Next Article in Journal
Synthesis and Evaluation of FeSX/TiO2 for the Photocatalytic Degradation of Phenol under Visible-Light Region
Next Article in Special Issue
Application of Bioelectrochemical System and Magnetite Nanoparticles on the Anaerobic Digestion of Sewage Sludge: Effect of Electrode Configuration
Previous Article in Journal
Advances in the Catalytic Conversion of Biomass Components to Ester Derivatives: Challenges and Opportunities
Previous Article in Special Issue
The Impact of Alternative Fuels on Ship Engine Emissions and Aftertreatment Systems: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigations on Structural, Electronic and Optical Properties of MoS2/CDs Heterostructure via First-Principles Study

1
Ansteel Beijing Research Institute Co., Ltd., Beijing 102209, China
2
Pangang Group Vanadium Titanium & Resources Co., Ltd., Panzhihua 617067, China
3
School of Chemical Engineering & Technology, Hebei University of Technology, Tianjin 300401, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 11 March 2022 / Revised: 11 April 2022 / Accepted: 15 April 2022 / Published: 20 April 2022

Abstract

:
Much effort has been made for MoS2/CDs heterostructure application in the field of photocatalysts. However, the impacts of functional groups of CDs on the properties of the heterostructure are ambiguous. Here, the impacts of hydroxyl, carbonyl, and carboxyl groups of CDs on the structural, electronic, and optical properties of MoS2/CDs’ heterostructure were investigated by conducting a first-principles study. The calculated energy band structure and band gap of monolayer MoS2 were consistent with the experimental values. The band gap of MoS2 was obviously decreased after the construction of MoS2/CDs and MoS2/CDs–hydroxyl/carboxyl, thus helping to improve the light adsorption range. However, the band gap of MoS2/CDs–carbonyl was slightly increased compared with that of monolayer MoS2. The CDs with functional groups can spontaneously bind on 2D-MoS2 and form a stable MoS2/CDs heterostructure. It was confirmed that the MoS2/CDs’ heterostructure belongs to the typical type-II band alignment, which contributes to the separation of photogenerated charge and hole. Notably, the carbonyl and carboxyl groups on the CDs obviously reduced the optical absorption intensity of the MoS2/CDs in the ultraviolet region. The hydroxyl groups have little effect on optical absorption intensity. Thus, the CDs with more hydroxyl groups are beneficial to produce a higher photocatalytic performance. This paper reveals the impacts of surface functional groups and provides a promising approach for designing the MoS2/CDs’ heterostructure to enhance the photocatalytic properties.

1. Introduction

Molybdenum disulfide (MoS2) is one of the most studied novel material among transition metal sulfides, due to its potential applications in electronic and optoelectronic devices [1,2,3], hydrogen evolution reaction [4,5,6], and energy storage [7,8,9]. However, MoS2 also has some disadvantages in practical applications, including the inert basal plane, untunable band gap, and low conductivity [10,11]. Therefore, in order to activate the basal plane, enhance conductivity, and tune band-gap, MoS2-based heterostructures and hybrid systems have been built. The MoS2-based heterostructures include carbon materials (graphene, graphene oxide, graphene quantum dot, carbon dot, and carbon nanotube), noble metals (Ag, Au, and Pt), TiO2, Cu2O, MoO3, and other two-dimensional materials (WS2, WSe, BN, and Mxenes) [12,13,14,15,16,17,18,19,20,21], exhibiting the better performance in terms of electronics, optoelectronics, catalytics, and energy storage.
Carbon materials play a key role in these MoS2-based heterostructures, due to their unique electronic properties. However, some unexpected functional groups always are produced from the preparation of carbon materials such as graphene, graphene oxide, and carbon dot. The impacts of functional groups on MoS2-based heterostructures are vague. In this paper, we chose carbon dots (CDs) as a representative carbon material for researching the impacts of functional groups on the electronic properties. Carbon dots are a common carbon material, because they have extensive applications in photocatalysis [22], bio-imaging [23], sensor [24], and light-emitting diodes [25], due to high quantum yields and tunable light-emission bands. Chen et al. [26] prepared 1L-MoS2/CDs ultra-thin light emission films with full-color tunable photoluminescence, which greatly extended the light emission bands of MoS2. Badhulika et al. [16] first fabricated a 2D-MoS2-CDs-based flexible broadband photodetector covering the entire range of the electromagnetic spectrum, due to the combination of UV illumination of CDs and visible and NIR illumination of MoS2. To date, the facile and green preparation of CDs is the carbonization of carbon sources, including glucose [27], citric acid [26], sugar [28], and chia seed [16]. Many experiments have demonstrated that the hydroxyl, carbonyl, and carboxyl groups exist in the edge of CDs due to the incomplete carbonization [27,29]. It is well-known that the surface functional groups have a strong influence on electronic structures, optical characteristics, and conductivity [26,30,31,32,33]. However, due to the difficulty of synthesizing CDs with a specific functional group, the influence of a specific functional group is hard to study through experimentation. Therefore, it is of great significance to study it through simulation calculations.
Density functional theory (DFT) calculation is a powerful and scientific technology that is used to research the intrinsic properties of nanomaterials and provide theoretical guidance for experiments and applications. Some papers have reported the effects of vacancies or heterostructure on the structural, electronic, and optical properties of monolayer MoS2 by DFT [30,31]. However, the influences of the surface properties of CDs on the properties of MoS2/CDs were less investigated through simulation. In this study, the effects of hydroxyl, carbonyl, and carboxyl groups of CDs on the structural, electronic, and optical properties of MoS2/CDs’ heterostructure were revealed by applying density functional theory. This study contributes to improving the optical property of MoS2/CDs’ heterostructure.

2. Computational Details

All the first-principles calculations were performed within the framework of the plane-wave pseudopotential density functional theory (DFT) implemented in CASTEP code with ultrasoft pseudopotential method on the basis of DFT [34,35,36]. The generalized gradient approximation (GGA) with Perdew–Bruke–Ernzerhof (PBE) functional and van der Waals correction proposed by the Grimme potential were employed [34,37]. This method has widely been employed to study the structural, electronic, and photoelectronic properties of MoS2-based heterostructure [34,38,39,40]. All structures were relaxed until the force on each atom was less than 0.01 eV Å−1. The energy cutoff was 400 eV. The two-dimensional (2D) periodic boundary conditions were considered along the growth directions of the MoS2/CDs layer. The Monkhorst–Pack k-point mesh was sampled with a separation of 0.05 and 0.015 Å−1 in the Brillouin zone during the relaxation and calculation periods, respectively.

3. Results and Discussions

First, a 4 × 4 supercell of MoS2 monolayer composed of 16 molybdenum atoms and 32 sulfur atoms was constructed as a substrate (Figure 1a). A zigzag-edged C24H12 structure was used, and all edge carbon atoms of CDs were assumed to be fully passivated by H atoms to eliminate the dangling bonds (Figure 1b) [34,41]. It has been widely accepted that the carboxyl, hydroxyl, and carbonyl are three dominant functional groups on the CDs. Therefore, the CDs-T (T is carboxyl, hydroxyl, and carbonyl, respectively; CDs-T represents that only one functional group T is on the CDs) were constructed for studying the impacts of functional groups on the structure, electronic, and optical absorption of MoS2/CDs. In order to match the experimental data, the functional groups were place on the edge of CDs. After geometry optimization, the lattice constants of the MoS2 monolayer were a = b = 3.191 Å and α = β = 90°, γ = 120°, which is comparable to the experimentally measure for the bulk MoS2 [42]. To evaluate the structural stability of MoS2/CDs-T, the binding energies were calculated by using the Equation (1) [39].
Eb = E(MoS2/CDs-T) − E(MoS2) − E(CDs-T)
where E(MoS2/CDs-T), E(MoS2), and E(CDs-T) are the total energies of MoS2/CDs-T, MoS2, and CDs-T, respectively.
Figure 2a shows the Eb and h (the distance between MoS2 and CDs and CDs-T) in the heterostructures. The optimized h value ranges from 3.267 to 3.358 Å, and the closest distance is 3.267 Å of MoS2/CDs-OH. The bigger carboxyl groups on the CDs increase the distance due to the steric effect. The negative binding energy suggests that CDs spontaneously form the MoS2/CDs’ heterostructure and keep it in a stable state. A negative binding energy indicates that the two substances can spontaneously bind and form a stable heterojunction. Conversely, a positive binding energy indicates that the two substances cannot bind spontaneously. Therefore, the carboxyl, hydroxyl, and carbonyl groups on the CDs can still form the stability of the MoS2/CDs-T heterostructures. Figure 2b,c shows the band structure and partial density of the MoS2 states. The direct band gap (the conduction band minimum and the valence band maximum are located at the K symmetry point indicating) of 1.75 eV (Figure 2b) for monolayer MoS2 is in good agreement with other theoretical values (1.78 eV [43], 1.74 eV [34], and 1.70 eV [43]) and experimental results (1.75 eV [44] and 1.80 eV [45]). As shown in Figure 2c, the valence bands and the conduction bands are essentially hybridized from the Mo 4d and S 3p orbitals, and this is in accordance with the previous calculation results [34,46].
Figure 3 displays the PDOS of CDs and CDs-T; the insets in these figures are the corresponding band structures. Electron transfer and sharing between functional groups and carbon atoms lead to changes in the electronic structure. The valence bands of CDs from −5 to −2 eV are mainly hybridized from C 2p and H 1s orbitals, while the valence bands from −2 to 0 eV and the conduction bands from 0 to 4 eV are composed only of C 2p orbital (Figure 3a). Moreover, the energy levels of CDs are discrete, owing to the quantum confinement [34]. The energy gap of CDs between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) is 2.859 eV, which is really agreement with previously reported results (2.890 eV [34] and 2.90 eV [41]). It is worth noting that the energy gaps calculated by DFT are lower than the experimental result (5.56 eV [47] and 6.82 eV [48]), which results from the most used local and semi-local approximations for exchange–correlation functional in standard DFT calculation liable to underestimate energy gaps [46]. After hydroxyl, carbonyl, and carboxyl were added to CDs, the energy gaps of CDs-T between LUMO and HOMO were 2.637, 2.236, and 2.110 eV, respectively. As seen in Figure 3b–d, the valence bands of CDs-T from −5 to −2 eV and the conduction bands from 3.5 to 5 eV are hybridized from C 2p, O 2p, and H 1s orbitals, while the valence bands from −2 to 0 eV and the conduction bands from 0 to 3.5 eV are hybridized from major C 2p and a small part portion of O 2p orbitals. The bottom of the conduction bands and top of the valence bands of MoS2/CDs-T consist of the hybridization from the C 2p and O 2p orbitals.
The PDOS and band structure of MoS2/CDs and MoS2/CDs-T heterostructures are shown in Figure 4. In the LUMO, the conduction bands of these heterostructures are mainly hybridized from Mo 4d and S 3p orbitals. There are energy levels under the Femi level. For MoS2/CDs, MoS2/CDs-OH, and COOH, these energy levels are mainly composed of the C 2p orbital. However, the energy level at the Femi level of MoS2/CDs-O is hybridized from the C 2p and O 2p orbitals. The electronic structure indicates that the Mo 4d, S 3p, C 2p, and O 2p orbitals are the main way of photo-induced electron. Under the illumination, the electrons are transferred from CDs to MoS2 and leave holes on the CDs, which contribute to the separation of electron and hole. The energy bands of MoS2/CDs, MoS2/CDs–hydroxyl, and carboxyl are reduced to 1.496, 1.320, and 1.248 eV, respectively, while the energy band in MoS2/CDs–carbonyl is 1.772 eV.
The optical absorption of the catalyst plays a key role in the photocatalytic process. Figure 5a shows the optical absorption spectra of these samples, namely CDs-T, MoS2, and MoS2/CDs-T. The frequency-dependent dielectric matrix, ε (ω) = ε1 (ω) + iε2 (ω), is calculated, where ε1 (ω) and ε2 (ω) are the real and imaginary parts of the dielectric function, respectively. The ε1 (ω) is determined by summation over electronic states, using the Equation (2). [49]:
ε 1   ( ω ) = 1 + 2 π P 0 ε 2 ( ω ) ω   ω 2 ω 2 + i η d ω
The ε2 (ω) is obtained by using the Kramers–Kronig relationship [45]:
ε 2   ( ω ) = 4 π 2 e 2 Ω lim q 0 1 q 2 c , v ,   k 2 ω k δ ( ε c k ε v k ω )   ×   μ c k   +   e q | μ v k   μ v k   +   e β q | μ v k
The light adsorption intensity is evaluated according to the Equation (4). [45]:
α   ( E ) = 4 π e h c { [ ε 1 2 + ε 2 2 ] 1 / 2 ε 1 2 } 1 / 2
The calculated ε1 (ω) and ε2 (ω) of dielectric function versus energy of MoS2, CDs-T, and MoS2/CDs-T are shown in Figure 5. The peak A (3.01 eV), B(4.96 eV), and C(10.94 eV) can be observed in the imaginary part of MoS2. These peaks are located from 0 to 30 eV, due to the absorptive transition from the valence bands to the conduction bands. According to the structure analysis, the peaks of A, B, and C result from the transition of S 3p into Mo 4d conduction bands [50], the hybridization orbitals between S 3p and Mo 4d into Mo 4d conduction bands [51], and σ bonding between S 3p and Mo 5s into Mo 4d conduction bands [52], respectively. For MoS2/CDs-T, they have similar profiles with that of MoS2. Notably, the imaginary parts of MoS2/CDs-T move slightly toward lower energies because of the effects of heterostructure [53]. The calculated static constants of MoS2 is about 5.0 eV, which is consistent with other reported values of 5.3 eV calculated by the generalized gradient approximation (GGA) with Perdew–Bruke–Ernzerhof functional [54]. It is obvious that the static dielectric constants of MoS2/CDs-T are bigger than those of of MoS2. Figure 6a shows the optical absorption spectra of all samples. There is strong absorption in the range of 200 to 700 nm for the MoS2 monolayer, which is potential for applications in sunlight-driven photocatalysis. The main peak at about 450 nm is attributed to the transition from the occupied Mo-4d orbital to the unoccupied S-3p orbital [55]. However, the MoS2 monolayer has a low adsorption capacity at the long wavelength end, and this results from the electronic structure of the monolayer. The absorption peaks of CDs are mainly located at the ultraviolet region from 200 to 400 nm, and the absorption intensity is relatively low. The obvious red shifts of the absorption peaks and the reduction of the absorption intensities can be observed after the MoS2/CDs’ heterostructures were constructed. Compared to the MoS2 monolayer, the absorption intensities ranging from 200 nm to the long wavelength were obviously increased by CDs. However, the carbonyl and carboxyl groups of CDs decrease the absorption intensities of MoS2/CDs in the ultraviolet region. In contrast, the hydroxyl groups have a slight impact on the absorption. It can be concluded that CDs are suitable for promoting the photocatalysis of MoS2 due to the enhanced adsorption intensity. However, the carbonyl and carboxyl groups have a negative effect on the photocatalysis of MoS2/CDs, causing a decrease in adsorption intensity.
To investigate the redox ability of MoS2/CDs-T, the valence band maximum (VBM) and conduction band minimum (CBM) relative to the vacuum energy level were calculated according to the Equations (5) and (6). [55]:
EVB = φ = V (∞) − EF
ECB = EVB − Eg
where V (∞), EF, and Eg are the electrostatic potential in a vacuum region, the Fermi level of the neutral surface system, and the band gap, respectively. As shown in Figure 6b, the LUMO and HOMO of CDs-T are placed at higher energy states than the CBM and VBM of MoS2 monolayer, respectively. It indicated that MoS2/CDs heterostructures can construct the typical type-II band alignments [34], which promote photo-induced charge separation and transfer. Combining with the results from PDOS and band structures (Figure 4), the photo-generated electron of C 2p orbitals of CDs and CDs-T will transfer to the S 3p and Mo 4d hybrid orbitals (Figure 6c).

4. Conclusions

In summary, the impacts of hydroxyl, carbonyl, and carboxyl of the CDs on the structural, electronic, and optical properties of MoS2/CDs were studied by employing density function theory. The more negative binding energies (−0.36 eV to −1.17 eV) between MoS2 and CDs-T suggest that the hydroxyl, carbonyl, and carboxyl groups contribute to spontaneous combination and the stability structure. The three groups show opposite effects on the band gap of MoS2/CDs-T. The hydroxyl and carboxyl greatly reduce the band gap (1.32 and 1.25 eV, respectively) compared with the band gap of MoS2 and MoS2/CDs. However, the carbonyl group slightly increases the band gap (1.77 eV). The heterostructure of MoS2/CDs-T belongs to typical type-II band alignment, which is beneficial to the separation of photo-induced charge and hole. Notably, the carbonyl and carboxyl groups of CDs reduce the absorption intensity of MoS2/CDs. In the preparation of MoS2/CDs, the generation of carbonyl and carboxyl groups should be avoided as much as possible. Increasing the content of hydroxyl groups in carbon quantum dots is more conducive to improving the photocatalytic performance.

Author Contributions

Conceptualization, X.Y. and A.T.; methodology X.Y.; software, P.Y., A.T., X.Y., D.Z. and J.Y.; validation, Z.C., X.Y. and A.T.; formal analysis, P.Y., D.Z., J.Y., X.Y. and A.T.; investigation, X.Y. and A.T.; data curation, A.T.; writing—original draft preparation, X.Y. and A.T.; writing—review and editing, Z.C.; visualization, Z.C. and D.Z.; supervision, J.Y.; project administration, X.Y. and A.T.; funding acquisition, P.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by High Level Talents Funding Project of Hebei: Postdoctoral Research Project-Preferential Funding: B2020003021; Hebei Natural Science Foundation for Young Scientists: E2021202169. The APC was funded by Ansteel Beijing Research Institute Co., Ltd.

Data Availability Statement

Data are contained within the article.

Acknowledgments

Ansteel Beijing Research Institute Co., Ltd.; Pangang Group Vanadium Titanium & Resources Co., Ltd.; High Level Talents Funding Project of Hebei: Postdoctoral Research Project-Preferential Funding (B2020003021), and Hebei Natural Science Foundation for Young Scientists (E2021202169).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Singh, E.; Singh, P.; Kim, K.S.; Yeom, G.Y.; Nalwa, H.S. Flexible Molybdenum Disulfide (MoS2) Atomic Layers for Wearable Electronics and Optoelectronics. ACS Appl. Mater. Interfaces 2019, 11, 11061–11105. [Google Scholar] [CrossRef] [PubMed]
  2. Yuan, J.; Wang, F.; Patel, S.; Hu, Z.-L.; Tang, M.; Lou, J. Minimizing the Water Effect in Synthesis of High-Quality Monolayer MoS2 Nanosheets: Implications for Electronic and Optoelectronic Devices. ACS Appl. Nano Mater. 2021, 4, 8094–8100. [Google Scholar] [CrossRef]
  3. Tang, K.; Qi, W.; Li, Y.; Wang, T. Electronic properties of van der Waals heterostructure of black phosphorus and MoS2. J. Phys. Chem. C 2018, 122, 7027–7032. [Google Scholar] [CrossRef]
  4. Morales-Guio, C.; Hu, X. Amorphous Molybdenum Sulfides as Hydrogen Evolution Catalysts. Acc. Chem. Res. 2014, 47, 2671–2681. [Google Scholar] [CrossRef] [Green Version]
  5. Huang, Y.; Sun, Y.; Zheng, X.; Aoki, T.; Pattengale, B.; Huang, J.; He, X.; Bian, W.; Younan, S.; Williams, N.; et al. Atomically engineering activation sites onto metallic 1T-MoS2 catalysts for enhanced electrochemical hydrogen evolution. Nat. Commun. 2019, 10, 982. [Google Scholar] [CrossRef]
  6. Lau, T.H.M.; Lu, X.; Kulhavý, J.; Wu, S.; Lu, L.; Wu, T.-S.; Kato, R.; Foord, J.S.; Soo, Y.-L.; Suenaga, K.; et al. Transition metal atom doping of the basal plane of MoS2 monolayer nanosheets for electrochemical hydrogen evolution. Chem. Sci. 2018, 9, 4769–4776. [Google Scholar] [CrossRef] [Green Version]
  7. Qin, W.; Chen, T.; Pan, L.; Niu, L.; Hu, B.; Li, D.; Li, J.; Sun, Z. MoS2-reduced graphene oxide composites via microwave assisted synthesis for sodium ion battery anode with improved capacity and cycling performance. Electrochim. Acta 2015, 153, 55–61. [Google Scholar] [CrossRef]
  8. Zhou, Q.; Li, W.; Gao, M.; Xu, H.; Guo, Y.; Sun, L.; Zheng, D.; Lin, J. A truncated octahedron metal-organic framework derived TiO2@C@MoS2 composite with superior lithium-ion storage properties. J. Power Sources 2021, 518, 230746. [Google Scholar] [CrossRef]
  9. Poudel, M.B.; Karki, H.P.; Kim, H.J. Silver nanoparticles decorated molybdenum sulfide/tungstate oxide nanorods as high performance supercapacitor electrode. J. Energy Storage 2020, 32, 101693. [Google Scholar] [CrossRef]
  10. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nelsen, 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]
  11. Ganatra, R.; Zhang, Q. Few-Layer MoS2: A Promising Layered Semiconductor. ACS Nano 2014, 8, 4074–4099. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, X.; Huangfu, L.; Gu, Z.; Xiao, S.; Zhou, J.; Nan, H.; Gu, X.; Ostrikov, K. Controllable Epitaxial Growth of Large-Area MoS 2 /WS 2 Vertical Heterostructures by Confined-Space Chemical Vapor Deposition. Small 2021, 17, 2007312. [Google Scholar] [CrossRef] [PubMed]
  13. Zafiropoulou, I.; Katsiotis, M.S.; Boukos, N.; Karakassides, M.A.; Stephen, S.; Tzitzios, V.; Fardis, M.; Vladea, R.V.; Alhassan, S.M.; Papavassiliou, G. In Situ Deposition and Characterization of MoS2 Nanolayers on Carbon Nanofibers and Nanotubes. J. Phys. Chem. C 2013, 117, 10135–10142. [Google Scholar] [CrossRef]
  14. Jing, Y.; Ortiz-Quiles, E.O.; Cabrera, C.R.; Chen, Z.; Zhou, Z. Layer-by-Layer Hybrids of MoS2 and Reduced Graphene Oxide for Lithium Ion Batteries. Electrochim. Acta 2014, 147, 392–400. [Google Scholar] [CrossRef]
  15. Yin, Z.; Zhang, X.; Cai, Y.; Chen, J.; Wong, J.I.; Tay, Y.-Y.; Chai, J.; Wu, J.; Zeng, Z.; Zheng, B.; et al. Preparation of MoS2-MoO3Hybrid Nanomaterials for Light-Emitting Diodes. Angew. Chem. Int. Ed. 2014, 53, 12560–12565. [Google Scholar] [CrossRef]
  16. Sahatiya, P.; Jones, S.S.; Badhulika, S. 2D MoS2–carbon quantum dot hybrid based large area, flexible UV–vis–NIR photodetector on paper substrate. Appl. Mater. Today 2018, 10, 106–114. [Google Scholar] [CrossRef]
  17. Thangasamy, P.; Partheeban, T.; Sudanthiramoorthy, S.; Sathish, M. Enhanced Superhydrophobic Performance of BN-MoS2 Heterostructure Prepared via a Rapid, One-Pot Supercritical Fluid Processing. Langmuir 2017, 33, 6159–6166. [Google Scholar] [CrossRef]
  18. Zhao, Y.-F.; Yang, Z.-Y.; Zhang, Y.-X.; Jing, L.; Guo, X.; Ke, Z.; Hu, P.; Wang, G.; Yan, Y.-M.; Sun, K.-N. Cu2O Decorated with Cocatalyst MoS2 for Solar Hydrogen Production with Enhanced Efficiency under Visible Light. J. Phys. Chem. C 2014, 118, 14238–14245. [Google Scholar] [CrossRef]
  19. Ramakrishnan, S.; Karuppannan, M.; Vinothkannan, M.; Ramachandran, K.; Kwon, O.J.; Yoo, D.J. Ultrafine Pt Nanoparticles Stabilized by MoS2/N-Doped Reduced Graphene Oxide as a Durable Electrocatalyst for Alcohol Oxidation and Oxygen Reduction Reactions. ACS Appl. Mater. Interfaces 2019, 11, 12504–12515. [Google Scholar] [CrossRef]
  20. Poudel, M.B.; Ojha, G.P.; A Kim, A.; Kim, H.J. Manganese-doped tungsten disulfide microcones as binder-free electrode for high performance asymmetric supercapacitor. J. Energy Storage 2021, 47, 103674. [Google Scholar] [CrossRef]
  21. Guo, J.; Zhu, H.; Sun, Y.; Tang, L.; Zhang, X. Doping MoS2 with Graphene Quantum Dots: Structural and Electrical Engineering towards Enhanced Electrochemical Hydrogen Evolution. Electrochim. Acta 2016, 211, 603–610. [Google Scholar] [CrossRef]
  22. Heng, Z.W.; Chong, W.C.; Pang, Y.L.; Koo, C.H. An overview of the recent advances of carbon quantum dots/metal oxides in the application of heterogeneous photocatalysis in photodegradation of pollutants towards visible-light and solar energy ex-ploitation. J. Environ. Chem. Eng. 2021, 9, 105199. [Google Scholar] [CrossRef]
  23. Chen, W.; Li, D.; Tian, L.; Xiang, W.; Wang, T.; Hu, W.; Hu, Y.; Chen, S.; Chen, J.; Dai, Z. Synthesis of graphene quantum dots from natural polymer starch for cell imaging. Green Chem. 2018, 20, 4438–4442. [Google Scholar] [CrossRef]
  24. Luo, Z.; Qi, G.; Chen, K.; Zou, M.; Yuwen, L.; Zhang, X.; Huang, W.; Wang, L. Microwave-Assisted Preparation of White Fluorescent Graphene Quantum Dots as a Novel Phosphor for Enhanced White-Light-Emitting Diodes. Adv. Funct. Mater. 2016, 26, 2739–2744. [Google Scholar] [CrossRef]
  25. Guo, X.; Wang, C.-F.; Yu, Z.-Y.; Chen, L.; Chen, S. Facile access to versatile fluorescent carbon dots toward light-emitting diodes. Chem. Commun. 2012, 48, 2692–2694. [Google Scholar] [CrossRef]
  26. Su, W.; Wang, Y.; Wu, W.; Qin, H.; Song, K.; Huang, X.; Zhang, L.; Chen, D. Towards full-colour tunable photoluminescence of monolayer MoS2/carbon quantum dot ultra-thin films. J. Mater. Chem. C 2017, 5, 6352–6358. [Google Scholar] [CrossRef]
  27. Gan, Z.; Wu, X.; Hao, Y. The mechanism of blue photoluminescence from carbon nanodots. CrystEngComm 2014, 16, 4981–4986. [Google Scholar] [CrossRef]
  28. Zuo, P.; Lu, X.; Sun, Z.; Guo, Y.; He, H. A review on syntheses, properties, characterization and bioanalytical applications of fluorescent carbon dots. Mikrochim. Acta 2015, 183, 519–542. [Google Scholar] [CrossRef]
  29. Sevilla, M.; Fuertes, A.B. Chemical and Structural Properties of Carbonaceous Products Obtained by Hydrothermal Carbonization of Saccharides. Chem. Eur. J. 2009, 15, 4195–4203. [Google Scholar] [CrossRef]
  30. Poudel, M.B.; Awasthi, G.P.; Kim, H.J. Novel insight into the adsorption of Cr(VI) and Pb(II) ions by MOF derived Co-Al layered double hydroxide @hematite nanorods on 3D porous carbon nanofiber network. Chem. Eng. J. 2021, 417, 129312. [Google Scholar] [CrossRef]
  31. Poudel, M.B.; Kim, H.J. Confinement of Zn-Mg-Al-layered double hydroxide and α-Fe2O3 nanorods on hollow porous carbon nanofibers: A free-standing electrode for solid-state symmetric supercapacitors. Chem. Eng. J. 2021, 429, 132345. [Google Scholar] [CrossRef]
  32. Yan, J.-A.; Chou, M.-Y. Oxidation functional groups on graphene: Structural and electronic properties. Phys. Rev. B 2010, 82, 125403. [Google Scholar] [CrossRef]
  33. Tang, S.; Wu, W.; Zhang, S.; Ye, D.; Zhong, P.; Li, X.; Liu, L.; Li, Y.-F. Tuning the activity of the inert MoS2 surface via graphene oxide support doping towards chemical functionalization and hydrogen evolution: A density functional study. Phys. Chem. Chem. Phys. 2018, 20, 1861–1871. [Google Scholar] [CrossRef] [PubMed]
  34. Li, N.; Liu, Z.; Hu, S.; Chang, Q.; Xue, C.; Wang, H. Electronic and photocatalytic properties of modified MoS2/graphene quantum dots heterostructures: A computational study. Appl. Surf. Sci. 2019, 473, 70–76. [Google Scholar] [CrossRef]
  35. Danovich, M.; Aleiner, I.L.; Drummond, N.D.; Fal’Ko, V. Fast Relaxation of Photo-Excited Carriers in 2-D Transition Metal Dichalcogenides. IEEE J. Sel. Top. Quantum Electron. 2016, 23, 168–172. [Google Scholar] [CrossRef] [Green Version]
  36. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  37. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  38. Su, J.; Feng, L.-P.; Zheng, X.; Hu, C.; Lu, H.; Liu, Z. Promising Approach for High-Performance MoS2 Nanodevice: Doping the BN Buffer Layer to Eliminate the Schottky Barriers. ACS Appl. Mater. Interfaces 2017, 9, 40940–40948. [Google Scholar] [CrossRef]
  39. Zhang, J.; Lang, X.Y.; Zhu, Y.F.; Jiang, Q. Strain tuned InSe/MoS2 bilayer van der Waals heterostructures for photovoltaics or photocatalysis. Phys. Chem. Chem. Phys. 2018, 20, 17574–17582. [Google Scholar] [CrossRef]
  40. Kośmider, K.; Fernández-Rossier, J. Electronic properties of the MoS2-WS2heterojunction. Phys. Rev. B 2013, 87, 075451. [Google Scholar] [CrossRef] [Green Version]
  41. Li, Y.; Shu, H.; Wang, S.; Wang, J. Electronic and Optical Properties of Graphene Quantum Dots: The Role of Many-Body Effects. J. Phys. Chem. C 2015, 119, 4983–4989. [Google Scholar] [CrossRef]
  42. Petkov, V.; Billinge, S.J.L.; Larson, P.; Mahanti, S.D.; Vogt, T.; Rangan, K.K.; Kanatzidis, M.G. Structure of nanocrystalline materials using atomic pair distribution function analysis: Study of LiMoS2. Phys. Rev. B 2002, 65, 092105. [Google Scholar] [CrossRef] [Green Version]
  43. Lauritsen, J.; 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]
  44. 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]
  45. Lee, H.S.; Min, S.-W.; Chang, Y.-G.; Park, M.K.; Nam, T.; Kim, H.; Kim, J.H.; Ryu, S.; Im, S. MoS2 Nanosheet Phototransistors with Thickness-Modulated Optical Energy Gap. Nano Lett. 2012, 12, 3695–3700. [Google Scholar] [CrossRef]
  46. Feng, L.-P.; Sun, H.-Q.; Li, A.; Su, J.; Zhang, Y.; Liu, Z.-T. Influence of Mo-vacancy concentration on the structural, electronic and optical properties of monolayer MoS2: A first-principles study. Mater. Chem. Phys. 2018, 209, 146–151. [Google Scholar] [CrossRef]
  47. Ong, W.-J.; Putri, L.K.; Tan, Y.-C.; Tan, L.-L.; Li, N.; Ng, Y.H.; Wen, X.; Chai, S.-P. Unravelling charge carrier dynamics in protonated g-C3N4 interfaced with carbon nanodots as co-catalysts toward enhanced photocatalytic CO2 reduction: A combined experimental and first-principles DFT study. Nano Res. 2017, 10, 1673–1696. [Google Scholar] [CrossRef]
  48. Dürr, H. Handbook of Photochemistry; CRC Press: Boca Raton, FL, USA, 1994. [Google Scholar]
  49. Gajdoš, M.; Hummer, K.; Kresse, G.; Furthmüller, J.; Bechstedt, F. Linear optical properties in the projector-augmented wave methodology. Phys. Rev. B 2006, 73, 045112. [Google Scholar] [CrossRef] [Green Version]
  50. Shi, H.; Pan, H.; Zhang, Y.-W.; Yakobson, B.I. Quasiparticle band structures and optical properties of strained monolayer MoS2 and WS2. Phys. Rev. B 2013, 87, 155304. [Google Scholar] [CrossRef] [Green Version]
  51. Kumar, A.; Ahluwalia, P. Ahluwalia, A first principle comparative study of electronic and optical properties of 1H-MoS2 and 2H-MoS2. Mater. Chem. Phys. 2012, 135, 755–761. [Google Scholar] [CrossRef]
  52. Liang, W.Y.; A Wilson, J.; Yoffe, A.D. Optical studies of metal-semiconductor transmutations produced by intercalation. J. Phys. C: Solid State Phys. 1971, 4, L18–L20. [Google Scholar] [CrossRef] [Green Version]
  53. Feng, L.-P.; Su, J.; Liu, Z.-T. Effect of vacancies on structural, electronic and optical properties of monolayer MoS2: A first-principles study. J. Alloys Compd. 2014, 613, 122–127. [Google Scholar] [CrossRef]
  54. Yang, C.-L.; Wang, M.-S.; Ma, X.G. First-principles study on the electronic and optical properties of WS2 and MoS2 monolayers. Chin. J. Phys. 2017, 55, 1930–1937. [Google Scholar] [CrossRef]
  55. Li, Y.; Li, Y.-L.; Araujo, C.M.; Luo, W.; Ahuja, R. Single-layer MoS2 as an efficient photocatalyst. Catal. Sci. Technol. 2013, 3, 2214–2220. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Geometrical structures of MoS2 monolayer (a), CDs (b), MoS2/CDs heterostructures without groups (c), MoS2/CDs heterostructures with carboxyl (d), hydroxyl (e), and carbonyl (f).
Figure 1. Geometrical structures of MoS2 monolayer (a), CDs (b), MoS2/CDs heterostructures without groups (c), MoS2/CDs heterostructures with carboxyl (d), hydroxyl (e), and carbonyl (f).
Catalysts 12 00456 g001aCatalysts 12 00456 g001b
Figure 2. The binding energy (Eb) and distance (h) of MoS2/CDs-T (a), band structure of MoS2 monolayer (b), and partial density of states (PDOS) of MoS2 monolayer (c).
Figure 2. The binding energy (Eb) and distance (h) of MoS2/CDs-T (a), band structure of MoS2 monolayer (b), and partial density of states (PDOS) of MoS2 monolayer (c).
Catalysts 12 00456 g002
Figure 3. Partial density of states (PDOS) of pure CDs (a), CDs-T (T is hydroxyl (b), carbonyl (c), and carboxyl (d)); the insets in these figures are the corresponding band structures.
Figure 3. Partial density of states (PDOS) of pure CDs (a), CDs-T (T is hydroxyl (b), carbonyl (c), and carboxyl (d)); the insets in these figures are the corresponding band structures.
Catalysts 12 00456 g003
Figure 4. Partial density of states (PDOS) of MoS2/CDs heterostructure (a), MoS2/CDs-T heterostructures (T is hydroxyl (b), carbonyl (c), and carboxyl (d)); the insets in these figures are the corresponding band structures.
Figure 4. Partial density of states (PDOS) of MoS2/CDs heterostructure (a), MoS2/CDs-T heterostructures (T is hydroxyl (b), carbonyl (c), and carboxyl (d)); the insets in these figures are the corresponding band structures.
Catalysts 12 00456 g004
Figure 5. Real parts, ε 1   ( ω ) of (a) MoS2/CDs, (c) MoS2/CDs-OH, (e) MoS2/CDs-OH, and (g) MoS2/CDs-COOH. Imaginary parts, ε 2   ( ω ) of (b) MoS2/CDs, (d) MoS2/CDs-OH, (f) MoS2/CDs-OH, and (h) MoS2/CDs-COOH. Imaginary parts.
Figure 5. Real parts, ε 1   ( ω ) of (a) MoS2/CDs, (c) MoS2/CDs-OH, (e) MoS2/CDs-OH, and (g) MoS2/CDs-COOH. Imaginary parts, ε 2   ( ω ) of (b) MoS2/CDs, (d) MoS2/CDs-OH, (f) MoS2/CDs-OH, and (h) MoS2/CDs-COOH. Imaginary parts.
Catalysts 12 00456 g005aCatalysts 12 00456 g005b
Figure 6. Optical absorption spectra (a); the calculated valence band maximum (VBM) and conduction band minimum (CBM) of MoS2, CDs, CDs-OH, CDs-O, and CDs-COOH; (b) and MoS2/CDs-T (c).
Figure 6. Optical absorption spectra (a); the calculated valence band maximum (VBM) and conduction band minimum (CBM) of MoS2, CDs, CDs-OH, CDs-O, and CDs-COOH; (b) and MoS2/CDs-T (c).
Catalysts 12 00456 g006aCatalysts 12 00456 g006b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yin, X.; Teng, A.; Chang, Z.; Yuan, P.; Zhang, D.; Yu, J. Investigations on Structural, Electronic and Optical Properties of MoS2/CDs Heterostructure via First-Principles Study. Catalysts 2022, 12, 456. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050456

AMA Style

Yin X, Teng A, Chang Z, Yuan P, Zhang D, Yu J. Investigations on Structural, Electronic and Optical Properties of MoS2/CDs Heterostructure via First-Principles Study. Catalysts. 2022; 12(5):456. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050456

Chicago/Turabian Style

Yin, Xianglu, Aijun Teng, Zhi Chang, Peng Yuan, Dongbin Zhang, and Jiyang Yu. 2022. "Investigations on Structural, Electronic and Optical Properties of MoS2/CDs Heterostructure via First-Principles Study" Catalysts 12, no. 5: 456. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050456

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