Next Article in Journal
High-Speed Centrifugal Spinning Polymer Slip Mechanism and PEO/PVA Composite Fiber Preparation
Previous Article in Journal
Enhancement Mechanism of Pt/Pd-Based Catalysts for Oxygen Reduction Reaction
Previous Article in Special Issue
Effect of Nano Nd2O3 on the Microstructure and High-Temperature Resistance of G@Ni Laser Alloying Coatings on Ti-6Al-4V Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Optical Properties of Tungsten Disulfide Nanoscale Films Grown by Sulfurization from W and WO3

1
Department of Physics, National Cheng Kung University, Tainan 701, Taiwan
2
Department of Applied Physics, National University of Kaohsiung, Kaohsiung 811, Taiwan
3
Taiwan Consortium of Emergent Crystalline Materials, Ministry of Science and Technology, Taipei 10601, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(7), 1276; https://doi.org/10.3390/nano13071276
Submission received: 20 February 2023 / Revised: 28 March 2023 / Accepted: 31 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue Processing, Surfaces and Interfaces of Nanomaterials)

Abstract

:
Tungsten disulfide (WS2) was prepared from W metal and WO3 by ion beam sputtering and sulfurization in a different number of layers, including monolayer, bilayer, six-layer, and nine-layer. To obtain better crystallinity, the nine-layer of WS2 was also prepared from W metal and sulfurized in a furnace at different temperatures (800, 850, 900, and 950 °C). X-ray diffraction revealed that WS2 has a 2-H crystal structure and the crystallinity improved with increasing sulfurization temperature, while the crystallinity of WS2 sulfurized from WO3 (WS2-WO3) is better than that sulfurized from W-metal (WS2-W). Raman spectra show that the full-width at half maximum (FWHM) of WS2-WO3 is narrower than that of WS2-W. We demonstrate that high-quality monocrystalline WS2 thin films can be prepared at wafer scale by sulfurization of WO3. The photoluminescence of the WS2 monolayer is strongly enhanced and centered at 1.98 eV. The transmittance of the WS2 monolayer exceeds 80%, and the measured band gap is 1.9 eV, as shown by ultraviolet-visible-infrared spectroscopy.

1. Introduction

Tungsten disulfide (WS2) is a two-dimensional (2-D) material consisting of a covalently bonded sheet of W atoms filled between two trigonal sheets of S atoms. WS2 is mainly composed of three structures, hexagonal (2-H), trigonal (1T), or rhombohedral (3R) phase. Among them, the 2H phase structure is relatively stable and exhibits better optical properties [1]. The interlayers of WS2 are bound only by weak van der Waals forces, and the interlayer spacing is ~0.6 nm [2,3].
WS2 has been extensively studied due to its unique layer-dependent properties, such as the ability to absorb 5–10% of incident sunlight [4], and its unique band structure. Monolayer WS2 exhibits a direct band gap [5] of 1.98 eV [6], while multilayer WS2 shows an indirect band gap [7] of about 1.3 eV [8,9]. This phenomenon can be attributed to the lack of Coulomb repulsion between the pz orbitals of the chalcogenide elements in adjacent layers, leading to stabilization of the Γ-state valence band [10]. WS2 has been investigated for applications such as solar cells [11], hydrogen evolution reactions (HER) [12], electrocatalysis [13], batteries [14], and transistors [15]. However, the efficiency of most of the devices remains low. Therefore, a better understanding of the formation and physical properties of WS2 is essential. Many efforts have been made to improve the application value of WS2, such as controlling the formation mechanism, including studying the growth mode of the films to obtain a uniform, large-area [16], and well-oriented crystals. There are three common methods for depositing WS2 thin films: the stripping method [17], the chemical vapor deposition (CVD) [18,19], and the direct vulcanization synthesis method [20]. In this study, we seek to obtain high-quality WS2 films by comparing tungsten (W) metal and oxide (WO3) sulfurization processes. The structural and optical properties of the WS2 thin films were investigated in detail for future applications in next-generation optoelectronic devices.

2. Experimental

The W and WO3 films were prepared on c-axis Al2O3 (1 cm2, provided by Bangjie Material Technology). First, the Al2O3 substrate was ultrasonicated in acetone for 5 min to remove dirt and then rinsed with methanol. The cleaned c-axis sapphire was then inserted into the ion beam sputtering (IBS, Commonwealth Scientific) chamber for the growth of W or WO3 films. Subsequently, the grown W or WO3 films were sulfurized to obtain WS2 films, as schematically shown in Figure S1a,b of the Supplementary Document. The sputtering of W films was performed at a base pressure of about 5 × 10−6 Torr. Ar (99.999% purity) was introduced into the chamber at a flow rate of 5 sccm to initiate sputtering process. On the other hand, the WO3 films were grown by flowing Ar and oxygen at flow rates of 5 and 3 sccm, respectively. The sulfurization of the W and WO3 films was carried out in a horizontal quartz-tube furnace. The W or WO3 films were placed in the center of the quartz-tube furnace while about 3 g of sulfur (99.999% purity) was placed next to the tube lid. The tube was then evacuated to 5 × 10−2 Torr. During the sulfurization process, nitrogen gas (99.999%) flowed into the chamber while maintaining a pressure of 0.7 Torr. In this work, the sulfurization process is maintained at 900 °C for thickness-dependent studies. To study the crystallinity behavior, the nine-layer sample of WS2-W was sulfurized at different temperatures of 800, 850, 900, and 950 °C. All sulfurization processes were carried out for 20–30 min and cooled naturally to room temperature. The elemental compositions of the WS2 films were examined using X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific with an Al Kα light source). The binding energies were referenced to the NIST-XPS database. The WS2 phase was identified by micro-Raman (Horiba Jobin Yvon Lab RAM HR) equipped with an 1800-cycle grating, an objective lens with 100× magnification, and a laser wavelength of 532 nm. X-ray absorption near-edge spectroscopy (XANES) was performed to identify the local electron structure of WS2. The XANES was performed at the 07A1 beamline of the National Synchrotron Radiation Research Center (Hsinchu, Taiwan). The crystal structure of WS2 films was examined by X-ray diffraction (XRD, D8 by Bruker AXS Gmbh) equipped with a Cu 2 keV light source. A high-resolution transmission electron microscope (HR-TEM, JEOL JEM-2100F CS STEM) equipped with a dual-beam focus ion beam was conducted to study the microstructure and thickness at the atomic scale. The electrons were accelerated at 12 kV and a magnification of 500 K. The TEM was also equipped with an energy-dispersive X-ray spectrometer (EDS). Further analysis was performed using micro-photoluminescence spectroscopy (PL) to study the band gap of WS2 and ultraviolet-visible-infrared spectroscopy to understand the light absorption properties of the samples.

3. Results and Discussion

3.1. Structural Analysis

Figure 1a,b shows the XRD of WS2-W at different sputtering thicknesses and sulfurization temperatures. All the XRD peaks marked with an asterisk match the substrate Al2O3 peaks. The presence of lattice planes (002), (004), (006), and (008) suggests that the structure of WS2 is a 2H (hexagonal) phase [5]. Moreover, the interlayer spacing of the WS2 was estimated to be ~0.62 nm using the Bragg equation, which agrees with the theory and previous results [21,22].
With increasing film thickness, the intensity of XRD signals (002), (004), (006), and (008) increases, while the FWHM of each peak is reduced, as shown in Figure 1a, indicating that the crystallinity of the films improves with the number of WS2 layers. The absence of the (002) peak for the monolayer of WS2 is due to the diffraction limit. The tungsten films were sulfurized at different temperatures of 800, 850, 900, and 950 °C to obtain the optimum temperature for formation WS2 (nine-layer sample). The corresponding XRD results are shown in Figure 1b, indicating that the crystallinity of WS2 improves with increasing sulfurization temperature. Additionally, to study the local structure of WS2, the X-ray absorption near-edge structure (XANES) was also performed at W L3-edge. Figure 1c exhibits the spectrum of WS2 with tungsten and tungsten trioxide as references. The XANES of WS2 agrees with the previous result [23]. In addition, it also confirms that the WS2 (black path) has a 2H (hexagonal) structure.
TEM investigation has been carried out to observe the atomic stacking of WS2 (nine-layer sample). In order to improve the conductivity of the film, about 5 nm of platinum was deposited on the surface of WS2. Platinum was also used to protect the sample from oxidation. Figure 1c shows the TEM image of the nine-layer of WS2 stacked along the c-axis with the platinum capping layer. The total thickness of the WS2 layer was estimated to be ~5.37 nm with an interplanar spacing of 0.62 nm (Figure 1d), which agrees with the XRD result.

3.2. Elemental Composition Analysis

The elemental composition of WS2 was qualitatively examined by X-ray photoelectron spectroscopy (XPS). Figure 2 shows high-resolution XPS spectra in W4+ and S2− energy regions. Figure 2a exhibits the tungsten signals at 32.6, 34.8, and 38.2 eV, which can be ascribed to W 4f7/2, W 4f5/2, and W 5f3/2, respectively, with a spin-orbit splitting of Δ E P   (4f7/2 − 4f5/2) = 2.2 eV. The existence of high-intensity signals at 32.6 and 34.8 eV also suggests the formation of 2H-WS2, which concurs with previous reports [24,25]. Figure 2b shows the spectrum of the doublet S 2p at 163.5 eV and 162.3 eV, representing S 2p1/2 and S 2p3/2, respectively [26,27]. These peaks are attributed to the divalent sulfide ions (S2−), which are also associated with the formation of 2H-WS2. The elemental composition of WS2 was estimated to be 56.0, 28.7, and 15.3% (in atomic percentage) for S 2p, W 4f, and O 1s, in order. The oxygen signal is from the sapphire substrate. Therefore, the W and S ratio is very close to 2:1, confirming that the samples are in good stoichiometry. Additionally, the elemental composition of WS2 was also qualitatively examined using EDS (Figure S2).

3.3. Optical Properties

3.3.1. Micro-PL Spectroscopy

Micro-PL has been performed to study the optical and electronic properties of WS2. It is found that the PL signal comes from the energy gap of the direct transition when the thickness is reduced below two atomic layers. The PL of WS2 is strongly suppressed as thickness increases above two atomic layers, in which the indirect band gap is completely formed.
Figure 3 shows the PL peaks of the monolayer and bilayer of WS2. The peaks are centered at 1.98 and 1.95 eV, respectively, in good agreement with previous results (2.1–1.9 eV) [28] and with the DFT-LDA calculations for the monolayer and the bilayer of WS2 [29,30]. The PL intensity shows a slight redshift (~3 eV) and a threefold decrease in intensity as the layer increases from the monolayer to the bilayer of WS2 [3]. The peak shift is likely due to the proportional relationship between layer number and carrier concentration, which also contributes to Coulomb scattering [31]. This result is in line with the theoretical prediction that the WS2 monolayer has a direct band gap at the Γ point [32].

3.3.2. Ultraviolet-Visible-Infrared Spectroscopy

The optical properties of WS2 were further investigated using an ultraviolet-visible-infrared (UV/visible/NIR) spectrophotometer, as shown in Figure 4. It was observed that the transmittance of WS2 increased from 40% (nine-layer) to over 80% (monolayer) in the UV-visible region (Figure 4a). The band gap of the WS2 monolayer is about 1.9 eV, as estimated from the Tauc plot of the absorption spectrum (Figure 4b).

3.4. Comparison of WS2-W and WS2-WO3

3.4.1. XRD of WS2-W and WS2-WO3

In order to obtain high-quality WS2 films, we compared the sulfurization of W and WO3 films, termed as WS2-W and WS2-WO3, respectively. The samples were sulfurized side by side at 900 °C. The XRD results are shown in Figure 5. It was revealed that the diffraction peaks of WS2-WO3 are sharper and narrower than those of WS2-W with similar thickness, suggesting that the crystallinity of WS2-WO3 is better than that of WS2-W.

3.4.2. Raman Spectra of WS2-W and WS2-WO3

Raman spectroscopy is widely used for elemental analysis, layer stacking order, and doping effect of transition metal dichalcogenides [33]. Figure 6 shows the Raman spectra of WS2-W and WS2-WO3 with the different number of stacked layers.
Figure 6a,b shows two main Raman peaks centered at ~354 and ~420 cm−1, respectively. These Raman peaks agree with previous work [21,34,35]. It was observed that the A1g peaks overlap with the Raman peaks of sapphire (380 and 417cm−1) [36] when the Raman peaks were deconvoluted using a multi-peak Lorentzian fitting method [37]. Deconvolution of Raman peaks also revealed three strong peaks, which are E 2 g 1 (in-plane displacement of W and S), A1g (out-plane displacement of S-S atoms), and 2LA (M) (second order Raman peak), which are centered at ~356, ~418, and ~351 cm−1, respectively. Furthermore, the approximate distance between A1g and E 2 g 1 of WS2-W is ~61.8, 63.8, 61.6, and 64.1 cm−1 for monolayer, bilayer, six-layer, and nine-layer samples, respectively, which is not much different from the distance for WS2-WO3 of about 61.8, 63.7, 63.2, 62.7 in the same order. Figure 6c shows the I 2 L A / I A 1 g intensity ratio of WS2-W reduces from the monolayer (1.8) to the nine-layer (1.3). However, WS2-WO3 shows a slightly different trend, where the intensity ratio increases from the monolayer (1.3) to the bilayer (1.8), and then decreases to the nine-layer (1.3). The results suggest that the second-order Raman peak intensity ( I 2 L A ) is almost twice that of the first-order   I A 1 g . This is consistent with the previous results [38,39]. Additionally, for both WS2-W and WS2-WO3 samples, as the film thickness decreases, the A1g signals show a slight red shift of about ~2.3 cm−1 (Figure 6d). The redshift of A1g is likely associated with reducing the atomic restoring forces. When the number of WS2 layers is decreased, the long-range Coulomb interaction between effective charges and dielectric screening is enhanced, leading to increased restoring force between the S-S atoms [3].
Due to the dielectric screening effect, the Raman signal is sensitive to the interactions between atoms in the interlayer and long-range Coulomb interactions [40]. The Raman wavevector and the intensity of WS2 are strongly correlated with the layer thickness at the nanoscale level. In this work, we also used Raman spectra to calibrate the thickness of WS2 as a monolayer, bilayer, six-layer, and nine-layer [31,39]. The Raman spectra confirmed that WS2 was well-formed as WS2-W and WS2-WO3. It is clear that the peak intensity and FWHM of E 2 g 1 for WS2-WO3 are narrower than that of WS2-W (Table S1 in the Supplementary Materials). The average estimation of FWHM of WS2-WO3 is ~7.2; however, the WS2-W is ~10.1, suggesting WS2-WO3 has better crystallinity than WS2-W. This is consistent with the XRD results (Figure 5). This phenomenon can be attributed to the fact that sulfur can replace oxygen more effectively at a relatively lower temperature [41], while WS2 formed from tungsten metal requires a higher temperature to facilitate metal–sulfur bonding [42]. As a result, the defect density in WS2-WO3 is likely to be lower than in WS2-W.

4. Conclusions

We report a study of nanoscale WS2 films with various number of layers, prepared by sulfurization of W and WO3 at different temperatures (800 to 950 °C). The WS2 films were inferred to be the 2H phase and c-axis oriented. The XRD shows that the crystal quality of the WS2 films improved with increasing sulfurization temperature. The photoluminescence of the monolayer of WS2 is strongly enhanced and centered at 1.98 eV. The transmittance of the monolayer WS2 exceeds 80% and the band gap is 1.9 eV, revealed by ultraviolet-visible-infrared spectroscopy. The Raman analysis shows that the FWHM of WS2-WO3 is narrower than that of WS2-W, indicating the structure of WS2-WO3 is superior to that of WS2-W, which is in good agreement with the X-ray diffraction result. We conclude that a large-area, high-quality WS2 film can be prepared by sulfurizing WO3. The results are promising for applications in next-generation optoelectronic devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13071276/s1, Figure S1: Schematic diagram: (a). W-metal and WO3 prepared on sapphire substrate by ion beam sputtering technique; (b). W-metal or WO3 sulfurization process using a thermocouple-equipped furnace, the process was carried out inside a horizontal quartz tube with a diameter of 50 mm and length of 100 cm; Figure S2: shows the EDS depth analysis of the WS2 film that was carried out with Line scan mode. The small picture in the upper left corner is the relationship between the scanning position and the ratio of each element; Figure S3: the deconvolution of Raman spectra of bilayer, six-layer, and nine-layer samples using a multi-peak Lorentz fitting to separate the A1g peak from the substrate and 2LA from E_2 g1. The (a–c) shows the WS2 sulfurized from tungsten metal and (d–f) shows the WS2 sulfurized from tungsten trioxide; Table S1: Summary of the intensity ratio of I 2 L A / I A 1 g and FWHM as a function of layer numbers.

Author Contributions

J.-Y.C. and T.-T.H.; data curation, P.G.; Investigation and writing-original draft, J.-C.L. and S.-H.S.; Visualization, J.-C.A.H.; writing-review and editing. All authors discussed the results. All authors have read and agreed to the published version of the manuscript.

Funding

We would like to acknowledge the financial support from National Science and Technology Council supported this research financially under contracts NSTC 111-2124-M-006-008 and 109-2112-M-006-019-MY3.

Data Availability Statement

All data included in this study are available upon request by contact with the corresponding author.

Acknowledgments

The authors gratefully acknowledge the use of XRD 003100, SQUID 000200 and EM 012300 of MOST 110-2731-M-006-001 belonging to the Core Facility Center of National Cheng Kung University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahler, B.; Hoepfner, V.; Liao, K.; Ozin, G.A. Colloidal synthesis of 1T-WS2 and 2H-WS2 nanosheets: Applications for pho-tocatalytic hydrogen evolution. J. Am. Chem. Soc. 2014, 136, 14121–14127. [Google Scholar] [CrossRef] [PubMed]
  2. Li, W.; Wang, T.; Dai, X.; Wang, X.; Zhai, C.; Ma, Y.; Chang, S. Bandgap engineering different stacking WS2 bilayer under an external electric field. Solid State Commun. 2016, 225, 32–37. [Google Scholar] [CrossRef]
  3. Gutierrez, H.R.; Nestor, P.L.; Ana, L.E.; Ayse, B.; Bei, W.; Ruitao, L.; Florentino, L.U.; Vincent, H.C.; Humberto, T.; Mauricio, T. Extraordinary room-temperature photoluminescence in triangular WS2 monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bernardi, M.; Palummo, M.; Grossman, J.C. Extraordinary Sunlight Absorption and One Nanometer Thick Photovoltaics Using Two-Dimensional Monolayer Materials. Nano Lett. 2013, 13, 3664–3670. [Google Scholar] [CrossRef]
  5. Jo, S.H.; Ubrig, N.; Berger, H.; Kuzmenko, A.B.; Morpurgo, A.F. Mono- and bilayer WS2 light-emitting transistors. Nano Lett. 2014, 14, 2019–2025. [Google Scholar] [CrossRef] [Green Version]
  6. Liang, L.; Meunier, V. First-principles Raman spectra of MoS2, WS2 and their heterostructures. Nanoscale 2014, 21, 5394–5401. [Google Scholar] [CrossRef]
  7. Zibouche, N.; Kuc, A.; Heine, T. From layers to nanotubes: Transition metal disulfides TMS2. Eur. Phys. J. B 2012, 85, 49. [Google Scholar] [CrossRef] [Green Version]
  8. Kuc, A; Zibouche, N; Heine, T. Influence of quantum confinement on the electronic structure of the transition metal sulfide TS2. Phys. Rev. B 2011, 83, 245213. [Google Scholar] [CrossRef] [Green Version]
  9. Frey, G.L.; Tenne, R.; Matthews, M.J. Optical properties of MS2 (M = Mo, W) inorganic fullerene like and nanotube material optical absorption and resonance Raman measurements. J. Mater. Res. 1998, 13, 9. [Google Scholar] [CrossRef] [Green Version]
  10. Ghosh, S.; Brüser, V.; Kaplan, A.I.; Popovitz, B.R.; Peglow, S.; Martínez, J.I.; Alonso, J.A.; Zak, A. Cathodoluminescence in single and multiwall WS2 nanotubes: Evidence for quantum confinement and strain effect. Appl. Phys. Rev. 2020, 7, 041401. [Google Scholar] [CrossRef]
  11. Shanmugam, M.; Durcan, C.A.; Jacobs-Gedrim, R.; Yu, B. Layered semiconductor tungsten disulfide: Photoactive material in bulk heterojunction solar cells. Nano Energy 2013, 2, 419–424. [Google Scholar] [CrossRef]
  12. Cheng, L.; Huang, W.; Gong, Q.; Liu, C.; Liu, Z.; Li, Y.; Dai, H. Ultrathin WS2 Nanoflakes as a High-Performance Electrocatalyst forth Hydrogen Evolution Reaction. Angew. Chem. Int. Ed 2014, 53, 7860–7863. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, T.Y.; Chang, Y.H.; Hsu, C.L.; Wei, K.H.; Chiang, C.Y.; Li, L.J. Comparative study on MoS2 and WS2 for electro-catalytic water splitting. Hydrog. Energy Publ. 2013, 38, 12302–12309. [Google Scholar] [CrossRef]
  14. Feng, C.; Huang, L.; Guo, Z.; Liu, H. Synthesis of tungsten disulfide (WS2) nanoflakes for lithium ion battery application. Electrochem. Commun. 2007, 9, 119. [Google Scholar] [CrossRef]
  15. Huo, N.; Kang, J.; Wei, Z.; Li, S.S.; Li, J.; Wei, S.H. Novel and Enhanced Optoelectronic Performances of Multilayer MoS 2 –WS 2 Heterostructure Transistors. Adv. Funct. Mater 2014, 24, 7025–7031. [Google Scholar] [CrossRef]
  16. Elías, A.L.; Perea-López, N.; Castro-Beltrán, A.; Berkdemir, A.; Lv, R.; Feng, S.; Long, A.D.; Hayashi, T.; Kim, Y.A.; Endo, M.; et al. Controlled synthesis and transfer of large-area WS2 sheets: From single layer to few layers. ACS Nano 2013, 7, 5235–5242. [Google Scholar] [CrossRef]
  17. Miremadi, B.K.; Morrison, S.R. The intercalation and exfoliation of tungsten disulfide. J. Appl. Phys. 1988, 63, 4970–4974. [Google Scholar] [CrossRef]
  18. Cong, C.; Shang, J.; Wu, X.; Cao, B.; Peimyoo, N.; Qiu, C.; Sun, L.; Yu, T. Synthesis and optical properties of large-area sin-gle-crystalline 2D semiconductor WS2monolayer from chemical vapor deposition. Adv. Opt. Mater 2013, 2, 131–136. [Google Scholar] [CrossRef]
  19. Modtland, B.J.; Efren, N.M.; Xiang, J.; Marc, B.; Jing, K. Monolayer tungsten disulfide (WS2) via chlorine-driven chemical vapor transport. Small 2017, 1701232. [Google Scholar] [CrossRef]
  20. Wan, L.; Shi, C.W.; Yu, Z.B.; Wu, H.D.; Xiao, W.; Geng, Z.X.; Ren, T.Q.; Han, Q.; Yang, Z.X. Preparation of WS2/C composite material and its electrocatalytic hydrogen evolution performance. J. Fuel Chem. Technol. 2021, 49, 1362–1370. [Google Scholar] [CrossRef]
  21. Ramakrishna Matte, H.S.S.; Gomathi, A.; Manna, A.K.; Late, D.J.; Datta, R.; Pati, S.K.; Rao, C.N.R. MoS2 and WS2 Analogues of graphene. Angew. Chem. Int. Ed 2010, 49, 4059–4062. [Google Scholar] [CrossRef] [PubMed]
  22. Schutte, W.J.; Deboer, J.L.; Jellinek, F. Crystal-Structures of Tungsten Disulfide and Diselenide. J. Solid State Chem. 1987, 70, 207–209. [Google Scholar] [CrossRef]
  23. Prouzet, E.; Heising, J.; Kanatzidis, M.G. Structure of Restacked and Pillared WS2: An X-ray Absorption Study. Chem. Mater. 2003, 15, 412–418. [Google Scholar] [CrossRef]
  24. Wang, J.; Yu, L.; Zhou, Z.; Zeng, L.; Wei, M. Template-free synthesis of metallic WS2 hollow microspheres as an anode for the sodium-ion battery. J. Colloid Interface Sci. 2013, 557, 722–728. [Google Scholar] [CrossRef] [PubMed]
  25. Cao, H.; Wen, F.; De Hosson, J.T.M.; Pei, Y.T. Instant WS2 platelets reorientation of self-adaptive WS2/a-C tribocoating. Mater. Lett. 2018, 229, 64–67. [Google Scholar] [CrossRef]
  26. Yang, W.; Wang, J.; Si, C.; Peng, Z.; Frenzel, J.; Eggeler, G.; Zhang, Z. [001] Preferentially-oriented 2D tungsten disulfide nanosheets as anode materials for superior lithium storage. J. Mater. Chem A 2015, 3, 17811. [Google Scholar] [CrossRef]
  27. Xu, S.; Gao, X.; Hu, M.; Sun, J.; Wang, D.; Zhou, F.; Weng, L.; Liu, W. Morphology Evolution of Ag Alloyed WS2 Films and the Significantly Enhanced Mechanical and Tribological Properties. Surf. Coat. Technol. 2014, 238, 197–206. [Google Scholar] [CrossRef]
  28. Frey, G.L.; Elani, S.; Homyonfer, M.; Feldman, Y.; Tenne, R. Optical-absorption spectra of inorganic fullerene like MS2 (M = Mo, W). Phys. Rev. B 1998, 57, 6666–6671. [Google Scholar] [CrossRef]
  29. Beal, A.R.; Liang, W.Y. Excitons in 2H-WSe2 and 3R-WS2. J. Phys. C Solid State Phys. 1976, 9, 2459–2466. [Google Scholar] [CrossRef]
  30. Berkdemir, A.; Gutiérrez, H.R.; Botello-Méndez, A.R.; Perea-López, N.; Elías, A.L.; Chia, C.I.; Wang, B.; Crespi, V.H.; López-Urías, F.; Charlier, J.C.; et al. Identification of individual and few layers of WS2 using Raman Spectroscopy. Sci. Rep. 2013, 3, 1755. [Google Scholar] [CrossRef] [Green Version]
  31. Zhao, W.; Ghorannevis, Z.; Amara, K.K.; Pang, J.R.; Toh, M.; Zhang, X.; Kloc, C.; Tan, P.H.; Eda, G. Lattice dynamics in mono- and few-layer sheets of WS2 and WSe2. Nanoscale 2013, 5, 9677. [Google Scholar] [CrossRef] [Green Version]
  32. Tongay, S.; Zhou, J.; Ataca, C.; Lo, K.; Matthews, T.S.; Li, J.; Grossman, J.C.; Wu, J. Thermally Driven Crossover from Indirect toward Direct Bandgap in 2D Semiconductors: MoSe2 versus MoS2. Nano Lett. 2012, 12, 5576–5580. [Google Scholar] [CrossRef] [PubMed]
  33. Lui, C.H.; Li, Z.; Chen, Z.; Klimov, P.V.; Brus, L.E.; Heinz, T.F. Imaging stacking order in few-layer graphene. Nano Lett. 2011, 11, 164–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Stacy, A.M.; Hodul, D.T. Raman-Spectra of Ivb and Vib Transition-Metal Disulfides Using Laser Energies near the Absorption Edges. J. Phys. Chem. Solids 1985, 46, 405–409. [Google Scholar] [CrossRef]
  35. Molina-Sanchez, A.; Wirtz, L. Phonons in single-layer and few-layer MoS2 and WS2. Phys. Rev. B 2011, 84, 155413. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, Y.; Zhang, Y.; Ji, Q.; Ju, J.; Yuan, H.; Shi, J.; Gao, T.; Ma, D.; Liu, M.; Chen, Y.; et al. Controlled Growth of High-Quality Monolayer WS2 Layers on Sapphire and Imaging Its Grain Boundary. ACS Nano 2013, 7, 8963–8971. [Google Scholar] [CrossRef] [PubMed]
  37. Huang, X.; Gao, Y.; Yang, T.; Ren, W.; Cheng, H.M.; Lai, T. Quantitative Analysis of Temperature Dependence of Raman shift of monolayer WS2. Sci. Rep. 2016, 6, 32236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Song, J.G.; Park, J.; Lee, W.; Choi, T.; Jung, H.; Lee, C.W.; Hwang, S.H.; Myoung, J.M.; Jung, J.H.; Kim, S.H.; et al. Layer-Controlled, Wafer-Scale, and Conformal Synthesis of Tungsten Disulfide Nanosheets Using Atomic Layer Deposition. ACS Nano 2013, 7, 11333–11340. [Google Scholar] [CrossRef]
  39. Sekine, T.; Nakashizu, T.; Toyoda, K.; Uchinokura, K.; Matsuura, E. Raman scattering in layered compound 2H-WS2. Solid State Commun. 1980, 35, 371–373. [Google Scholar] [CrossRef]
  40. Sie, E.J.; Steinhoff, A.; Gies, C.; Lui, C.H.; Ma, Q.; Rosner, M.; Schönhoff, G.; Jahnke, F.; Wehling, T.O.; Lee, Y.H.; et al. Observation of exciton redshift–blueshift crossover in monolayer WS2. Nano Lett. 2017, 14, 4210–4216. [Google Scholar] [CrossRef] [Green Version]
  41. Cao, S.; Zhao, C.; Han, T.; Peng, L. The WO3/WS2 nanostructures: Preparation, characterization and optical absorption properties. Physical E 2016, 81, 235–239. [Google Scholar] [CrossRef]
  42. Sun, Y.; Darling, A.J.; Li, Y.; Fujisawa, K.; Holder, C.F.; Liu, H.; Janik, M.J.; Terrones, M.; Schaak, R.E. De-fect-mediated selective hydrogenation of nitroarenes on nanostructured WS2. Chem. Sci. 2019, 10, 10310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Structure and morphology of WS2. The XRD spectra of WS2 films for (a) different numbers of layers; (b) different sulfurization temperatures of nine-layer of WS2; (c) the morphology of the WS2 is presented by TEM image; (d) the HR-TEM shows the interlayer distance of WS2 stacking layer; (e) the XANES shows the local structure of WS2.
Figure 1. Structure and morphology of WS2. The XRD spectra of WS2 films for (a) different numbers of layers; (b) different sulfurization temperatures of nine-layer of WS2; (c) the morphology of the WS2 is presented by TEM image; (d) the HR-TEM shows the interlayer distance of WS2 stacking layer; (e) the XANES shows the local structure of WS2.
Nanomaterials 13 01276 g001
Figure 2. High-resolution XPS of WS2 films: (a) W 4f and W 5f signals; (b) S 2p.
Figure 2. High-resolution XPS of WS2 films: (a) W 4f and W 5f signals; (b) S 2p.
Nanomaterials 13 01276 g002
Figure 3. Photoluminescence spectra of WS2 monolayer and bilayer.
Figure 3. Photoluminescence spectra of WS2 monolayer and bilayer.
Nanomaterials 13 01276 g003
Figure 4. (a) Comparison of the transmittance of the WS2 thin film with different thicknesses; (b) Tauc diagram of monolayer WS2 from the absorption. Inset: the absorption measurement can be used to determine the band gap of WS2.
Figure 4. (a) Comparison of the transmittance of the WS2 thin film with different thicknesses; (b) Tauc diagram of monolayer WS2 from the absorption. Inset: the absorption measurement can be used to determine the band gap of WS2.
Nanomaterials 13 01276 g004
Figure 5. XRD of nine-layer WS2 sulfurized from W and WO3 at 900 °C It shows narrower FWHMs of 002, 004, and 008 signals from WS2-WO3 (2.1 ± 0.1; 3.4 ± 0.2; 2.6 ± 0.2, respectively) then from WS2-W (2.2 ± 0.1; 4.1 ± 0.2; 3.1 ± 0.2, respectively).
Figure 5. XRD of nine-layer WS2 sulfurized from W and WO3 at 900 °C It shows narrower FWHMs of 002, 004, and 008 signals from WS2-WO3 (2.1 ± 0.1; 3.4 ± 0.2; 2.6 ± 0.2, respectively) then from WS2-W (2.2 ± 0.1; 4.1 ± 0.2; 3.1 ± 0.2, respectively).
Nanomaterials 13 01276 g005
Figure 6. (a,b). Deconvolution of the Raman spectra of WS2-W and WS2-WO3, respectively, using Lorentz fitting method; (c) the intensity ratio of I 2 L A / I A 1 g as a function of the number layers of WS2; (d) the peak positions of A1g, E 2 g 1 , and 2LA with respect to the number layer of WS2. Additionally, the deconvolution of the Raman spectra for bilayer, six-layer, nine-layer can be seen in the Supplementary Documents.
Figure 6. (a,b). Deconvolution of the Raman spectra of WS2-W and WS2-WO3, respectively, using Lorentz fitting method; (c) the intensity ratio of I 2 L A / I A 1 g as a function of the number layers of WS2; (d) the peak positions of A1g, E 2 g 1 , and 2LA with respect to the number layer of WS2. Additionally, the deconvolution of the Raman spectra for bilayer, six-layer, nine-layer can be seen in the Supplementary Documents.
Nanomaterials 13 01276 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gultom, P.; Chiang, J.-Y.; Huang, T.-T.; Lee, J.-C.; Su, S.-H.; Huang, J.-C.A. Structural and Optical Properties of Tungsten Disulfide Nanoscale Films Grown by Sulfurization from W and WO3. Nanomaterials 2023, 13, 1276. https://doi.org/10.3390/nano13071276

AMA Style

Gultom P, Chiang J-Y, Huang T-T, Lee J-C, Su S-H, Huang J-CA. Structural and Optical Properties of Tungsten Disulfide Nanoscale Films Grown by Sulfurization from W and WO3. Nanomaterials. 2023; 13(7):1276. https://doi.org/10.3390/nano13071276

Chicago/Turabian Style

Gultom, Pangihutan, Jiang-Yan Chiang, Tzu-Tai Huang, Jung-Chuan Lee, Shu-Hsuan Su, and Jung-Chung Andrew Huang. 2023. "Structural and Optical Properties of Tungsten Disulfide Nanoscale Films Grown by Sulfurization from W and WO3" Nanomaterials 13, no. 7: 1276. https://doi.org/10.3390/nano13071276

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