Next Article in Journal
Advancement of Marginal Bone and Soft Tissue Aesthetics for Slope-Configured Implants
Previous Article in Journal
Preparation and Properties of UV and Aziridine Dual–Cured Polyurethane Acrylate Emulsion
Previous Article in Special Issue
Layer by Layer Optimization of Langmuir–Blodgett Films for Surface Acoustic Wave (SAW) Based Sensors for Volatile Organic Compounds (VOC) Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Transport Properties of Na2X (X = O and S) Monolayers

1
Department of Orthopedics, Renmin Hospital, Wuhan University, Wuhan 430060, China
2
Key Laboratory of Artificial Micro- and Nano-Structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(9), 1294; https://doi.org/10.3390/coatings12091294
Submission received: 23 July 2022 / Revised: 22 August 2022 / Accepted: 26 August 2022 / Published: 2 September 2022
(This article belongs to the Special Issue Thick and Thin Films for Functional Device Applications)

Abstract

:
Motivated by the excellent functional thin film devices made from two-dimensional materials, we investigated the thermal transport properties of Na2X (X = O and S) monolayers using first-principle calculations. The thermal conductivity at room temperature was 1.055 W/mK and 1.822 W/mK for the Na2O monolayers and Na2S monolayers, respectively. The high thermal conductivity for the Na2S monolayers is mainly contributed to by in-plane transverse acoustic (TA) phonons. The group velocity for the Na2S monolayers exhibits lower group velocity and a larger phonon relaxation time than the Na2O monolayers. Our results are helpful for functional thin film devices made using Na2X (X = O and S) monolayers.

1. Introduction

Two-dimensional (2D) materials usually consist of multiple thin layers that are connected by weak interlayer van der Waals forces. Owing to their ultra-thin structures, 2D materials have attracted increasing amounts of attention in a number of research areas, including in thermoelectric [1], photodetector [2], and spintronic devices [3]. In practice, various synthesis techniques are applied to build 2D material-based functional nanoscale devices [4,5,6]. As a typical representative of 2D materials, graphene shows great potential in photonic and electronic devices [7]. For example, Wang et al. [8] utilized graphene as a high-impedance surface to build reconfigurable terahertz leaky-wave antennae. The tunable conductivity of graphene can effectively modulate the reflection phase, resonant frequency, and radiation pattern. The unique physical and chemical properties of 2D materials make them well-suited for use in functional nanoscale devices.
Investigating new 2D functional materials has become one of the more popular topics in materials science [9]. Some 2D materials that have already been applied to fabricate thin film membranes include graphene [10], MoS2 [11], g-C3N4 [12], h-BN [13], and MXene [14]. Due to its zero band gap [15] and high thermal conductivity [16], graphene has been found to be inappropriate for photovoltaic and thermoelectric applications. Moreover, these 2D materials are easily oxidized in the air. Therefore, plenty of efforts have made to investigate and fabricate stable 2D materials [17,18].
Among these new 2D materials, dialkali-metal monochalcogenide monolayers provide new opportunities and challenges for thermoelectric and photovoltaic devices. Hua et al. [19] reported that dialkali-metal monochalcogenide monolayers exhibit a 2D gate-tunable magnetism. They also found that dialkali-metal monoxide monolayers host multiple symmetry-protected topological phases that can be tuned by strain engineering [20]. Recently, Rawat et al. [21] investigated the photovoltaic properties of M2X (M = Na, K, and Cs; X = O, S, Se, and Te) monolayers using density functional theory. They found that the van der Waals heterostructures of K2O/Cs2O, K2S/Cs2S, and K2Se/Cs2S exhibit high power conversion efficiency in tandem solar cells. These theoretical works prove that dialkali-metal monochalcogenide monolayers exhibit good semiconductor properties and show great potential in energy harvesting.
To achieve thermoelectric and photovoltaic applications, thermal transport properties are essential. Since there are many types of dialkali-metal monochalcogenide monolayers, we chose Na2O and Na2S monolayers as the research objects after a structure stability test. In this work, we present a theoretical prediction of the thermal transport properties in Na2X (X = O and S) monolayers. Our results show that Na2O and Na2S monolayers exhibit low thermal conductivity.

2. Materials and Methods

All first-principle calculations were completed using the Quantum ESPRESSO package [22,23]. The generalized gradient approximation (GGA) of the Perdew–Burke–Ernzerhof (PBE) [24] function was utilized, and the corresponding pseudopotential files were obtained from the standard solid-state pseudopotentials library [25]. To obtain the fully relaxed structures, the energy and force convergence threshold were set as 10−6 Ry and 10−4 Ry, respectively. The kinetic energy cutoff was set as 80 Ry, and the k-mesh was set as 13 × 13 × 1. We set the vacuum spacing to be 15 Å to avoid interlayer interactions. Spin–orbit coupling (SOC) has negligible effects and was not included in this work. The Heyd–Scuseria–Ernzerhof (HSE06) hybrid function [26] was added while calculating band structure.
The phononic thermal conductivity κph was calculated using the phonon Boltzmann transport theory [27]. In the phonon Boltzmann transport theory, κph is obtained by
κ p h = 1 N V λ κ λ = 1 N V λ C λ v λ v λ τ λ
where N is the total number of phonon modes, V is the volume, C λ is the heat capacity, v λ is the phonon group velocity, and τ λ is the phonon relaxation time. For small displacements, the total potential energy can be expanded into
E = E 0 + 1 2 ! i j , α β Φ i j α β u i α u j β + 1 3 ! i j k , α β γ Φ i j k α β γ u i α u j β u k γ +
where u i α is the displacement of atom i along the α(x, y, z) direction. The coefficients Φ in the n-th order term are the corresponding n-th order interatomic force constants (IFCs). Calculating higher-order interatomic force constants requires a lot of time [28]. Generally, third-order IFCs are enough to obtain suitable results. A supercell-based finite-difference method was used to calculate the third-order IFCs. Under a small displacement h , Φ i j k α β γ is expressed as
Φ i j k α β γ = 3 E u i α u j β u k γ 1 2 h [ 2 E u j β u k γ ( u i α = h ) 2 E u j β u k γ ( u i α = h ) ]
The Grüneisen parameter γ can also be calculated from third-order IFCs.
γ = 1 6 ω i j k , α β γ Φ i j k α β γ e i α * e j β m i m j e x p ( i q · R j ) u k γ
where m i is the mass of atom i , e is the phonon eigenvector, and R is the unit cell vector. Details about phonon Boltzmann transport theory can be found in Refs. [27,28,29].
Phononic thermal conductivity was calculated using the ShengBTE code [27]. The phonon dispersions and harmonic interatomic force constants (second-order IFCs) were calculated using the PHONOPY package [30] with a 7 × 7 × 1 supercell (147 atoms in total). The anharmonic interatomic force constants (third-order IFCs) were calculated using the maximum atomic interaction distance of the ninth neighbor with a 7 × 7 × 1 supercell. The q-point grid was set to 60 × 60 × 1, and the smearing parameter was set as 1.0.

3. Results

In Figure 1a, we present different views of Na2X (X = O and S) monolayers. The structure of the Na2X (X = O and S) monolayers is the same as that of the MoS2 monolayers. The space group is P-6m2 (No. 187), and the point group symmetry is D3h. Along the z-axis, the atomic layers follow the sequence Na-X-Na (X = O and S). As depicted in Figure 1a, we defined the thickness d for the distance between two Na atoms along the z-axis. The relaxed lattice parameter a = 3.451 Å and 4.249 Å and the relaxed thickness d = 2.306 Å and 2.581 Å for the Na2O and Na2S monolayers, respectively. Since the transport properties along the x-axis and y-axis are the same, we only present the transport properties along the x-axis in the following discussion.
We present the band structures in Figure 1b,c. It can be noted that the band structures for the Na2O and Na2S monolayers are nearly the same, except for the band gaps. The band gaps are 0.251 eV and 1.467 eV for the Na2O and Na2S monolayers, respectively. The Na2O monolayers exhibit stronger metallicity than the Na2S monolayers. The band gap can reflect the energy needed to break the bond. This confirms the normalized electron localization function shown in Figure 1d,e. The values of 0 and 1 represent accumulated and vanishing electron densities [31]. It can be seen that the electron densities are highly localized around the X (X = O and S) atoms and that the distributions between the Na atoms and X (X = O and S) atoms are nearly negligible. This means that the Na-X (X = O and S) bonds are ionic. Apparently, the strength of a Na-O bond is weaker than that of a Na-S bond. The weaker bond strength corresponds to a lower κph [32].
To evaluate the thermal transport properties, we first calculated the phonon dispersions of the Na2X (X = O and S) monolayers in Figure 2a,b. Obviously, there is no imaginary mode that implies the dynamic stability of the Na2O and Na2S monolayers. It can be noted that the phonon dispersion in the Na2O monolayers is similar to that of Na2S monolayers. There is a small gap in the Na2O monolayers, and this small gap disappears in the Na2S monolayers. The cutoff frequencies are 400 cm−1 and 275 cm−1 for the Na2O and Na2S monolayers, respectively. We also calculated the corresponding projected density of states (DOS) in Figure 2c,d. In the low frequency range, the DOS is contributed to by Na and X (X = O and S) atoms equally. For the middle frequency range, the DOS is mainly contributed to by Na atoms. At the high frequency range, the DOS for the Na2O monolayers is occupied by O atoms, and the DOS for the Na2S monolayers is occupied by Na and S atoms equally.
After obtaining the second- and third-order IFCs, we can calculate the thermal conductivity κph as a function of temperature, as shown in Figure 3a. Generally, the κph vlaies of the Na2S monolayers are higher than those of the Na2O monolayers. For example, the κph values for the Na2O and Na2S monolayers are 1.055 W/mK and 1.822 W/mK at room temperature, respectively. These κph values are much lower than those of MoS2 monolayers [33], which implies the potential thermoelectric properties of Na2X (X = O and S) monolayers. The κph decreases as the temperature increases and follows a 1 / T dependence. This is caused by the enhanced phonon–phonon scattering with temperature. When the temperature increases to 700K, the κph is reduced to 0.452 W/mK and 0.783 W/mK for the Na2O and Na2S monolayers, respectively. We also discussed the effect of system size on the phononic thermal transport properties. When the size of the system reduces to the mean free path level, boundary scattering is enhanced and leads to a lower κph. Figure 3b displays the cumulative κph as a function of the mean free path. The κph for the Na2O monolayers increases faster than that of the Na2S monolayers as the mean free path increases. This means that reducing the κph of Na2O monolayer requires a smaller particle size than Na2S monolayers do. To provide further discussion, we display the κph contributed to by the out-of-plane flexural acoustic (ZA), in-plane transverse acoustic (TA), in-plane longitudinal (LA) acoustic, and optical modes in Figure 3c. The TA mode occupies the main part of the κph for the Na2X (X = O and S) monolayers. The κph values occupied by optical modes for the Na2X (X = O and S) monolayers are similar.
To analyze the differences in the thermal transport properties, we present the phononic-related parameters for the Na2X (X = O and S) monolayers in Figure 4. Because the shapes of the phonon dispersions for the Na2O and Na2S monolayers are similar, the group velocity also shows similar shapes, as shown in Figure 4a. Na2O monolayers possess higher group velocity than the Na2S monolayers. According to the Equation (1), a higher group velocity means higher thermal conductivity. However, this is not consistent with the above κph results (the κph of Na2S is higher than that of Na2O). The phonon relaxation time is the main reason for this. In most frequency ranges, the phonon relaxation time for the Na2S monolayers is higher than that of the Na2O monolayers, as seen in Figure 4b.
The scattering rate 1 / τ can be reflected by the P3 parameter and Grüneisen parameter γ [34]. In three-phonon scattering processes, energy ( ω j ( q ) ± ω j ( q ) = ω j ( q ) ) and momentum ( q ± q = q + G ) conservation are required [35]. This limits the phase space available for the three-phonon scattering processes. The P3 parameter represents the available phase space for anharmonic three-phonon scattering. The square of the γ parameter represents the strength of anharmonic scattering [27]. We present the P3 and γ parameters contributed by different modes in Figure 5. The overall P3 parameter of the Na2S monolayers is much higher than that of the Na2O monolayers, especially for the acoustic part. The phonon modes for the Na2S monolayers provide more channels for anharmonic scattering than the Na2O monolayers. As for scattering strength, it can be noticed that the γ from the ZA modes of the Na2S monolayers is negative, which is common in 2D materials [36]. However, the γ from the ZA modes of the Na2S monolayers is much higher than in the other cases, as mentioned above. We noticed that the phonon dispersion of the ZA modes from Γ to K in the Na2S monolayers is softer compared to in the Na2O monolayers. Thus, the high κph value from TA modes in the Na2S monolayers arises from the increased available phase space for anharmonic scattering by S atoms.

4. Conclusions

In summary, we calculated the thermal conductivity of Na2X (X = O and S) monolayers using first-principle calculations. The shapes of the band structures and the phonon dispersions for the two structures were similar but resulted in different thermal transport properties. The thermal conductivity of the Na2S monolayers is higher than that of the Na2O monolayers and mainly contributes to the TA mode. The room temperature κph values are 1.055 W/mK and 1.822 W/mK for the Na2O and Na2S monolayers, respectively. This work provides a theoretical guide for thermal management in electronic or thermoelectric devices based on Na2O and Na2S monolayers.

Author Contributions

X.Y. and W.C. conceived the study. W.C. performed the calculations and analyzed the data. X.Y. and H.L. wrote, reviewed, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities (No. 2042022kf1078) and the Knowledge Innovation Program of Wuhan-Basic Research (No. 2022020801010473).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data provided in this study can be released upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kanahashi, K.; Pu, J.; Takenobu, T. 2d materials for large-area flexible thermoelectric devices. Adv. Energy Mater. 2020, 10, 1902842. [Google Scholar] [CrossRef]
  2. Long, M.; Wang, P.; Fang, H.; Hu, W. Progress, challenges, and opportunities for 2d material based photodetectors. Adv. Funct. Mater. 2019, 29, 1803807. [Google Scholar] [CrossRef]
  3. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.-J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, J.; Zhou, M.; Wu, W.; Tang, Y. Fabrication of diamond microstructures by using dry and wet etching methods. Plasma Sci. Technol. 2013, 15, 552–554. [Google Scholar] [CrossRef]
  5. Pardo, D.A.; Jabbour, G.E.; Peyghambarian, N. Application of screen printing in the fabrication of organic light-emitting devices. Adv. Mater. 2000, 12, 1249–1252. [Google Scholar]
  6. Choi, D.-G.; Jeong, J.-h.; Sim, Y.-s.; Lee, E.-s.; Kim, W.-S.; Bae, B.-S. Fluorinated organic−inorganic hybrid mold as a new stamp for nanoimprint and soft lithography. Langmuir 2005, 21, 9390–9392. [Google Scholar] [CrossRef]
  7. Wang, R.; Ren, X.-G.; Yan, Z.; Jiang, L.-J.; Sha, W.E.I.; Shan, G.-C. Graphene based functional devices: A short review. Front. Phys. 2018, 14, 13603. [Google Scholar] [CrossRef]
  8. Wang, X.C.; Zhao, W.S.; Hu, J.; Yin, W.Y. Reconfigurable terahertz leaky-wave antenna using graphene-based high-impedance surface. IEEE Trans. Nanotechnol. 2015, 14, 62–69. [Google Scholar] [CrossRef]
  9. Haastrup, S.; Strange, M.; Pandey, M.; Deilmann, T.; Schmidt, P.S.; Hinsche, N.F.; Gjerding, M.N.; Torelli, D.; Larsen, P.M.; Riis-Jensen, A.C.; et al. The computational 2d materials database: High-throughput modeling and discovery of atomically thin crystals. 2D Mater. 2018, 5, 042002. [Google Scholar]
  10. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar]
  11. Sun, L.; Huang, H.; Peng, X. Laminar mos2 membranes for molecule separation. Chem. Commun. 2013, 49, 10718–10720. [Google Scholar] [CrossRef] [PubMed]
  12. Ding, F.; Yang, D.; Tong, Z.; Nan, Y.; Wang, Y.; Zou, X.; Jiang, Z. Graphitic carbon nitride-based nanocomposites as visible-light driven photocatalysts for environmental purification. Environ. Sci. Nano 2017, 4, 1455–1469. [Google Scholar] [CrossRef]
  13. Abdikheibari, S.; Lei, W.; Dumée, L.F.; Milne, N.; Baskaran, K. Thin film nanocomposite nanofiltration membranes from amine functionalized-boron nitride/polypiperazine amide with enhanced flux and fouling resistance. J. Mater. Chem. A 2018, 6, 12066–12081. [Google Scholar] [CrossRef]
  14. Karahan, H.E.; Goh, K.; Zhang, C.; Yang, E.; Yıldırım, C.; Chuah, C.Y.; Ahunbay, M.G.; Lee, J.; Tantekin-Ersolmaz, Ş.B.; Chen, Y.; et al. Mxene materials for designing advanced separation membranes. Adv. Mater. 2020, 32, 1906697. [Google Scholar] [CrossRef]
  15. Das Sarma, S.; Adam, S.; Hwang, E.H.; Rossi, E. Electronic transport in two-dimensional graphene. Rev. Mod. Phys. 2011, 83, 407–470. [Google Scholar] [CrossRef]
  16. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef]
  17. Si, J.; Yu, J.; Shen, Y.; Zeng, M.; Fu, L. Elemental 2d materials: Progress and perspectives toward unconventional structures. Small Struct. 2021, 2, 2000101. [Google Scholar] [CrossRef]
  18. Geng, D.; Yang, H.Y. Recent advances in growth of novel 2d materials: Beyond graphene and transition metal dichalcogenides. Adv. Mater. 2018, 30, 1800865. [Google Scholar] [CrossRef]
  19. Hua, C.; Sheng, F.; Hu, Q.; Xu, Z.-A.; Lu, Y.; Zheng, Y. Dialkali-metal monochalcogenide semiconductors with high mobility and tunable magnetism. J. Phys. Chem. Lett. 2018, 9, 6695–6701. [Google Scholar] [CrossRef]
  20. Hua, C.; Li, S.; Xu, Z.-A.; Zheng, Y.; Yang, S.A.; Lu, Y. Tunable topological energy bands in 2d dialkali-metal monoxides. Adv. Sci. 2020, 7, 1901939. [Google Scholar] [CrossRef]
  21. Rawat, A.; Arora, A.; De Sarkar, A. Interfacing 2d M2X (M = Na, K, Cs; X = O, S, Se, Te) monolayers for 2d excitonic and tandem solar cells. Appl. Surf. Sci. 2021, 563, 150304. [Google Scholar] [CrossRef]
  22. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. Quantum espresso: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  23. Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Buongiorno Nardelli, M.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced capabilities for materials modelling with quantum espresso. J. Phys. Condens. Matter 2017, 29, 465901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  25. Prandini, G.; Marrazzo, A.; Castelli, I.E.; Mounet, N.; Marzari, N. Precision and efficiency in solid-state pseudopotential calculations. Npj Comput. Mater. 2018, 4, 72. [Google Scholar] [CrossRef]
  26. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  27. Li, W.; Carrete, J.; Katcho, N.A.; Mingo, N. Shengbte: A solver of the boltzmann transport equation for phonons. Comput. Phys. Commun. 2014, 185, 1747–1758. [Google Scholar] [CrossRef]
  28. Han, Z.; Yang, X.; Li, W.; Feng, T.; Ruan, X. Fourphonon: An extension module to shengbte for computing four-phonon scattering rates and thermal conductivity. Comput. Phys. Commun. 2022, 270, 108179. [Google Scholar] [CrossRef]
  29. Broido, D.A.; Ward, A.; Mingo, N. Lattice thermal conductivity of silicon from empirical interatomic potentials. Phys. Rev. B 2005, 72, 014308. [Google Scholar] [CrossRef]
  30. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  31. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. Elf: The electron localization function. Angew. Chem. Int. Ed. Engl. 1997, 36, 1808–1832. [Google Scholar] [CrossRef]
  32. Zeier, W.G.; Zevalkink, A.; Gibbs, Z.M.; Hautier, G.; Kanatzidis, M.G.; Snyder, G.J. Thinking like a chemist: Intuition in thermoelectric materials. Angew. Chem. Int. Ed. 2016, 55, 6826–6841. [Google Scholar] [CrossRef] [PubMed]
  33. Gandi, A.N.; Schwingenschlögl, U. Thermal conductivity of bulk and monolayer mos2. Europhys. Lett. 2016, 113, 36002. [Google Scholar] [CrossRef]
  34. Ziman, J.M. Electrons and Phonons: The Theory of Transport Phenomena in Solids; Oxford University Press: Oxford, UK, 2001; p. 288. [Google Scholar]
  35. Lindsay, L.; Broido, D.A. Three-phonon phase space and lattice thermal conductivity in semiconductors. J. Phys. Condens. Matter 2008, 20, 165209. [Google Scholar] [CrossRef]
  36. Sevik, C. Assessment on lattice thermal properties of two-dimensional honeycomb structures: Graphene, h-bn, h-mos2, and h-mose2. Phys. Rev. B 2014, 89, 035422. [Google Scholar] [CrossRef]
Figure 1. (a) Top and side views of Na2X (X = O and S) monolayers. The electronic band structures of (b) Na2O and (c) Na2S monolayers. The normalized electron localization function for (d) Na2O and (e) Na2S monolayers.
Figure 1. (a) Top and side views of Na2X (X = O and S) monolayers. The electronic band structures of (b) Na2O and (c) Na2S monolayers. The normalized electron localization function for (d) Na2O and (e) Na2S monolayers.
Coatings 12 01294 g001
Figure 2. The phonon dispersions of (a) Na2O and (b) Na2S monolayers. The corresponding projected density of states of (c) Na2O and (d) Na2S monolayers.
Figure 2. The phonon dispersions of (a) Na2O and (b) Na2S monolayers. The corresponding projected density of states of (c) Na2O and (d) Na2S monolayers.
Coatings 12 01294 g002
Figure 3. (a) The thermal conductivity of Na2O and Na2S monolayers as a function of temperature. (b) The mean free path dependence of the thermal conductivity for Na2O and Na2S monolayers at room temperature. (c) The thermal conductivity at room temperature contributed from ZA, TA, LA, and optical modes.
Figure 3. (a) The thermal conductivity of Na2O and Na2S monolayers as a function of temperature. (b) The mean free path dependence of the thermal conductivity for Na2O and Na2S monolayers at room temperature. (c) The thermal conductivity at room temperature contributed from ZA, TA, LA, and optical modes.
Coatings 12 01294 g003
Figure 4. The (a) group velocity and (b) phonon relaxation times for Na2O and Na2S monolayers at room temperature.
Figure 4. The (a) group velocity and (b) phonon relaxation times for Na2O and Na2S monolayers at room temperature.
Coatings 12 01294 g004
Figure 5. The (a,b) phase space for anharmonic scattering and (c,d) Grüuneisen parameter for different modes for Na2O and Na2S monolayers at room temperature.
Figure 5. The (a,b) phase space for anharmonic scattering and (c,d) Grüuneisen parameter for different modes for Na2O and Na2S monolayers at room temperature.
Coatings 12 01294 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, X.; Cao, W.; Li, H. Thermal Transport Properties of Na2X (X = O and S) Monolayers. Coatings 2022, 12, 1294. https://doi.org/10.3390/coatings12091294

AMA Style

Yan X, Cao W, Li H. Thermal Transport Properties of Na2X (X = O and S) Monolayers. Coatings. 2022; 12(9):1294. https://doi.org/10.3390/coatings12091294

Chicago/Turabian Style

Yan, Xinxin, Wei Cao, and Haohuan Li. 2022. "Thermal Transport Properties of Na2X (X = O and S) Monolayers" Coatings 12, no. 9: 1294. https://doi.org/10.3390/coatings12091294

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop