and are expected to help understand the effects of structural and substitutional defects in the performance of TMD-based nanoelectronic devices.13,18
Non-equilibrium molecular dynamics (NEMD) methods have been used for estimating the thermal conductivity, κ, of semiconductor materials due to their direct analogy to experimental procedures.10 For this purpose, the central/scattering region is connected to two semi-infinite, defect-free regions, labeled as the heat source and drain terminals, respectively. Then, the momentum of the hottest atom in the drain is exchanged with the momentum of the coldest atom in the source, although conserving the total energy and momentum of the system.20,21 This produces a heat current from source to drain that results in a temperature gradient along the length of the system.
Alternatively, the non-equilibrium Greens function (NEGF) method has also been employed for calculating thermal conductance, σ, by obtaining the phonon transmission spectrum from the phonon Green's function G(ω)22 and calculating κ = σl/S, where l is the length of the sample and S is the cross-sectional area.23
This work aims at getting a better understanding of the effects of structural and substitutional defects in the performance of nanoelectronic devices and may also provide a tool for phononic engineering by controllably manipulating the thermal and electrical conductivity properties of nanostructured devices,24 where defect engineering may become a necessary step in finely tuning the thermoelectric properties of these devices.8,10,25
Similar to Mortazavi et al.,6,11 we passively train our potential, but include the pristine and alloyed configurations in the training set. This consists of 780 1L configurations from small random atomic displacements and molecular dynamics (MD) trajectories, as detailed in our Supporting Information (ESI†).
To assess the accuracy of our trained MTP, we compute vibrational modes and phonon band diagrams with the supercell method.26–28 For this, we diagonalize the dynamical matrix, which is constructed by small point displacements to the relaxed geometries. We use 21 × 21 × 1 repetitions for 1L-MoS2 and WS2, and 11 × 11 × 1 repetitions for the 1L-MoxW1−xS2 alloys. Then, we calculate the finite differences of the forces with the central finite difference method, using 0.01 Å atomic displacements and 1 × 10−8 Hartree Bohr−2 force tolerance, as implemented in QuantumATK.29,30
Fig. 1 shows the computed phonon band structure diagrams for the 1L-MoxW1−xS2 systems. We find very good agreement between DFT calculations and our MTP. This includes capturing a “negative” energy band at the Γ point in both DFT and MTP calculations; the negative band results from benign computational artifacts and does not translate into any anomalous instability.9,11,31,32 We verify this by fully relaxing the 1L-MoS2 and WS2 unit cells (see Fig. S2, ESI†), where the imaginary frequencies disappear, also noting that the frequency of the optical modes is reduced by ∼20–30 cm−1 with the functional and self-consistent criteria that were used in this study. Interestingly, our trained MTP agrees well with the lattice parameters of the relaxed structures and the resulting phonon band diagrams. To further improve the accuracy of calculations one can add MD trajectories at higher temperatures, relax the unit cells for stress, and increase the accuracy of the reference DFT calculations. However, the random displacement method is sufficient for the modeling performed in this work.
![]() | ||
Fig. 1 Validation of the trained MTPs: Comparison of different potentials for the determination of phonon dispersion diagrams for 1L- (a) MoS2, (b) WS2, (c) Mo0.75W0.25S2, (d) Mo0.5W0.5S2 and € Mo0.25W0.75S2. These diagrams are calculated with our trained MTP (green-dotted lines) and compared to DFT calculations (solid blue lines). Panel (a) also includes the phonon dispersion of 1L-MoS2 obtained with the SW potential in ref. 19 (orange-dashed lines). |
For additional reference, we also compute the phonon dispersion diagram of 1L-MoS2 with the Stillinger–Weber (SW) potential in ref. 19 (see Fig. 1(a)) and compare the optical vibrational modes in Fig. 2(a), finding good agreement with our trained MTP. Furthermore, we verify the evolution of the optical vibrational modes of the compositional-dependent MoxW1−xS2 systems in Fig. 2(b). Consistent with our previous work in ref. 33, we observe a blue-shift of the A2′′ mode (at 477 and 454 cm−1 for pristine 1L-MoS2 and WS2, respectively) and red-shift of the E′′ mode (at 288 and 307 cm−1 for pristine 1L-MoS2 and WS2, respectively) with increased Mo content.
![]() | ||
Fig. 2 (a) Optical vibrational modes of 1L-MoS2 obtained with our trained MTP and SW potential in ref. 19, and compared to DFT calculations. (b) Evolution of the optical absorption spectra for the 1L-MoxW1−xS2 structures, estimated with our trained MTP (symbols and solid lines) and compared to DFT calculations (dotted lines). |
Our inter-atomic potential was then used for estimating the thermal conductivity of 1L-MoS2, WS2 and their alloys, for which we consider sharp and alloy interfaces, as depicted in Fig. 3(a). We setup system configurations of length l = 16, 33, 66 and 132 nm for all cases. For S, we used a fixed width of 25.3 nm for NEMD and 7.58 nm for NEGF calculations (due to larger computational complexity) and assumed a monolayer thickness of 0.6033 nm.34
![]() | ||
Fig. 3 (a) Top view of the 1L-MoS2/WS2 heterostructure with a sharp or alloy interface. Blue, green and yellow spheres correspond to W, Mo and S atoms, respectively. (b) Phonon thermal conductivity of 1L-MoS2 at T = 300 K as a function of length, using the NEMD and NEGF methods. Our calculations with the trained MTP (solid lines) are compared to calculations with the SW potential (dashed lines) in ref. 19. (c) Phonon thermal conductivity of 1L-MoxW1−xS2 systems at T = 300 K as a function of length, obtained with our trained MTP and the NEMD method. The infinite thermal conductivity, κ∞, in panels (b) and (c) is estimated from the finite length results, assuming a phonon mean free path Λ = 100 nm for all systems. (d) Effect of sulfur vacancies on the thermal conductivity of 1L-MoS2, WS2 and Mo0.5W0.5S2 systems. Alloy interfaces show little variations in the estimated κ with both the NEMD (solid lines) and NEGF (dashed lines) methods. |
The conventional approach for estimating the thermal conductivity of infinite systems, κ∞ consists in extrapolating the thermal conductivity results of finite lengths, 1/κ = (1 + Λ/l)/κ∞,1,6,35 where l is the length of the sample and Λ is the effective phonon mean free path, assumed to be 100 nm in all cases.
As summarized in Table S3 (ESI†), previous measurements of κ for 1L-MoS2 at 300 K range between 34.5 ± 4 and 84 ± 17 W K−1m−1,36–38 while theoretical values range between 1.35 and 131 W K−1 m−1, where several methods have been employed, including the self-consistent solution of the Boltzmann transport equation (BTE),28,34,39 and the solution of Slack expression for the acoustic phonon modes,40 in addition to the NEMD41–43 and NEGF23 methods. Our MTP calculations result in κ of 43.2 and 63.4 W K−1 m−1 with the NEMD and NEGF methods, respectively, showing very good agreement with the range of experimental values and our analogous calculations with the SW potential. Our results of finite-length calculations and extrapolation of infinite-length κ are shown in Fig. 3(b).
Fig. 3(c) compares κ for the 1L-MoxW1−xS2 systems using the NEMD method. For 1L-WS2, there is only one reference experimental result for κ to the best of our knowledge, 32 W K−1 m−1,44 which is lower than the experimental references for MoS2. Our MTP calculations result in κ of 33.7 and 51.4 W K−1 m−1 with the NEMD and NEGF methods, respectively, noting that the NEMD result is in better agreement with the experimental value and reference calculations with the Slack expression.40 Then, we verify that 1L-MoS2 has higher κ than 1L-WS2, which is also related to its lower average atomic mass.40
Furthermore, we observe that κ for sharp interfaces is lower than that of the pristine materials, where Mo0.5W0.5S2 has a lower value among the studied concentrations. Alloy interfaces have slightly lower κ and have very low variability among the cases considered in this study, as shown in the inset of Fig. 3(c). This could be useful for engineering the thermal transport properties of other alloy configurations, independent of content concentrations.
Lastly, we add S vacancies to our systems and compute κ for small defect concentrations and a fixed l = 66 nm. As observed in Fig. 3(d), κ is reduced between 25–45% for the pristine materials with a defect density of 0.5%. This is consistent with previous studies where it was observed that vacancies in few-layer MoS2 produced a sharp decrease in κ, that drops even further with increasing defect concentrations due to phonon-defect scattering,25 but noting that very large defect densities result in a transition from a crystalline to an amorphous structure.25 On the other hand, alloy structures are more resilient to S vacancies, with as much as 13% reduction in κ at 0.5% defect density.
We have shown that MTPs can accurately reproduce phononic structure and thermal transport properties of pristine and alloyed 1L-TMDs. Although we find distinctive phonon spectrum features for 1L-MoxW1−xS2 alloys at different concentrations, we observe that thermal transport properties have very low dependence on the alloy content, making it possible to engineer the thermal transport properties of similar materials by mixing different atomic species. Furthermore, we found that alloy configurations are more resilient to S vacancies, providing new opportunities for the fabrication of novel devices without much dependence on costly techniques for obtaining defect-free crystalline materials. Lastly, precision largely depends on the reference DFT method and training set, making it possible to obtain reasonable results without the need for long optimization sequences and higher computational costs.6,11 Future work includes the application of MTPs to multilayer systems and other related quantum materials.
The authors thank Montana State University, Bozeman, for startup support and computational resources within the Tempest Research Cluster, and the Synopsys' QuantumATK development team for helpful discussions. M.A.M. acknowledges support from the MonArk NSF Quantum Foundry supported by the National Science Foundation Q-AMASE-i program under NSF award No. DMR-1906383.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2cc02519a |
This journal is © The Royal Society of Chemistry 2022 |