Next Article in Journal
The Effects of New Zealand Grown Ginseng Fractions on Cytokine Production from Human Monocytic THP-1 Cells
Previous Article in Journal
Influence of Genistein on Hepatic Lipid Metabolism in an In Vitro Model of Hepatic Steatosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Properties of Triangle Molybdenum Disulfide (MoS2) Clusters with Different Sizes and Edges

1
School of Physics, Harbin Institute of Technology, Harbin 150001, China
2
Key Lab of Micro-Optics and Photonic Technology of Heilongjiang Province, Harbin 150001, China
3
School of Mechanical & Power Engineering, Harbin University of Science and Technology, Harbin 150080, China
4
Key Laboratory of Green Chemical Technology of College of Heilongjiang Province, College of Chemical and Environmental Engineering, Harbin University of Science and Technology, Harbin 150080, China
5
State Key Laboratory of Intense Pulsed Radiation Simulation and Effect, Xi’an 710024, China
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(4), 1157; https://doi.org/10.3390/molecules26041157
Submission received: 14 January 2021 / Revised: 29 January 2021 / Accepted: 2 February 2021 / Published: 22 February 2021

Abstract

:
The electronic structures and transition properties of three types of triangle MoS2 clusters, A (Mo edge passivated with two S atoms), B (Mo edge passivated with one S atom), and C (S edge) have been explored using quantum chemistry methods. The highest occupied molecular orbital (HOMO)–lowest unoccupied molecular orbital (LUMO) gap of B and C is larger than that of A, due to the absence of the dangling of edge S atoms. The frontier orbitals (FMOs) of A can be divided into two categories, edge states from S3p at the edge and hybrid states of Mo4d and S3p covering the whole cluster. Due to edge/corner states appearing in the FMOs of triangle MoS2 clusters, their absorption spectra show unique characteristics along with the edge structure and size.

1. Introduction

Since the successful isolation of graphene [1,2], atomic layer-thick two-dimensional (2D) nanomaterials have attracted a great deal of attention due to their excellent mechanical flexibility and optical transparency, as well as their ultrahigh specific surface, which are related to their strong in-plane covalent bond and atomic thickness [3,4,5,6,7]. Beyond graphene, single-layer transition-metal dichalcogenides (TMDs) have been receiving increasing interest. For example, the structural and physical properties of the MoS2 monolayer and bulk have been extensively studied both theoretically and experimentally. The monolayer of MoS2 consists of a layer of Mo atoms sandwiched between two layers of S atoms, forming a trilayer. The absence of inversion symmetry and dominant d-electron interactions of the heavy transition metal atom Mo, endow the monolayer (MoS2) with intriguing physical properties that are not found in sp-bonded nanomaterials [8,9,10,11].
In recent years, an increasing amount of effort has been devoted to 2D materials with ultrasmall sizes, due to increasing demand for miniaturizing photonic and optoelectronic devices. For MoS2, from the three-dimensional (3D) bulk structure to the 2D monolayer, the band gap transforms from an indirect band gap (∼1.2 eV) to a direct band gap (∼1.8 eV, in the visible frequency range) [12,13]. Compared with their 2D and one dimensional (1D) counterparts, zero dimensional (0D) MoS2 clusters possess tunable energy levels, more active edges, and larger surface-area-to-volume ratios, which give them fascinating properties. This makes them promising candidates for application in fields such as optoelectronics [14,15], electrochemical technology [16,17], biology [18], catalysis [19] and so forth [20,21,22]. Jaramillo et al. [23] pointed out that 2D TMDs are hydrogen evolution reaction (HER) active because of their highly active edges, which makes them potential candidates for electrocatalytic hydrogen evolution. Yin et al. [24] pointed out that a strong second-harmonic generation (SHG) can be observed at the edges and corners of MoS2 clusters, implying the important role that edges and corner states play in determining the properties of the 0D structures. However, the detailed transition natures are ambiguous. In order to shed light on the essence of the various applications and obtain full use of the MoS2 clusters, it is crucial to gain a comprehensive understanding of the electronic structures and their transition properties.
It is well known that when the physical size of the material is comparable to or smaller than the Bohr radius, the excitons are confined in all three dimensions, so there exists a 3D quantum confinement effect. Superimposed on the intrinsic size-dependent electronic structure, one thus has the effect of edges, corners, atomic vacancies and so forth, all these are likely to induce additional local effects. As a result, the electronic properties become increasingly sensitive not only to sizes, but to the shapes, edge atomic structures, and compositions [25,26,27]. Despite the breakthroughs in preparation of MoS2 clusters over the years, mainly including top-down [28,29,30,31,32,33] and bottom-up methods [34,35,36], a cost-effective yet efficient technique for the high yield production of MoS2 clusters with desirable sizes remains a challenge. Consequently, it is still difficult to provide a specific illustration about the properties of MoS2 clusters experimentally. Aiming to obtain the whole picture in relation to MoS2 clusters and to clarify the relationships between structures and properties, the density-of-states (DOS), frontier molecular orbitals (FMOs), electronic absorption spectra, and electron-hole characteristics of a series of triangular MoS2 clusters with different sizes and edges will be explored using quantum chemistry methods in the present paper. This will be of great significance not only in fundamental physics exploration but also in device applications.

2. Models and Computational Details

Based on different edge structures, there are mainly three types of triangle MoS2 clusters, as shown in Figure 1. A and B represent the Mo-terminated (commonly referred as Mo-edge) triangle MoS2 clusters, while C is the S-terminated cluster (commonly referred as S-edge). Considering that a bare Mo-edge is not stable, it is often passivated by one or two S atoms per edge-Mo atom. In A, each edge-Mo atom is passivated with two S atoms forming an S-dimer normal to the basal plane, usually called a Mo edge, with 100% S coverage.
In the case of B, two adjacent edge-Mo atoms share one S atom forming an S-monomer parallel to the basal plane, usually called a Mo edge, with 50% S coverage. Consequently, the S atoms of C dimerize along the direction normal to the basal plane, often called a S edge, with 100% coverage. By changing the number of the edge-Mo atoms n (n = 3, 4, 5, and 6) in A, B, and C, a series of different triangle MoS2 clusters with different sizes and edges could be gained. In the following discussions, A refers to four triangle MoS2 clusters with the same edges but different sizes, including 3A, 4A, 5A, and 6A, as B and C.
The density functional theory (DFT) based on Becke’s three-parameter hybrid exchange combined with the Lee–Yang–Parr correlation (B3LYP) [37,38] and the LANL2DZ basis set were adopted to optimize the geometries of the triangular MoS2 clusters. With both the merits of B3LYP and long-range corrected properties, CAM-B3LYP is able to predict molecular charge-transfer spectra properly due to the improved description of the long-range exchange interactions [39,40,41,42]. It is employed to evaluate the electron excitations with the LANL2DZ basis set in the present work.
The DFT calculations were performed using the Gaussian 09 program [43]. The data of DOS, electronic absorption spectra, and electron-hole distributions were obtained with the Multiwfn package [44].

3. Results and Discussion

In order to indicate the stability of A, B, and C, the formation energy was calculated according to the formula: Eformation = Ecluster (MomSn) − mE (Mo) − nE (S), where Ecluster (MomSn), E (Mo), and E (S) denote the energy of the cluster having m Mo atoms and n S atoms, the energy of a single Mo atom, and the energy of a single S atom, respectively. The calculated Eformation values of A, B, and C are listed in Table 1, Table 2 and Table 3, respectively. The negative Eformation values imply that the clusters are stable. The Eformation decrease with increasing the sizes of the cluster, suggesting that the larger the cluster, the more energy is released.
For the three different classes of clusters, their highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energies and their energy gaps, as well as the orbital compositions of HOMO and LUMO are listed in Table 1, Table 2 and Table 3, respectively. The DOS is shown in Figure 2.
For these ultrasmall MoS2 clusters, the band gap does not decrease monotonically with size as is the case for graphite flakes, which might be attributed the deformation of structures due to large the ratio of number of edge atoms to core atoms in ultrasmall clusters. The energy of HOMO (EH) in 3A is −5.06 eV, while that of LUMO (EL) is −6.84 eV, resulting in a 1.78 eV energy gap (Egap). The Egap value of 4A, 5A, and 6A is 0.70, 0.66, and 0.73 eV, respectively, far smaller than that of 3A. For the Mo-terminated triangle MoS2 clusters with 100% S coverage, some FMOs contributed from S3p at the edge exhibit evident edge states, and the other FMOs exhibit hybrid states of Mo4d and S3p whose charge distribution covers the whole cluster. By comparing the electronic structures of three classes of clusters, it can be seen that the energy gap of the ultrasmall MoS2 clusters is dependent on the edge structure, which is similar to graphite flake. For Mo-terminated triangle MoS2 clusters with 50% S coverage (B) and S-terminated triangle MoS2 clusters (C), their S atoms at the edge of the cluster are shared by two adjacent Mo atoms. In contrast to the A system, the absence of dangling edge S atoms in the two class clusters increases their HOMO–LUMO gap and evident edge/corner states can be observed in the FMOs of these clusters. For B, all the FMOs comes from the hybridization between Mo4d and S3p. While the occupied FMOs mainly comes from the S3p, and the unoccupied FMOs from the hybridization between Mo4d and S3p in the case of C.
The electronic absorption spectra of A, B, and C and their corresponding electron-hole distributions where the process of single-electron excitation is described as “an electron goes to the electron from the hole” are given in Figure 3, Figure 4, Figure 5 and Figure 6, respectively. Moreover, the percentage of Mo4d and S3p in the electron and hole are provided in Table 4, Table 5 and Table 6, respectively. The maximum absorption shows a red shift along with increasing size of the clusters in the A, B and C systems. For 3A, 4A, 5A, and 6A, the maximum absorption is ~623.61, ~591.50, ~670.54, and 796.87 nm, respectively. From Table 4, the percentage of S3p is larger than that of Mo4d in both the electron and hole in 3A, Mo4d takes the larger proportion in the electron and the hole in 4A and 5A. The S3p and Mo4d make almost the same contribution to the electron and the hole in 6A. Combined with the data in Figure 4, there are two types of electron transitions: one is a d-d transition of the Mo atoms, and the other occurs at the S atoms along the edges, which is associated with the edge states.
The maximum absorption is ~422.32, ~609.91, ~656.23, and 678.05 nm for 3B, 4B, 5B, and 6B, respectively. The electrons and holes distribute along the edges/corners. The percentage of Mo4d in the electron is larger than that of S3p, while the percentage of S3p in the hole is larger than that of Mo4d, leaving S atoms with the feature of hole and Mo atoms with the feature of the electron. In other words, the electron prefers to transfer from the S3p to the connected Mo4d along the edges/corners of B. The maximum absorption is ~502.21, ~654.45, ~633.81, and 845.93 nm for 3C, 4C, 5C, and 6C, respectively. The electron and hole distribute at the corners with the S atoms at the top vertex acting as hole and the Mo atoms neighboring acting as the electron (see Table 6 and Figure 6).
Overall, the S0→Sn transitions corresponding to the maximum absorption of the Mo-terminated triangle MoS2 clusters with 100% S coverage A come from two aspects, the d-d transitions of the Mo atom and transitions of S3p along the edge. The S0→Sn transitions of B and C occur at the edges/corners, with an electron transferring from the S atoms to the neighboring Mo atoms. In other words, the S atoms act as the hole and the adjacent Mo atoms act as the electron in the case of the clusters with no dangling of edge S atoms in them. The S0→S1 transition directly leads to the lowest excited state and is always the excited state for luminescence. For the S0→S1 transition of A, the electron and hole mainly distribute throughout the whole triangular clusters; this can be observed faintly at the S atoms along the edges (see Figure 4). Referred to the data in Table 4, the percentage of Mo4d in electron and hole not only takes a larger proportion but also has nearly the same values, which implies obvious d-d transitions of the Mo atoms. The S0→S1 transitions of B also occur at the edges/corners, as shown in Figure 5. By referring to the data in Table 5, both the Mo and S atoms possess the distribution of electrons and holes, which ascribes the S0→S1 from local transition. For the S0→S1 transitions of C, the percentage of Mo4d in the electron is larger than that in the hole, while the percentage of S3p in the hole is larger than that in the electron, implying significant electron transfer from the S atoms to the neighboring Mo atoms along the edges (see Figure 6).

4. Conclusions

The electronic properties of three types of triangle MoS2 clusters, A (Mo-edge passivated with two S atoms), B (Mo-edge passivated with one S atom), and C (S-edge) have been explored using quantum chemistry methods in the present paper. The DOS appears to have different features in A, B, and C. For A, the MOs can be divided into two types: the edge state from S3p along the edges, the hybrid state from Mo4d and S3p covering the whole cluster. Evident edge/corner states appear in the FMOs of B and C. The hybridization of Mo4d and S3p in B generate the occupied and unoccupied MOs. The occupied MOs mainly come from S3p, and the unoccupied MOs from the hybridization of Mo4d and S3p in C.
For the electron excitation processes of A, B, and C, the absorption peaks show a red shift along with increasing the size of the clusters. Electron–hole analysis indicated that the S0→S1 transitions appear to have different characteristics compared with the S0→Sn transitions (associated with the intense absorption in the electronic absorption spectrum) in A, B, and C. In A, the S0→S1 transitions mainly stem from the d-d transitions of the Mo atoms, while the S0→Sn transitions come from the d-d transitions of the Mo atom or transitions of the S3p along the edges. Both the S0→S1 and S0→Sn transitions of B occur at the edges/corners, with the local S0→S1 transitions originating from Mo4d and S3p, but charge transfer S0→Sn transitions caused by the electron transferring from S3p to the connected Mo4d. In C, the S0→S1 transitions mainly occur along the edge with the S atoms acting as hole and the adjacent Mo atoms as electron, while the S0→Sn transitions mainly occur at the corners, where the two S atoms at the vertex acts as the hole and the adjacent Mo atoms act as the electron.

Author Contributions

Conceptualization, W.L. and Y.J.; validation, C.H.; formal analysis, W.L. and G.Z.; data curation, C.H. and L.Y.; writing—original draft preparation, S.W. and Y.H.; writing—review and editing, S.W. and Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Heilongjiang Province of China, grant number B2018007.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  3. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-Like Two-Dimensional Materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef]
  4. Tan, C.; Cao, X.; Wu, X.-J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.-H.; et al. Recent Advances in Ultrathin Two-Dimensional Nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef]
  5. Zhang, X.; Lai, Z.; Ma, Q.; Zhang, H. Novel structured transition metal dichalcogenide nanosheets. Chem. Soc. Rev. 2018, 47, 3301–3338. [Google Scholar] [CrossRef]
  6. Jariwala, D.; Sangwan, V.K.; Lauhon, L.J.; Marks, T.J.; Hersam, M.C. Emerging Device Applications for Semiconducting Two-Dimensional Transition Metal Dichalcogenides. ACS Nano 2014, 8, 1102–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Rao, C.N.R.; Gopalakrishnan, K.; Maitra, U. Comparative Study of Potential Applications of Graphene, MoS2, and Other Two-Dimensional Materials in Energy Devices, Sensors, and Related Areas. ACS Appl. Mater. Interfaces 2015, 7, 7809–7832. [Google Scholar] [CrossRef]
  8. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  10. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  11. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelec-tronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Lauritsen, J.V.; Kibsgaard, J.; Helveg, S.; Topsøe, H.; Clausen, B.S.; Laegsgaard, E.; Besenbacher, F. Size-dependent structure of MoS2 nanocrystals. Nat. Nanotechnol. 2007, 2, 53–58. [Google Scholar] [CrossRef]
  14. Yin, W.; Liu, X.; Zhang, X.; Gao, X.; Colvin, V.L.; Zhang, Y.; Yu, W.W. Synthesis of Tungsten Disulfide and Molybdenum Disulfide Quantum Dots and Their Applications. Chem. Mater. 2020, 32, 4409–4424. [Google Scholar] [CrossRef]
  15. Lu, G.-Z.; Wu, M.-J.; Lin, T.-N.; Chang, C.-Y.; Lin, W.-L.; Chen, Y.T.; Hou, C.-F.; Cheng, H.-J.; Lin, T.-Y.; Shen, J.-L.; et al. Electrically Pumped White-Light-Emitting Diodes Based on Histidine-Doped MoS2 Quantum Dots. Small 2019, 15, 1901908. [Google Scholar] [CrossRef]
  16. Liu, A.; Lv, H.; Liu, H.; Li, Q.; Zhao, H. Two dimensional MoS2/CNT hybrid ink for paper-based capacitive energy storage. J. Mater. Sci. Mater. Electron. 2017, 28, 8452–8459. [Google Scholar] [CrossRef]
  17. Hota, P.; Miah, M.; Bose, S.; Dinda, D.; Ghorai, U.K.; Su, Y.-K.; Saha, S.K. Ultra-small amorphous MoS2 decorated reduced graphene oxide for supercapacitor application. J. Mater. Sci. Technol. 2020, 40, 196–203. [Google Scholar] [CrossRef]
  18. Guo, Y.; Li, J. MoS2 quantum dots: Synthesis, properties and biological applications. Mater. Sci. Eng. C 2020, 109, 110511. [Google Scholar] [CrossRef] [PubMed]
  19. Tian, X.; Sun, Y.; Fan, S.; Boudreau, M.D.; Chen, C.; Ge, C.; Yin, J.-J. Photogenerated Charge Carriers in Molybdenum Disulfide Quantum Dots with Enhanced Antibacterial Activity. ACS Appl. Mater. Interfaces 2019, 11, 4858–4866. [Google Scholar] [CrossRef]
  20. Guo, X.; Huang, J.; Zeng, Q.; Wei, Y.; Liu, X.; Wang, L. Boronic acid functionalized molybdenum disul-fide quantum dots for ultrasensitive analysis of dopamine basing on synergistic quenched effects from IFE and aggregation. J. Mater. Chem. B 2019, 7, 2799–2807. [Google Scholar] [CrossRef] [PubMed]
  21. Yadav, S.; Chaudhuary, P.; Uttam, K.N.; Varma, A.; Vashistha, M.; Yadav, B.C. Facile Synthesis of Molyb-denum disulfide (MoS2) Quantum Dots and its Application in Humidity Sensing. Nanotechnology 2019, 30, 295501. [Google Scholar] [CrossRef]
  22. Dong, H.; Tang, S.; Hao, Y.; Yu, H.; Dai, W.; Zhao, G.; Cao, Y.; Lu, H.; Zhang, X.; Ju, H. Fluorescent MoS2 Quantum Dots: Ultrasonic Preparation, Up-Conversion and Down-Conversion Bioimaging, and Photodynamic Therapy. ACS Appl. Mater. Interfaces 2016, 8, 3107–3114. [Google Scholar] [CrossRef] [PubMed]
  23. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [Green Version]
  24. Yin, X.; Ye, Z.; Chenet, D.A.; Ye, Y.; O’Brien, K.; Hone, J.C.; Zhang, X. Edge Nonlinear Optics on a MoS2 Atomic Monolayer. Science 2014, 344, 488–490. [Google Scholar] [CrossRef] [PubMed]
  25. El-Sayed, M.A. Small Is Different: Shape-, Size-, and Composition-Dependent Properties of Some Colloidal Semi-conductor Nanocrystals. Acc. Chem. Res. 2004, 37, 326–333. [Google Scholar] [CrossRef] [PubMed]
  26. Yoffe, A.D. Low-dimensional systems: Quantum size effects and electronic properties of semiconductor microcrystallites (zero-dimensional systems) and some quasi-two-dimensional systems. Adv. Phys. 2002, 51, 799–890. [Google Scholar] [CrossRef]
  27. Lv, R.; Terrones, H.; Elías, A.L.; Perea-López, N.; Gutiérrez, H.R.; Cruz-Silva, E.; Rajukumar, L.P.; Dresselhaus, M.S.; Mauricio, T. Two-dimensional transition metal dichalcogenides: Clus-ters, ribbons, sheets and more. Nano Today 2015, 10, 559–592. [Google Scholar] [CrossRef] [Green Version]
  28. Štengl, V.; Henych, J. Strongly luminescent monolayered MoS2 prepared by effective ultrasound exfoliation. Nanoscale 2013, 5, 3387–3394. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, S.; Li, D.; Wu, P. One-Pot, Facile, and Versatile Synthesis of Monolayer MoS2/WS2 Quantum Dots as Bioimaging Probes and Efficient Electrocatalysts for Hydrogen Evolution Reaction. Adv. Funct. Mater. 2015, 25, 1127–1136. [Google Scholar] [CrossRef]
  30. Zeng, Z.; Yin, Z.; Huang, X.; Li, H.; He, Q.; Lu, G.; Boey, F.; Zhang, H. Single-Layer Semicon-ducting Nanosheets: High-Yield Preparation and Device Fabrication. Angew. Chem. Int. Ed. 2011, 50, 11093–11097. [Google Scholar] [CrossRef] [PubMed]
  31. Shaijumon, M.; Damien, D.; Li, B.; Gullappalli, H.; Pillai, V.K.; Ajayan, P.M.; Shaijumon, M.M. Electrochemical synthesis of luminescent MoS2 quantum dots. Chem. Commun. 2015, 51, 6293–6296. [Google Scholar] [CrossRef]
  32. Gu, W.; Yan, Y.; Cao, X.; Zhang, C.; Ding, C.; Xian, Y. A facile and one-step ethanol-thermal syn-thesis of MoS2 quantum dots for two-photon fluorescence imaging. J. Mater. Chem. B 2016, 4, 27–31. [Google Scholar] [CrossRef] [PubMed]
  33. Ali, J.; Siddiqui, G.U.; Choi, K.H.; Jang, Y.; Lee, K. Fabrication of blue luminescent MoS 2 quantum dots by wet grinding assisted co-solvent sonication. J. Lumin. 2016, 169, 342–347. [Google Scholar] [CrossRef]
  34. Ren, X.; Pang, L.; Zhang, Y.; Ren, X.; Fan, H.; Liu, S. (Frank) One-step hydrothermal synthesis of monolayer MoS2 quantum dots for highly efficient electrocatalytic hydrogen evolution. J. Mater. Chem. A 2015, 3, 10693–10697. [Google Scholar] [CrossRef]
  35. Zhu, H.; Zhang, H.; Xia, Y. Planar Is Better: Monodisperse Three-Layered MoS2 Quantum Dots as Fluorescent Re-porters for 2,4,6-Trinitrotoluene Sensing in Environmental Water and Luggage Cases. Anal. Chem. 2018, 90, 3942–3949. [Google Scholar] [CrossRef]
  36. Han, G.G.D.; Tu, K.-H.; Niroui, F.; Xu, W.; Zhou, S.; Wang, X.; Bulović, V.; Ross, C.A.; Warner, J.H.; Grossman, J.C. Photoluminescent Arrays of Nanopatterned Monolayer MoS2. Adv. Funct. Mater. 2017, 27, 1703688. [Google Scholar] [CrossRef]
  37. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  38. Lee, C.; Yang, W.; Parr, R.G. Development of the Colic-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange-correlation functional using the Cou-lomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  40. Tawada, Y.; Tsuneda, T.; Yanagisawa, S.; Yanai, T.; Hirao, K. A long-range-corrected time-dependent density functional theory. J. Chem. Phys. 2004, 120, 8425–8433. [Google Scholar] [CrossRef] [PubMed]
  41. Hu, Y.-Y.; Gu, J.; Li, W.; Yang, L.; Feng, J.-K.; Tian, W.Q. From Graphene to Carbon Nanotubes: Variation of the Electronic States and Nonlinear Optical Responses. ChemPhysChem 2015, 16, 2151–2158. [Google Scholar] [CrossRef] [PubMed]
  42. Sheng, X.; Zhu, H.; Yin, K.; Chen, J.; Wang, J.; Wang, C.; Shao, J.; Chen, F. Excited-State Absorption by Linear Response Time-Dependent Density Functional Theory. J. Phys. Chem. C 2020, 124, 4693–4700. [Google Scholar] [CrossRef]
  43. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Fox, Gaussian 09, Revision A02; Gaussian, Inc.: Walingford, CT, USA, 2009. [Google Scholar]
  44. Lu, T.; Chen, F.J. Multiwfn: A Multifunctional Wavefunction Analyzer. Comp. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Top view of the triangle MoS2 clusters with different sizes and edges. The Mo and S atoms are represented by green and yellow spheres, respectively.
Figure 1. Top view of the triangle MoS2 clusters with different sizes and edges. The Mo and S atoms are represented by green and yellow spheres, respectively.
Molecules 26 01157 g001
Figure 2. The density-of-states (DOS) map, as well as HOMO and LUMO distributions of A, B, and C. The discrete represents the MOs, the black, blue, and red curves represent the total DOS, partial DOS of S3p, and partial DOS of Mo4d, respectively.
Figure 2. The density-of-states (DOS) map, as well as HOMO and LUMO distributions of A, B, and C. The discrete represents the MOs, the black, blue, and red curves represent the total DOS, partial DOS of S3p, and partial DOS of Mo4d, respectively.
Molecules 26 01157 g002
Figure 3. Electronic absorption spectra with full width at half maximum (FWHM) 0.26 of A, B, and C.
Figure 3. Electronic absorption spectra with full width at half maximum (FWHM) 0.26 of A, B, and C.
Molecules 26 01157 g003
Figure 4. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3A, 4A, 5A and 6A.
Figure 4. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3A, 4A, 5A and 6A.
Molecules 26 01157 g004
Figure 5. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3B, 4B, 5B and 6B.
Figure 5. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3B, 4B, 5B and 6B.
Molecules 26 01157 g005
Figure 6. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3C, 4C, 5C and 6C.
Figure 6. Electron (green) and hole (blue) distributions with isosurface 0.0006 for 3C, 4C, 5C and 6C.
Molecules 26 01157 g006
Table 1. Formation energy Eformation (eV), energy of the highest occupied molecular orbital (HOMO) EH (eV), energy of the lowest unoccupied molecular orbital (LUMO) EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3A, 4A, 5A, and 6A.
Table 1. Formation energy Eformation (eV), energy of the highest occupied molecular orbital (HOMO) EH (eV), energy of the lowest unoccupied molecular orbital (LUMO) EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3A, 4A, 5A, and 6A.
3A4A5A6A
Eformation−151.84−240.28−348.85−476.47
EH−5.06−5.55−5.29−5.71
EL−6.84−6.25−5.95−6.45
Egap1.780.700.660.73
CHS: px, pyS: px, pyMo: dxy, dx2y2, dz2
S: px, py
S: px, py
CLMo: dxy, dx2y2
S: px, py
S: px, pyS: px, pyMo: dxy, dx2y2, dz2
S: px, pz
Table 2. Formation energy Eformation (eV), energy of HOMO EH (eV), energy of LUMO EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3B, 4B, 5B, and 6B.
Table 2. Formation energy Eformation (eV), energy of HOMO EH (eV), energy of LUMO EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3B, 4B, 5B, and 6B.
3B4B5B6B
Eformation−120.27−200.28−297.42−415.25
EH−5.80−5.23−5.72−5.96
EL−7.00−7.33−6.74−7.07
Egap1.192.101.021.11
CHMo: dx2y2, dz2, dxz
S: px, py, pz
Mo: dz2
S: px, py, pz
Mo: dyz, dx2y2, dz2
S: px, py, pz
Mo: dxy, dz2
S: py, pz
CLMo: dxy, dx2y2, dz2
S: px, py, pz
Mo: dxz, dyz
S: px, py
Mo: dxy, dx2y2, dz2
S: px, py
Mo: dxy, dx2y2, dz2
S: px, py, pz
Table 3. Formation energy Eformation (eV), energy of HOMO EH (eV), energy of LUMO EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3C, 4C, 5C, and 6C.
Table 3. Formation energy Eformation (eV), energy of HOMO EH (eV), energy of LUMO EL (eV), energy gap between HOMO and LUMO Egap (eV), as well as orbital composition of HOMO CH and LUMO CL of 3C, 4C, 5C, and 6C.
3C4C5C6C
Eformation−137.42−219.60−321.09−441.37
EH−5.59−5.74−5.57−5.71
EL−6.54−6.67−6.74−6.79
Egap0.960.921.171.08
CHS: pxS: px, pzS: px, py, pzS: px, pz
CLMo: dxy, dx2y2, dz2
S: py, pz
Mo: dx2y2, dz2, dxz
S: px, py, pz
Mo: dxy, dx2y2, dz2
S: px, py, pz
Mo: dx2y2, dz2
S: px, py, pz
Table 4. Transition between the ground state and the lowest excited state S0→S1, transition between the ground state and the state with the largest oscillator strength S0→Sn (normally associate with the intense absorption in the electronic absorption spectrum, n is usually different in different clusters), and the percentage of Mo4d and S3p in electron and hole of 3A, 4A, 5A, and 6A.
Table 4. Transition between the ground state and the lowest excited state S0→S1, transition between the ground state and the state with the largest oscillator strength S0→Sn (normally associate with the intense absorption in the electronic absorption spectrum, n is usually different in different clusters), and the percentage of Mo4d and S3p in electron and hole of 3A, 4A, 5A, and 6A.
Electron, %Hole, %
3AS0→S1Mo4d: 60.29
S3p: 24.17
Mo4d: 72.76
S3p: 9.36
S0→S21Mo4d: 16.97
S3p: 64.86
Mo4d: 17.85
S3p: 62.80
4AS0→S1Mo4d: 64.27
S3p: 20.03
Mo4d: 60.14
S3p: 15.68
S0→S48Mo4d: 41.28
S3p: 28.71
Mo4d: 47.34
S3p: 26.80
5AS0→S1Mo4d: 72.11
S3p: 8.84
Mo4d: 70.92
S3p: 6.44
S0→S36Mo4d: 45.32
S3p: 23.84
Mo4d: 50.55
S3p: 12.13
6AS0→S1Mo4d: 65.27
S3p: 0
Mo4d: 60.44
S3p: 0
S0→S27Mo4d: 37.62
S3p: 31.75
Mo4d: 29.56
S3p: 35.33
Table 5. S0→S1 transition, S0→Sn transition, and the percentage of Mo4d and S3p in electron and hole of 3B, 4B, 5B, and 6B.
Table 5. S0→S1 transition, S0→Sn transition, and the percentage of Mo4d and S3p in electron and hole of 3B, 4B, 5B, and 6B.
Electron, %Hole, %
3BS0→S1Mo4d: 33.09
S3p: 56.59
Mo4d: 48.57
S3p: 39.28
S0→S73Mo4d: 53.17
S3p: 31.18
Mo4d: 31.98
S3p: 45.78
4BS0→S1Mo4d: 65.12
S3p: 18.77
Mo4d: 51.49
S3p: 23.19
S0→S17Mo4d: 53.68
S3p: 23.46
Mo4d: 6.98
S3p: 64.08
5BS0→S1Mo4d: 52.29
S3p: 32.33
Mo4d: 56.36
S3p: 24.52
S0→S25Mo4d: 49.56
S3p: 15.81
Mo4d: 12.53
S3p: 57.54
6BS0→S1Mo4d: 50.53
S3p: 31.77
Mo4d: 45.67
S3p: 26.18
S0→S20Mo4d: 57.88
S3p: 16.71
Mo4d: 6.52
S3p: 66.06
Table 6. S0→S1 transition, S0→Sn transition, and the percentage of Mo4d and S3p in electron and hole of 3C, 4C, 5C, and 6C.
Table 6. S0→S1 transition, S0→Sn transition, and the percentage of Mo4d and S3p in electron and hole of 3C, 4C, 5C, and 6C.
TransitionElectron, %Hole, %
3CS0→S1Mo4d: 58.26
S3p: 27.97
Mo4d: 20.09
S3p: 64.86
S0→S17Mo4d: 55.81
S3p: 26.89
Mo4d: 34.75
S3p: 43.73
4CS0→S1Mo4d: 47.55
S3p: 36.23
Mo4d: 2.55
S3p: 86.51
S0→S31Mo4d: 54.91
S3p: 41.67
Mo4d: 21.46
S3p: 74.25
5CS0→S1Mo4d: 57.86
S3p: 23.43
Mo4d: 2.79
S3p: 83.63
S0→S42Mo4d: 49.91
S3p: 31.92
Mo4d: 28.54
S3p: 49.41
6CS0→S1Mo4d: 55.02
S3p: 17.42
Mo4d: 2.45
S3p: 76.37
S0→S19Mo4d: 42.57
S3p: 21.73
Mo4d: 12.33
S3p: 53.19
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, S.; Han, C.; Ye, L.; Zhang, G.; Hu, Y.; Li, W.; Jiang, Y. Electronic Properties of Triangle Molybdenum Disulfide (MoS2) Clusters with Different Sizes and Edges. Molecules 2021, 26, 1157. https://doi.org/10.3390/molecules26041157

AMA Style

Wang S, Han C, Ye L, Zhang G, Hu Y, Li W, Jiang Y. Electronic Properties of Triangle Molybdenum Disulfide (MoS2) Clusters with Different Sizes and Edges. Molecules. 2021; 26(4):1157. https://doi.org/10.3390/molecules26041157

Chicago/Turabian Style

Wang, Songsong, Changliang Han, Liuqi Ye, Guiling Zhang, Yangyang Hu, Weiqi Li, and Yongyuan Jiang. 2021. "Electronic Properties of Triangle Molybdenum Disulfide (MoS2) Clusters with Different Sizes and Edges" Molecules 26, no. 4: 1157. https://doi.org/10.3390/molecules26041157

Article Metrics

Back to TopTop