Next Article in Journal
Thin-Film Coating Methods: A Successful Marriage of High-Quality and Cost-Effectiveness—A Brief Exploration
Next Article in Special Issue
Recent Advancements in Surface Modification, Characterization and Functionalization for Enhancing the Biocompatibility and Corrosion Resistance of Biomedical Implants
Previous Article in Journal
Modification of Cold-Sprayed Cu-Al-Ni-Al2O3 Composite Coatings by Friction Stir Technique to Enhance Wear Resistance Performance
Previous Article in Special Issue
Effect of La2O3 on Microstructure and Properties of Laser Cladding SMA Coating on AISI 304 Stainless Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Calculations to Investigate the Oxidation Mechanism of Pristine MoS2 and Ti-Doped MoS2

1
School of Materials and Chemical Engineering, Tongren University, Tongren 554300, China
2
School of Materials Science and Engineering, East China Jiaotong University, Nanchang 330013, China
3
College of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
4
Institute for Nanotechnology and Water Sustainability, College of Science, Engineering and Technology, University of South Africa, Johannesburg 1710, South Africa
5
Chemistry Department, College of Science, King Khalid University, Abha 61413, Saudi Arabia
6
Chemistry Department, Faculty of Sciences, University of Sfax, Sfax 3038, Tunisia
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(8), 1114; https://doi.org/10.3390/coatings12081114
Submission received: 13 July 2022 / Revised: 27 July 2022 / Accepted: 31 July 2022 / Published: 4 August 2022
(This article belongs to the Special Issue Recent Advances in Functional Surfaces and Interfaces)

Abstract

:
Generally, MoS2 is easily oxidized when exposed to oxygen, and the antioxidation mechanism of MoS2 is still a challenge. Thus, more efforts were made to greatly improve its antioxidation performance. It was reported that the Ti atom doped with MoS2 was treated as the effective method to enhance its antioxidation performance; however, the detailed antioxidation mechanism was not well understood. Superior to experimental methods, the first-principles method could provide deep insight into the atomic information and serve as a useful tool to gain an understanding of the antioxidation mechanisms of the doped MoS2; thus, the antioxidation behavior of the Ti-doped MoS2 was investigated in detail using first-principles calculations. However, an opposing conclusion was obtained from the calculated results compared to the previous experimental results; that is, the incorporation of the Ti atom was not helpful for improving the antioxidation performance of MoS2. The strange phenomenon was well probed and discussed in detail, and understanding the oxidation mechanism of the Ti-doped MoS2 would be helpful for expanding its applications in the ambient atmosphere.

1. Introduction

A high-quality MoS2 sheet was successfully deposited using the chemical vapor deposition system [1], which has been applied in many fields due to its excellent optical [2,3], magnetic [4], electrocatalytic and photocatalytic [5,6], and mechanical [7] properties. In particular, the ultrathin thickness of MoS2 endowed the excellent tribological performance [8,9] due to its easily shearing. However, its tribological behaviors were closely sensitive to the environment conditions [10] because the formation of MoO3 would result in an increasing friction coefficient when exposed to oxygen [11]. Ren et al. recently reported that MoS2-based superlattice films exhibited superlubricity (0.006) in a low vacuum while showing a high friction coefficient (>0.04) in the air [12]. The poor antioxidation of MoS2 limited its further applications in the air; thus, improving its antioxidation performance was imperative.
The oxidation mechanism of MoS2 should initially be well understood [13]. It is a well-established oxidation mechanism that, owing to no dangling bonds on the MoS2 plane that are terminated by S atoms, when it is exposed to air, it leads to the formation of defects that enhance its surface activity, further resulting in the strong interactions between defected MoS2 and oxygen [14,15,16]. Indeed, Pu et al., using the first-principles method, confirmed that the defects in the MoS2 was a virulent attack on its related properties and led to failure [17]. It is worth noting that the metal-doped MoS2 was found as an efficient way to enhance antioxidation [18,19,20,21,22]. Ti-doped MoS2 has been confirmed to slow down its failure in ambient atmosphere [23,24,25] while the real interactions mechanism between the oxygen and the Ti-doped MoS2 were rarely reported, theoretically. Thus, one wonders if the incorporation of the Ti element really slowed down the failure of MoS2 in the ambient atmosphere.
Thereafter, in this work, interactions between pristine MoS2, Ti-doped MoS2, and O atoms were systematically investigated using first-principles calculations in order to obtain the proposed antioxidation mechanism. The complex microscopic antioxidation mechanism was revealed by the related calculations. However, we obtained opposing results compared to the previous reports [23,24,25]; that is, the incorporation of the Ti element was not conducive to improving the antioxidation performance of MoS2.

2. Computational Methods

As known, the first-principles method is treated as an effective tool to investigate the various properties of solid materials at an atomic level [26,27,28,29]. All calculations were performed using the CASTEP code [30]. The calculated model of pristine MoS2 (1T) is shown in Figure 1. The electronic configurations of Mo, S, and Ti was 4d5 5s1, 3s2 3p4, and 3s2 3p6 3d2 4s2, respectively. The cutoff energy of 400 eV was selected for the plane-wave expansion, and the k-point of the Brillouin zone was sampled with a 5 × 5 × 10 Monkhorst-Pack grid. The convergence of energy, force, stress, and displacement was set as 2.0 × 10−6 eV/atom, 5.0 × 10−2 eV/Å, 0.1 GPa, and 2.0 × 10−3 Å. To minimize interlayer interaction, MoS2 sheets in the neighboring cells were separated by a distance of at least 20 Å. The generalized gradient approximation (GGA) with Perdew−Burke−Ernzerhof (PBE) [31] was selected for all calculations. The ultrasoft pseudopotential was selected to manage the interaction between the ionic and valence electrons [32]. The phonon dispersion and related density of state of Ti-doped MoS2 were calculated to determine the dynamical stability of the doped models [33,34]. The finite displacement method was used to calculate phonons.

3. Results and Discussion

3.1. Phonon Dispersion

To investigate the stability of Ti-doped MoS2 (the S or Mo atom is replaced with the Ti atom), called MSTi-S and MSTi-Mo, the phonon dispersion curves are calculated as shown in Figure 2. Figure 2a shows the phonon dispersion of MoS2, which is in accord with the calculated results reported by Tornatzky et al. [35]. Generally, the occurrence of imaginary frequency corresponds to the dynamical instability [36]. Figure 2b displays the phonon dispersion of MSTi-S. Several imaginary frequencies are observed in MSTi-S, revealing the dynamical instability. Figure 2c shows the phonon dispersion of MSTi-Mo. No imaginary frequency indicates that MSTi-Mo is the dynamical stability, which is consistent with the previous work [37]. Thus, subsequently, the adsorption behaviors of O atoms are just investigated for the dynamical stable MoS2 and MSTi-Mo.

3.2. Adsorption Energy

To better understand the antioxidation mechanism of Ti-doped MoS2, we explore the stable adsorption configuration of the O atom on a MoS2(002) slab. The stability of the O atom is determined by the adsorption energy ( E a d O ) [38], calculated by:
E a d O = E tot ( MoS 2 ( 002 ) / O ) E MoS 2 ( 002 ) μ ( O )
where E tot ( MoS 2 ( 002 ) / O ) , E MoS 2 ( 002 ) , and μ ( O ) are the total energy of the oxygen adsorption system, the clean relaxed MoS2(002) surface, and the chemical potential of the O atom. Figure 3 shows the oxygen-adsorbed configuration on the MoS2(002) surface. Four possible adsorbed sites are considered to search the most stable adsorbed sites: the Mo top site, S top site, dopant top site, and hollow site (H site). To the best of our knowledge, the stability of the adsorbed site can be reflexed from the adsorption energy [39]. The lower adsorption energy corresponds to the good stability of the adsorbed materials.
The calculated E a d O values are listed in Table 1. The calculated ones of those adsorbed sites for the MoS2(002) surface are less than zero, which indicates that O atoms easily adsorb onto the MoS2(002) surface. The calculated E a d O for the MSTi-Mo(002) surface is −1.33, −0.70, −1.90, and −1.04 eV for the H, Mo top, S top, and Ti top sites, respectively. Generally, a more negative E a d O value reveals a stronger interaction between the adsorbent and the adsorbate [40,41]. The adsorption energies of O adsorbed on the MoS2(002) surface are more negative than that of the MSTi-Mo(002) surface, indicating a weaker interaction between the O atom and atoms in the MSTi-Mo(002) surface. As shown in Table 1, the distance between the Mo and O atom is 2.204 and 2.336 Å for the MoS2: Mo top site and MSTi-Mo: H site, respectively, which is larger than that of the calculated equilibrium bond length of the Mo−O bond (2.08 Å) [42], indicating that the adsorption of the O atom onto the Mo top site and H site (MSTi-Mo) results in a breaking of the Mo−S bond, leading to the damage of the MoS2 nanosheet. In Figure 4, for the H site of the MoS2(002) surface, the O atom moves from the hollow position to the S top site, forming the covalent bond; the nanosheet structure is maintained perfectly after the adsorption of the O atom. The distance between the Ti and O atoms is 2.010 and 2.062 Å for the MSTi-Mo: H site and MSTi-Mo: Ti top site, respectively, which is more than that of the reported equilibrium bond length of the Ti−O bond (1.933 Å) [43], indicating that no chemical bonds are formed between the Ti and O atoms. In Figure 5, for the H site of the MSTi-Mo(002) surface, after the O adsorption, the nanosheet structure is damaged after an attack of the O atom; the same phenomenon is found as the O atom adsorbed above the Ti top site. The nanosheet structure remains in good condition as the O atom adsorbed above the Mo and S top site. It can be concluded that the incorporation of Ti into MoS2 only improves the oxidation resistance as O adsorbed on the Mo site.

3.3. Charge Density Difference

To further investigate the nature of the charge transfer and chemical bonds of the O-adsorbed MoS2(002) and MSTi-Mo(002) surface, the charge density difference is analyzed and discussed as displayed in Figure 6. Because the electronegativity of the O atom is stronger than that of S atoms [44], the O atom can gain more valence charge from S atoms as shown in Figure 6a–c. Three different S−O bonds are related to the electronic interactions between the S and O atoms as shown in Table 1. Furthermore, in Figure 6d–g, due to the larger electronegativity of S compared to Mo [45] and Ti [46], the O atom also gains more valence charge from the S, Mo, and Ti atoms, as supported by Figure 6d–g. Four different S−O bonds also assign to the electronic interactions among the S and O atoms, listed in Table 1. Moreover, in Figure 6d,g, there is no localized hybridization between the Ti and O atoms, which further reveals no chemical bonds between the Ti and O atoms.

3.4. Density of States (DOS)

To further probe the nature of the oxidation behavior of MoS2, the total and partial density of states of the O-adsorbed MoS2 and MSTi-Mo are calculated as shown in Figure 7 and Figure 8. The vertical dotted line represents the Fermi energy level (0 eV). For the O-adsorbed MoS2(002) surface, in Figure 7a (H site), valence bands consist of two parts; the first part (−13.42 to −11.24 eV) is composed of O-2s, S-3s, S-3p, Mo-5s, and Mo4d states, and the second part (−6.23 to 1.97 eV) mainly consists of O-2p, S-3p, and Mo-4d states. The hybridization of the O-2p and S-3p states in the valence bands is obviously observed, indicating the formation of covalence bonds between the O and S atoms. Figure 7b,c shows that the valence bands mainly consist of O-2p, S-3s, S-3p, and Mo-4d states. The hybridization of O-2p and S-3p are also detected as shown in Figure 7b,c. The covalence bonds between O and S are also formed for the Mo top and S top sites. The results are in agreement with that of the charge density difference.
Figure 8a–d shows the total density and partial density of the state of the O-adsorbed MSTi-Mo(002) surface. The valence bands for these four adsorption sites are mainly composed of O-2p, S-3s, S-3p, Mo-5s, and Mo-4d states. The obvious hybridization between the O-2p and S-3p states is found, which reveals the formation of the covalence bonds between the O and S atoms. However, the hybridization between the O-2p and Ti-3p is not observed, indicating that no chemical bonds are formed between the O and the Ti atoms.
According to what was previously mentioned, the Ti element in the MoS2 does not directly bond with the O atom. The covalence bonds are formed between the S and the O atoms. The O atom inclines to attack and damage Mo−S bonds in pristine MoS2; although the oxygen resistance of the Mo top site is improved, the O atom favors to attack and damage Ti−S bonds in the H and Ti top sites with the incorporation of Ti into MoS2. Thus, the antioxidation of MoS2 does not improve under the incorporation of the Ti atom. Because the sequence of electronegativity is Ti (1.54) < Mo (2.16) < S (2.58) < O (3.44), when O is adsorbed onto the MoS2 surface, O atoms tend to move toward the S atom as supported by the charge density difference analysis, which results in the charge transferring between the O atom and the MoS2 surface. Furthermore, the covalence bonds are easily formed between the O and S atoms leading to the broken Mo−S bonds.

4. Conclusions

In summary, the influence of the Ti atom on the antioxidation of MoS2 were investigated using the first-principles method. The main conclusions obtained are as follows:
(1)
The oxidation mechanism of the pristine MoS2 was well probed. Two Ti-doped models, MSTi-S and MSTi-Mo, were designed based on the structural feature of MoS2. Due to the dynamical instability of MSTi-S, the antioxidation behavior of MSTi-Mo was investigated in detail.
(2)
Although the antioxidation property of the Mo site improved, the H site and Ti top site were easily attacked by the O atom, leading to broken Ti–S bonds. The incorporation of the Ti element into the MoS2 was not helpful for improving its antioxidation performance.

Author Contributions

Conceptualization, L.G. and E.E.E.; methodology, S.L.; software, Q.Z. and R.M.; validation, S.L., Y.H. and Q.Z.; formal analysis, E.E.E.; investigation, S.L. and Y.H.; resources, L.G.; data curation, R.M.; writing—original draft preparation, S.L.; writing—review and editing, E.E.E.; visualization, R.M.; supervision, L.G.; project administration, L.G.; funding acquisition, S.L., R.M. and L.G. All authors have read and agreed to the published version of the manuscript.

Funding

Riadh Marzouki extends appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the General Research Project under grant number (RGP. 2/224/43). This work is supported by the Foundation of the Department of Science and Technology of the Guizhou province (No. CG [2021]110 and No. QKHPTRC [2021]5643) and the Foundation of the Department of Education of the Guizhou province (No. KY [2018]030 and No. QJJ [2022]003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, Y.H.; Zhang, X.Q.; Zhang, W.; Chang, M.T.; Lin, C.T.; Chang, K.D.; Yu, Y.C.; Wang, J.T.W.; Chang, C.S.; Li, L.J.; et al. Synthesis of large-area MoS2 atomic layers with chemical vapor deposition. Adv. Mater. 2012, 24, 2320–2325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Salimi, M.; Shokrgozar, M.A.; Hamid, D.H.; Vossoughi, M. Photothermal properties of two-dimensional molybdenum disulfide (MoS2) with nanoflower and nanosheet morphology. Mater. Res. Bull. 2022, 152, 111837. [Google Scholar] [CrossRef]
  3. Huang, W.; Li, Y.; Fu, Q.; Chen, M. Fabrication of a novel biochar decorated nano-flower-like MoS2 nanomaterial for the enhanced photodegradation activity of ciprofloxacin: Performance and mechanism. Mater. Res. Bull. 2022, 147, 111650. [Google Scholar] [CrossRef]
  4. Uddin, M.K.; Mashkoor, F.; Al Arifi, I.M.; Nasar, A. Simple one-step synthesis process of novel MoS2@bentonite magnetic nanocomposite for efficient adsorption of crystal violet from aqueous solution. Mater. Res. Bull. 2021, 139, 111279. [Google Scholar] [CrossRef]
  5. Tan, X.; Duan, Z.; Liu, H.; Wu, X.; Cho, Y.R. Core-shell structured MoS2/Ni9S8 electrocatalysts for high performance hydrogen and oxygen evolution reactions. Mater. Res. Bull. 2022, 146, 111626. [Google Scholar] [CrossRef]
  6. Krishnan, U.; Kaur, M.; Kaur, G.; Singh, K.; Dogra, A.R.; Kumar, M.; Kumar, A. MoS2/ZnO nanocomposites for efficient photocatalytic degradation of industrial pollutants. Mater. Res. Bull. 2019, 111, 212–221. [Google Scholar] [CrossRef]
  7. Imani Yengejeh, S.; Liu, J.; Kazemi, S.A.; Wen, W.; Wang, Y. Effect of structural phases on mechanical properties of molybdenum disulfide. ACS Omega 2020, 5, 5994–6002. [Google Scholar] [CrossRef]
  8. Li, H.; Wang, J.; Gao, S.; Chen, J.; Peng, L.; Liu, K.; Wei, X. Superlubricity between MoS2 monolayers. Adv. Mtaer. 2017, 29, 1701474. [Google Scholar] [CrossRef]
  9. Ren, S.; Li, H.; Cui, M.; Wang, L.; Pu, J. Functional regulation of Pb-Ti/MoS2 composite coatings for environmentally adaptive solid lubrication. Appl. Surf. Sci. 2017, 401, 362–372. [Google Scholar] [CrossRef]
  10. Li, Y.; Xie, M.; Sun, Q.; Xu, X.; Fan, X.; Zhang, G.; Li, H.; Zhu, M. The effect of atmosphere on the tribological behavior of magnetron sputtered MoS2 coatings. Surf. Coat. Technol. 2019, 378, 125081. [Google Scholar] [CrossRef]
  11. Le, D.; Rawal, T.B.; Rahman, T.S. Single-layer MoS2 with sulfur vacancies: Structure and catalytic application. J. Phys. Chem. C 2014, 118, 5346–5351. [Google Scholar] [CrossRef]
  12. Ren, S.M.; Cui, M.J.; Martini, A.; Shi, Y.B.; Wang, H.X.; Pu, J.B.; Li, Q.Y.; Wang, L.P. Macroscale superlubricity enabled by rationally designed MoS2-based superlattice films. Res. Sq. 2022. preprint. [Google Scholar] [CrossRef]
  13. Santosh, K.; Roberto, C.; Robert, M.; Kyeongjae, C. Surface oxidation energetics and kinetics on MoS2 monolayer. J. Appl. Phys. 2015, 117, 136301. [Google Scholar]
  14. Sen, H.S.; Sahin, H.; Peeters, F.M.; Durgun, E. Monolayers of MoS2 as an oxidation protective nanocoating material. J. Appl. Phys. 2014, 116, 083508. [Google Scholar] [CrossRef] [Green Version]
  15. Ataca, C.; Ciraci, S. Dissociation of H2O at the vacancies of single-layer MoS2. Phys. Rev. B 2012, 85, 195410. [Google Scholar] [CrossRef]
  16. Li, H.; Huang, M.; Cao, G. Markedly different adsorption behaviors of gas molecules on defective monolayer MoS2: A first-principles study. Phys. Chem. Chem. Phys. 2016, 18, 15110–15117. [Google Scholar] [CrossRef]
  17. Li, Q.; Zheng, S.; Pu, J.; Wang, W.; Li, L.; Wang, L. Revealing the failure mechanism and designing protection approach for MoS2 in humid environment by first-principles investigation. Appl. Surf. Sci. 2019, 487, 1121–1130. [Google Scholar] [CrossRef]
  18. Chen, Z.W.; Yan, J.M.; Zheng, W.T.; Jiang, Q. Cu4 cluster doped monolayer MoS2 for CO oxidation. Sci. Rep. 2015, 5, 11230. [Google Scholar] [CrossRef] [Green Version]
  19. Cong, W.T.; Tang, Z.; Zhao, X.G.; Chu, J.H. Enhanced magnetic anisotropies of single transition-metal adatoms on a defective MoS2 nonolayer. Sci. Rep. 2015, 5, 9361. [Google Scholar] [CrossRef] [Green Version]
  20. Zeng, C.; Pu, J.; Wang, H.; Zheng, S.; Wang, L.; Xue, Q. Study on atmospheric tribology performance of MoS2–W films with self-adaption to temperature. Ceram. Int. 2019, 45, 15834–15842. [Google Scholar] [CrossRef]
  21. Li, L.; Lu, Z.; Pu, J.; Wang, H.; Li, Q.; Chen, S.; Zhang, Z.; Wang, L. The superlattice structure and self-adaptive performance of C–Ti/MoS2 composite coatings. Ceram. Int. 2020, 46, 5733–5744. [Google Scholar] [CrossRef]
  22. Zeng, C.; Pu, J.; Wang, H.; Zheng, S.; Chen, R. Influence of microstructure on tribological properties and corrosion resistance of MoS2/WS2 films. Ceram. Int. 2020, 46, 13774–13783. [Google Scholar] [CrossRef]
  23. Li, H.; Xie, M.; Zhang, G.; Fan, X.; Li, X.; Zhu, M.; Wang, L. Structure and tribological behavior of Pb-Ti/MoS2 nanoscaled multilayer films deposited by magnetron sputtering method. Appl. Surf. Sci. 2018, 435, 48–54. [Google Scholar] [CrossRef]
  24. Wang, P.; Yue, W.; Lu, Z.; Zhang, G.; Zhu, L. Friction and wear properties of MoS2-based coatings sliding against Cu and Al under electric current. Tribol. Int. 2018, 127, 379–388. [Google Scholar] [CrossRef]
  25. Lu, X.; Yan, M.; Yan, Z.; Chen, W.; Sui, X.; Hao, J.; Liu, W. Exploring the atmospheric tribological properties of MoS2-(Cr, Nb, Ti, Al, V) composite coatings by high throughput preparation method. Tribol. Int. 2021, 156, 106844. [Google Scholar] [CrossRef]
  26. Pan, Y. Influence of N-vacancy on the electronic and optical properties of bulk GaN from first-principles investigations. Int. J. Energy Res. 2021, 45, 15512–15520. [Google Scholar] [CrossRef]
  27. Chen, S.; Pan, Y. Influence of group III and IV elements on the hydrogen evolution reaction of MoS2 disulfide. J. Phys. Chem. C 2021, 125, 11848–11856. [Google Scholar] [CrossRef]
  28. Zhang, R.; Lu, Z.; Yang, Y.; Shi, W. First-principles investigation of the monoclinic NaMnO2 cathode material for rechargeable Na-ion batteries. Curr. Appl. Phys. 2018, 18, 1431–1435. [Google Scholar] [CrossRef]
  29. Zhang, R.; Zhao, J.; Yang, Y.; Shi, W.; Lu, Z.; Wang, J. Understanding the friction behavior of sulfur-terminated diamond-like carbon films under high vacuum by first-principles calculations. Curr. Appl. Phys. 2018, 18, 317–323. [Google Scholar] [CrossRef]
  30. Zhang, R.; Zhao, J.; Pu, J.; Lu, Z. First-Principles Investigation on the Tribological Properties of h-BN Bilayer Under Variable Load. Tribol. Lett. 2018, 66, 124. [Google Scholar] [CrossRef]
  31. Pan, Y.; Wang, S.; Mao, P.; Jin, C. Role of Si concentration on the thermodynamic properties of molybdenum silicides. Vacuum 2017, 141, 170–172. [Google Scholar] [CrossRef]
  32. Pan, Y. The structural, mechanical and thermodynamic properties of the orthorhombic TMAl (TM=Ti, Y, Zr and Hf) aluminides from first-principles calculations. Vacuum 2020, 181, 109742. [Google Scholar] [CrossRef]
  33. Pan, Y.; Yu, E. First-principles investigation of electronic and optical properties of H-doped FeS2. Int. J. Energy Res. 2021, 45, 11284–11293. [Google Scholar] [CrossRef]
  34. Pan, Y. First-principles investigation of the structural, mechanical, and thermodynamic properties of hexagonal and cubic MoAl5 alloy. J. Mater. Eng. Perform. 2021, 30, 8289–8295. [Google Scholar] [CrossRef]
  35. Tornatzky, H.; Gillen, R.; Uchiyama, H.; Maultzsch, J. Phonon dispersion in MoS2. Phys. Rev. B 2019, 99, 144309. [Google Scholar] [CrossRef] [Green Version]
  36. Pu, D.; Pan, Y. First-principles investigation of oxidation mechanism of Al-doped Mo5Si3 silicide. Ceram. Int. 2022, 48, 11518–11526. [Google Scholar] [CrossRef]
  37. Williamson, I.; Li, S.; Correa Hernandez, A.; Lawson, M.; Chen, Y.; Li, L. Structural, electrical, phonon, and optical properties of Ti- and V-doped two-dimensional MoS2. Chem. Phys. Lett. 2017, 674, 157–163. [Google Scholar] [CrossRef]
  38. Pan, Y.; Wen, M. Insight into the oxidation mechanism of Nb3Si (111) surface: First-principles calculations. Mater. Res. Bull. 2018, 107, 484–491. [Google Scholar] [CrossRef]
  39. Das, N.K.; Shoji, T. Early-stage oxidation of Ni–Cr binary alloy (111), (110) and (100) surfaces: A combined density functional and quantum chemical molecular dynamics study. Corros. Sci. 2013, 73, 18–31. [Google Scholar] [CrossRef]
  40. Elgendy, A.; Elkholy, A.E.; El Basiony, N.M.; Migahed, M.A. Monte Carlo simulation for the antiscaling performance of Gemini ionic liquids. J. Mol. Liq. 2019, 285, 408–415. [Google Scholar] [CrossRef]
  41. Guo, L.; Tan, J.; Kaya, S.; Leng, S.; Li, Q.; Zhang, F. Multidimensional insights into the corrosion inhibition of 3,3-dithiodipropionic acid on Q235 steel in H2SO4 medium: A combined experimental and in silico investigation. J. Colloid Interface Sci. 2020, 570, 116–124. [Google Scholar] [CrossRef]
  42. Islam, Z.; Haque, A. Defects and grain boundary effects in MoS2: A molecular dynamics study. J. Phys. Chem. Solids 2021, 148, 109669. [Google Scholar] [CrossRef]
  43. Zhang, R.; Wang, Q.; Liang, J.; Li, Q.; Dai, J.; Li, W. Optical properties of N and transition metal R (R=V, Cr, Mn, Fe, Co, Ni, Cu, and Zn) codoped anatase TiO2. Phys. B Condens. Matter 2012, 407, 2709–2715. [Google Scholar] [CrossRef]
  44. Michalczyk, M.; Zierkiewicz, W.; Scheiner, S. Interactions of (MY)6 (M=Zn, Cd; Y=O, S, Se) quantum dots with N-bases. Struct. Chem. 2019, 30, 1003–1014. [Google Scholar] [CrossRef] [Green Version]
  45. Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 2009, 21, 084204. [Google Scholar] [CrossRef]
  46. Liu, W.; Zhang, Z.; Peng, X.; Zhong, J. Tuning the Dirac cone of the topological insulator Bi2Te3 thin films by substitutional nonmagnetic atoms. Phys. B Condens. Matter 2015, 456, 355–358. [Google Scholar] [CrossRef]
Figure 1. The calculated model of hexagonal MoS2.
Figure 1. The calculated model of hexagonal MoS2.
Coatings 12 01114 g001
Figure 2. Phonon dispersion curves of (a) MoS2, (b) Ti-doped MoS2 with a replaced S atom, and (c) Ti-doped MoS2 with a replaced Mo atom.
Figure 2. Phonon dispersion curves of (a) MoS2, (b) Ti-doped MoS2 with a replaced S atom, and (c) Ti-doped MoS2 with a replaced Mo atom.
Coatings 12 01114 g002
Figure 3. The optimized oxygen-adsorbed site on the MoS2(002) 2 × 2 × 1 supercell; note: the vacuum layer 20 Å is added along the perpendicular direction to avoid the interaction between their period images, and the dopant (Ti atom) replaces the S or Mo atom.
Figure 3. The optimized oxygen-adsorbed site on the MoS2(002) 2 × 2 × 1 supercell; note: the vacuum layer 20 Å is added along the perpendicular direction to avoid the interaction between their period images, and the dopant (Ti atom) replaces the S or Mo atom.
Coatings 12 01114 g003
Figure 4. The relaxed configurations for the (ac) top view and (df) side view of the O-adsorbed MoS2(002) surfaces.
Figure 4. The relaxed configurations for the (ac) top view and (df) side view of the O-adsorbed MoS2(002) surfaces.
Coatings 12 01114 g004
Figure 5. The relaxed configurations for the (ad) top view and (eh) side view of the O-adsorbed MSTi-Mo (002) surfaces.
Figure 5. The relaxed configurations for the (ad) top view and (eh) side view of the O-adsorbed MSTi-Mo (002) surfaces.
Coatings 12 01114 g005
Figure 6. The calculated charge density difference of the (ac) O-adsorbed MoS2(002) surface and the (dg) O-adsorbed MSTi-Mo(002) surface. Note: the blue represents the maximum electronic delocalization; the yellow represents the maximum electronic localization.
Figure 6. The calculated charge density difference of the (ac) O-adsorbed MoS2(002) surface and the (dg) O-adsorbed MSTi-Mo(002) surface. Note: the blue represents the maximum electronic delocalization; the yellow represents the maximum electronic localization.
Coatings 12 01114 g006
Figure 7. The total and partial density of state (DOS) of the O-adsorbed MoS2(002) surface: (a) H site, (b) Mo top site, and (c) S top site.
Figure 7. The total and partial density of state (DOS) of the O-adsorbed MoS2(002) surface: (a) H site, (b) Mo top site, and (c) S top site.
Coatings 12 01114 g007
Figure 8. The total and partial density of state (DOS) of the O-adsorbed MSTi-Mo(002) surface: (a) H site, (b) Mo top site, (c) S top site, and (d) Ti top site.
Figure 8. The total and partial density of state (DOS) of the O-adsorbed MSTi-Mo(002) surface: (a) H site, (b) Mo top site, (c) S top site, and (d) Ti top site.
Coatings 12 01114 g008
Table 1. The calculated E a d O values for O adsorbed on the MoS2(002) and MSTi-Mo (002) surface.
Table 1. The calculated E a d O values for O adsorbed on the MoS2(002) and MSTi-Mo (002) surface.
ConfigurationsAdsorbed Site E a d O (eV) dS−O (Å)dMo−O (Å)dTi−O (Å)dTi−Mo (Å)
MoS2H−5.531.484---
Mo top−3.961.5682.204--
S top−5.521.483---
Ti−MoS2H−1.331.6392.3362.010-
Mo top−0.701.578---
S top−1.901.487---
Ti top−1.041.601-2.062-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Leng, S.; Zhang, Q.; Guo, L.; Huang, Y.; Ebenso, E.E.; Marzouki, R. First-Principles Calculations to Investigate the Oxidation Mechanism of Pristine MoS2 and Ti-Doped MoS2. Coatings 2022, 12, 1114. https://doi.org/10.3390/coatings12081114

AMA Style

Leng S, Zhang Q, Guo L, Huang Y, Ebenso EE, Marzouki R. First-Principles Calculations to Investigate the Oxidation Mechanism of Pristine MoS2 and Ti-Doped MoS2. Coatings. 2022; 12(8):1114. https://doi.org/10.3390/coatings12081114

Chicago/Turabian Style

Leng, Senlin, Qiao Zhang, Lei Guo, Yue Huang, Eno E. Ebenso, and Riadh Marzouki. 2022. "First-Principles Calculations to Investigate the Oxidation Mechanism of Pristine MoS2 and Ti-Doped MoS2" Coatings 12, no. 8: 1114. https://doi.org/10.3390/coatings12081114

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