Next Article in Journal
A Ferroelectric-Photovoltaic Effect in SbSI Nanowires
Next Article in Special Issue
Designing a Novel Monolayer β-CSe for High Performance Photovoltaic Device: An Isoelectronic Counterpart of Blue Phosphorene
Previous Article in Journal
Direct Spectroscopy for Probing the Critical Role of Partial Covalency in Oxygen Reduction Reaction for Cobalt-Manganese Spinel Oxides
Previous Article in Special Issue
Strain-Tunable Visible-Light-Responsive Photocatalytic Properties of Two-Dimensional CdS/g-C3N4: A Hybrid Density Functional Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile and Controllable Synthesis of Large-Area Monolayer WS2 Flakes Based on WO3 Precursor Drop-Casted Substrates by Chemical Vapor Deposition

1
Chongqing Key Lab of Multi-Scale Manufacturing Technology, Chongqing Institute of Green and Intelligent Technology, Chinese Academy of Sciences, Chongqing 400714, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 26 February 2019 / Revised: 29 March 2019 / Accepted: 2 April 2019 / Published: 9 April 2019

Abstract

:
Monolayer WS2 (Tungsten Disulfide) with a direct-energy gap and excellent photoluminescence quantum yield at room temperature shows potential applications in optoelectronics. However, controllable synthesis of large-area monolayer WS2 is still challenging because of the difficulty in controlling the interrelated growth parameters. Herein, we report a facile and controllable method for synthesis of large-area monolayer WS2 flakes by direct sulfurization of powdered WO3 (Tungsten Trioxide) drop-casted on SiO2/Si substrates in a one-end sealed quartz tube. The samples were thoroughly characterized by an optical microscope, atomic force microscope, transmission electron microscope, fluorescence microscope, photoluminescence spectrometer, and Raman spectrometer. The obtained results indicate that large triangular monolayer WS2 flakes with an edge length up to 250 to 370 μm and homogeneous crystallinity were readily synthesized within 5 min of growth. We demonstrate that the as-grown monolayer WS2 flakes show distinctly size-dependent fluorescence emission, which is mainly attributed to the heterogeneous release of intrinsic tensile strain after growth.

Graphical Abstract

1. Introduction

The isolation and synthesis of atomically thin two-dimensional transition metal dichalcogenides (TMDs), such as MS2 (M = Mo, W), have attracted a lot of interest due to their unprecedented properties compared with their bulk counterparts [1,2,3]. Unlike graphene and h-BN (h-Born Nitride), atomically thin MS2 possess semiconducting behavior with an intrinsic direct-energy gap corresponding to the visible frequency range, and show strong spin-orbit coupling and band splitting [4]. Of these, molybdenum disulfide (MoS2) has received tremendous attention because of its unique crystal structure, and physical, chemical, and electrical properties [5,6,7,8,9,10,11,12]. Due to these fascinating properties, MoS2 has been used in numerous applications, including photodetection, gas-sensing, and hydrogen evolution [13,14,15,16,17]. Atomically thin tungsten disulfide (WS2) has a similar structure to MoS2, but exhibits stronger photoluminescence (PL) quantum yield at room temperature and larger spin-orbit coupling compared to MoS2 [18,19,20], which makes it promising for applications in new optoelectronics and spintronics. Nevertheless, the research on WS2 lags far behind that of MoS2 and the development of simple and controllable strategies for preparing high-quality and large-area atomically thin WS2 is still a big challenge, which has impeded further fabrication of optoelectronics/spintronics devices.
As reported, large-area continuous mono- and multi-layer WS2 were synthesized by sulfurizing the pre-deposited ultrathin tungsten or WOx films obtained by various routes, including thermal evaporation, magnetron sputtering, and atomic layer deposition (ALD) [21,22,23,24,25,26,27]. This two-step chemical vapor deposition (CVD) method can provide large-area polycrystalline WS2 film for the fabrication of devices. However, the size of the single-crystal domain in the polycrystalline film is limited to the nanoscale and its electrical properties could be seriously degraded because of the existence of abundant grain boundaries (GBs) [28]. Moreover, the pre-deposition of ultrathin films is time-consuming and more efforts should be paid to precisely control the thickness, which restricts the wide use of two-step CVD in research communities. To simplify the synthesis process, one-step CVD that is based on direct sulfurization of powdered WO3 on diverse substrates, such as SiO2/Si, sapphire, Au foils, and h-BN, has been employed for the synthesis of WS2 [29,30,31,32,33,34]. In this direction, powdered sulfur and WO3 in quartz/ceramic boats are placed at separate locations; normally the sulfur is located upstream and the WO3 is placed underneath or ahead of the substrates. In addition, growth parameters, such as the temperature, heating rate, gas flow, and the distance between the precursor and substrate, can be accurately controlled between batches. However, this strategy presents some drawbacks since the distribution of the solid-phase precursors loaded in quartz/ceramic boats is not easy to control, and this could significantly influence the concentration of gaseous species on the growth interface. Furthermore, due to the difference in melting points of sulfur and WO3, it is difficult to simultaneously and precisely control the evaporation of these two precursors and the subsequent transportation to the growth interface. As a result, the dimension and morphology of the CVD-grown WS2 always present significant differences even under the same growth parameters.
Lee et al. [35] reported that continuous polycrystalline monolayer WS2 films in the centimeter-scale were grown on SiO2/Si substrate with the seeding of perylene-3,4,9,10-tetracarboxylic acid tetrapotassium salt (PTAS). However, the size of the isolated single-crystal domains was limited in the range of 10 to 20 μm. Zhang et al. [31] reported the synthesis of mono- and multi-layer WS2 flakes with the domain size exceeding 50 × 50 μm2 on the sapphire substrate at low-pressure with mixed hydrogen and argon gases. Yue et al. [36] reported that uniform triangular monolayer WS2 flakes with a side length of ~233 μm were synthesized on SiO2/Si substrate by carefully adjusting the introduction time of the sulfur precursor and the distance between the sources and substrates. Recently, Zhou et al. [37] demonstrated that molten-salt-assisted CVD can be used to synthesize large-area triangular monolayer WS2 flakes with an edge length of 300 μm at moderate temperatures due to the reduction of the WO3 melting point by using sodium chloride. To this end, significant attempts have been made towards the large-area synthesis of monolayer WS2 by one-step CVD. However, the layer controllability and universality of the synthesis method are the major existing obstacles that prevent practical applications. Therefore, controllable synthesis of large-area monolayer WS2 is still challenging.
In this study, we report a facile and controllable method for synthesis of large-area monolayer WS2 flakes by one-step CVD at atmospheric pressure. To promote the distribution of WO3, we first dispersed the powdered WO3 in ethanol to form a suspension solution, then we used a pipette to drop-cast the solution onto the SiO2/Si substrates. Moreover, we simultaneously loaded the powdered precursors and SiO2/Si substrates in a small one-end sealed quartz tube; and we placed the small tube inside a bigger quartz tube to ensure adequate amounts of sulfur-precursors to participate in the whole process of WS2 growth. Through this method, we obtained a series of triangular monolayer WS2 flakes with edge lengths up to 250 to 370 μm and homogeneous crystallinity. The morphology, thickness, atomic structure, and light emissions of the as-grown WS2 samples were characterized by various tools, such as an optical microscope (OM), atomic force microscope (AFM), transmission electron microscope (TEM), fluorescence (FL) microscope, PL spectrometer, and Raman spectrometer. It was found that the as-grown monolayer WS2 flakes with edge lengths greater than 95 μm show suppressed fluorescence emission in the inner region, while the smaller one presents homogeneous fluorescence emission. Raman mapping results indicate that the distinctly size-dependent fluorescence emission is attributed to the heterogeneous release of intrinsic tensile strain within WS2 after growth. Furthermore, the results of five independent experiments are consistent with each other, confirming the validity and reproducibility of this method.

2. Materials and Methods

In this research, the synthesis of WS2 was achieved by one-step CVD in a horizontal single-zone furnace (Hefei FACEROM Co., Ltd., Hefei, China) at atmospheric pressure, as shown in Figure 1a. The CVD system mainly consists of the heating zone and a quartz tube with a 60 mm diameter. Upstream of the quartz tube is connected to the high-purity (99.999%) argon cylinder, while downstream is connected to the exhausted gas treatment system. We did not use hydrogen in this research, due to considerations of safety, although previous reports demonstrated that hydrogen could facilitate the reduction of WO3 and lead to the growth of high quality WS2 [38,39]. Furthermore, a small quartz tube with a diameter of 15 mm sealed at one-end was intentionally employed for holding precursors and substrates simultaneously, this procedure was different from previous reports [36,40,41], aiming at obtaining large-area WS2 in a short period of time.
The SiO2 (300 nm)/Si (5 mm × 5 mm) sliced from a 4-inch wafer was used as substrate and was sequentially washed in DI (Deionized) water, acetone, DI water, ethanol, and DI water in an ultrasonic bath for 10 min each. The residual water and organic solvent were removed by compressed UHP (Ultra-high Purity) nitrogen gas blowing. Alcoholic tungsten suspension solution with a concentration of 1.5 mg/mL was prepared by WO3 (>99.9%, Adamas-beta, Shanghai, China) and ethanol (AR, KESHI, Chengdu, China). Before the growth process, a drop (10 μL) of alcoholic tungsten suspension solution was directly dropped onto the cleaned substrate using a pipette, as shown in Figure 1b, and the substrate was dried on a heater (40 °C) for 3 min. A relatively uniform distribution of powdered WO3 on the SiO2/Si substrate was obtained (as shown in Supplementary Materials Figure S1) after the evaporation of ethanol. Then, another fresh and cleaned SiO2/Si substrate was placed face-down above the drop-casted substrate to form a sandwich-structure. Subsequently, 600 mg of sulfur powder (S sublimed, >99.5%, KESHI, Chengdu, China) was introduced at the bottom of the sealed end of the small quartz tube, while the sandwich-structure substrates were carefully placed downstream, 24 cm away from the powdered sulfur. After that, the small quartz tube together with the precursors and substrates was loaded into the bigger quartz tube carefully, ensuring the substrates were exactly located at the center of the heating zone. Before heating, the furnace chamber was pumped to low pressure (<10 Pa) and then purged by 200 sccm argon to atmospheric pressure. Figure 1c shows the temperature profile used in this study. Initially, the furnace was heated from room temperature to 200 °C with a heating rate of 10 °C/min to remove the contaminants, such as water or residual organics. Subsequently, a higher heating rate of 28 °C/min was used to increase the temperature to 950 °C to obtain a high nucleation rate of WS2. The growth of WS2 was kept at 950 °C for 5 min under 200 sccm of argon. After growth, the furnace was cooled naturally to room temperature.
The morphology and size of the as-grown WS2 samples were characterized by an OM (50i POL, Nikon, Tokyo, Japan). The thickness and atomic structure of the as-grown WS2 were analyzed via an AFM (Dimension EDGE, Bruker, Billerica, MA, USA) in tapping mode and TEM (Tecnai G2 F20, FEI, Hillsboro, OR, USA), respectively. Raman spectra were collected in the backscattering geometry at room temperature with an excitation wavelength of 532 nm (InVia Reflex, Renishaw, Gloucestershire, UK). Before Raman characterization, the system was calibrated with the Raman peak of Si at 520 cm−1. Photoluminescence spectra were obtained from a home-made laser scanning confocal microscope system. A 532 nm CW laser with an average power of 100 μW was employed as the excitation source to avoid sample damage. Fluorescence images were captured by using a fluorescence microscope (BX53, Olympus, Tokyo, Japan) under excitation of a green light source.

3. Results and Discussion

Figure 2a shows a typical optical image of the as-grown WS2 sample on the covering SiO2/Si substrate, which is blank without WO3-ethanol drop-casting. The as-grown WS2 exhibits the feature of an equilateral triangle with sharp edges. Furthermore, the GBs can also be observed, which is caused by the coalescence of adjacent triangles and is hard to avoid in CVD-grown MoS2 and WS2 flakes [6,28]. It is noted that all the triangles regardless of their sizes show uniform contrast under OM, indicating good uniformity of the thickness of the as-grown WS2. Figure 2b presents a representative optical image of the WS2 sample grown on the bottom SiO2/Si substrate, which was drop-casted by WO3-ethanol solution. In contrast with the covering substrate, small triangular WS2 with uniform contrast can be occasionally observed, however, irregular-shape WS2 with heterogeneous contrast and particulate WS2 (as shown in Supplementary Materials Figure S2) are predominant on the bottom substrate. These distinctly morphological features could be caused by the existence of the powdered WO3 and highly concentrated reaction sources on the bottom substrate [33]. Due to the differences in the nucleation density and local environment on the substrate, the dimension distribution of the triangular flakes is quite wide, from tens to hundreds of micrometers, while large triangular WS2 flakes with edge lengths of ~270 μm are readily obtained on the covering SiO2/Si substrate, as shown in Figure 2a. In addition, truncated-triangle, hexagon, and butterfly-shaped flakes are formed on the covering SiO2/Si substrate as well (as shown in Supplementary Materials Figure S3), which was observed in previous CVD-grown WS2 and MoS2 and explained from the point of the changes in the M:S (M = Mo, W) ratio of the precursors within the growth interface [6,31,42,43].
In order to accurately determine the thickness of the as-grown WS2 sample, AFM was performed in tapping mode. Figure 2c,d show the AFM image and its corresponding height profiles of an equilateral triangular WS2 flake with uniform contrast under OM. It can be clearly observed that there are numerous small particles with several hundreds of nanometers in length and dozens of nanometers in height along the triangle edges. This phenomenon was mentioned by a previous report [37], and it is probably caused by the continuous attachment of the precursors to the growing edge. The thickness of the as-grown WS2 flake determined by the height profile is ~0.8 nm, which is consistent with the reported thickness of monolayer WS2 [36,44], corroborating the monolayer nature of the triangular WS2 flake with uniform contrast obtained under OM. To further identify the atomic structure of our sample, the triangular WS2 was transferred to a holy carbon-coated copper grid via the wet transfer method and analyzed by TEM. Figure 2e shows the high resolution TEM image of our WS2 sample, which clearly shows the hexagonal ring lattice consisting of alternating tungsten and sulfur atoms. Furthermore, the corresponding selective area electron diffraction (SAED) displayed in Figure 2f reveals only one set of diffraction spots, demonstrating the single-crystal nature of our WS2 sample. The inter-planar distances of (100) and (110) planes deduced from high resolution TEM measurement are ~0.269 nm and 0.155 nm, respectively, which coincide with the previous reported results of CVD-grown monolayer WS2 [40].
As a powerful and nondestructive tool, the Raman spectrometer has been widely employed to study the properties of TMDs, such as determination of the layer number [45,46,47], electrostatic doping [48], assessment of crystallinity [49,50], as well as internal and external strain [51,52]. Figure 3a shows a typical Raman spectrum of an equilateral triangular WS2 flake collected in the backscattering geometry at room temperature. As shown in Figure 3a, except the peak at 520 cm−1 from Si substrate, the other peaks are attributed to the WS2 flake. The strongest peak at ~350 cm−1 can be fitted well with three sub-peaks with a maximum frequency located at 344.7, 350.0, and 354.6 cm−1, and they can be identified as the first-order optical mode of   E 2 g 1 ( M ) , the second-order longitudinal acoustic mode of 2LA (M), and the first-order optical mode of E 2 g 1 ( Г ) (in-plane vibration between sulfur and tungsten atoms as shown in the inset of Figure 3a), respectively, according to the theoretical and experimental studies [45,53]. In addition, strong combination modes of 2 LA   ( M ) 2 E 2 g 2 ( Г ) and 2 LA   ( M ) E 2 g 2 ( Г ) were observed at 294.8 and 321.7 cm−1, respectively. The appearance of these strong combination modes is mainly attributed to the strong resonance between the phonon and B-exciton in WS2 excited by the 532 nm laser. The peak located at 416.8 cm−1 is assigned to the A 1 g ( Г ) mode, which is caused by the out-of-plane vibration between sulfur atoms and is sensitive to the layer number of WS2 [46]. Moreover, the intensity of 2LA (M) is much stronger than that of the A 1 g ( Г ) mode, giving rise to an intensity ratio ( R   =   I 2 LA   ( M ) / I   A 1 g ( Г ) ) of ~5.6, and the difference between the frequencies of the A 1 g ( Г ) and E 2 g 1 ( Г ) modes (Δ = A 1 g ( Г ) E 2 g 1 ( Г ) ) is ~62.2 cm−1. These characteristics are quite consistent with the reported results for monolayer WS2 excited by the 532 nm laser [33,36,41]. To probe the light emission and further determine the layer number of the triangular WS2 flake, the room temperature PL spectrum was collected with a 532 nm laser excitation. As shown in Figure 3b, a single and strong peak with a maxima wavelength of 632 nm (~1.96 eV) is observed, which is consistent with the reported PL peak position for monolayer WS2 [41,54,55], again confirming the monolayer nature of the as-grown WS2.
To investigate the light emission uniformity of the as-grown triangular WS2 flake, FL image was captured under excitation of a green light source. Figure 4a clearly shows that the outer region of the triangular WS2 flake with an edge length of ~270 μm exhibits intense FL emission, and it becomes weaker towards the inner region. In contrast, the two coalescent small triangles show relatively uniform FL emission across the entire region. This inhomogeneous FL emission in CVD-grown monolayer WS2 was observed by other groups. Peimyoo et al. [18] reported a similar observation of the suppression of FL emission for the center region as compared with the edge and they speculated that the suppression was relative to the structural imperfection and n-doping induced by charged defects. Liu et al. [56] observed inhomogeneous FL patterns consisting of alternating dark and bright concentric triangles in monolayer WS2, and they attributed the darker region to the high concentration of sulfur vacancies. Recently, Feng et al. [57] reported a novel FL aging behavior in monolayer WS2 with a large size, and they attributed that behavior to the partial release of intrinsic tensile strain after CVD growth.
To better understand the heterogeneity of FL emission within the large-area triangular WS2, Raman mapping with a step of 3 μm was performed and more than 10,000 data points across the entire flake were collected. As shown in Figure 4b, the distribution of the frequency difference between the A 1 g ( Г ) and E 2 g 1 ( Г ) modes is in the range of 61 to 62.5 cm−1, which is accordance with the reported results for monolayer WS2 [33,36,45], demonstrating that the entire triangular WS2 possesses a monolayer nature. It has been reported that the A 1 g ( Г ) mode is not only sensitive to the layer number, but also to the electrostatic doping in the typical 2H-type TMDs, and it will red-shift and broaden with electron doping [48]. As shown in Figure 4c,d, the frequency and full width of half maximum (FWHM) mappings of the A 1 g ( Г ) mode show relatively uniform features, confirming good uniformity of the crystallinity and electron doping for this large WS2 triangle. As a result, the heterogeneous FL emission observed in this WS2 triangle is not related to the n-doping. In addition, it has been reported that the frequency of the E 2 g 1 ( Г ) mode will blue-shift and the intensity of the A 1 g ( Г ) mode will decrease after release of the tensile strain in monolayer WS2 [51,57]. To investigate the stress distribution, the normalized intensity mapping of the A 1 g ( Г ) mode and frequency mapping of the E 2 g 1 ( Г ) mode for this WS2 triangle were analyzed, as shown in Figure 4e,f, respectively. It can be clearly seen that the intensity of the A 1 g ( Г ) mode decreases obviously and the frequency of the E 2 g 1 ( Г ) mode blue-shifts ~1 cm−1 for the inner region as compared with the outer region. This indicates that the tensile strain in the outer region is stronger than that of the inner region. By carefully comparing the FL image with the Raman mappings of the A 1 g ( Г ) mode intensity and E 2 g 1 ( Г ) mode frequency, it can be seen that they match each other well. Therefore, the heterogeneity of FL emission observed in the large WS2 triangle is mainly ascribed to the inhomogeneous tensile strain in the WS2.
Figure 5 illustrates the FL images of the as-grown WS2 flakes with different edge lengths. The FL image of the WS2 triangle with an edge length of ~171 μm displayed in Figure 5a possesses similar heterogeneity of the FL emission as the larger WS2 triangle observed in Figure 4a. Furthermore, it can be seen that the red large triangle is divided into three smaller and equilateral triangles by three suppressed lines, which is similar to the previous observation in a WS2 triangle with an edge size ~100 μm and disappears with aging time [57]. The optical image and Raman mappings of this WS2 triangle are shown in Supplementary Materials Figure S4. It can be seen that the Raman mapping results for this WS2 triangle present similar features to the larger one observed in Figure 4a, so that the heterogeneity of the FL emission observed in this WS2 triangle is ascribed to the inhomogeneous tensile strain as well. When the edge length of the WS2 triangle is reduced to ~95 μm, three lines with slightly weak FL emission can also be observed in the red triangle displayed in Figure 5b, whereas the FL emission shows better homogeneity across the entire triangle. Interestingly, the FL emission becomes more homogeneous and intense when the edge length of the WS2 triangle is less than ~32 μm, as shown in Figure 5c. Therefore, it is evident that the monolayer triangular WS2 flakes synthesized in this study present distinctly size-dependent FL emission.
Intrinsic tensile strain could be introduced in monolayer MS2 grown on SiO2/Si substrate during the fast cooling process from the high growth temperature to room temperature due to the higher thermal expansion coefficient of MS2 compared with that of silica substrate [57,58]. Feng et al. [57] reported that the intrinsic tensile strain could be partially released from the edge towards the center with the aging time, and consequently concentric FL patterns could be formed in large size (~100 μm) monolayer WS2 crystals after 72 h of aging. In this study, all the FL images were captured a week after synthesis, however, we did not detect the concentric FL patterns, probably because the release of intrinsic tensile strain was related to the geometry of the WS2 flake and the interaction between WS2 and the substrates.
To evaluate the validity and reproducibility of this method, we performed an extra four batches of experiments under the same conditions (950 °C for 5 min). Figure 6 shows the representative optical images of the as-grown WS2 flakes obtained from different bathes of experiments under the same conditions. We clearly noted that triangular WS2 flakes with an edge length of 250 to 370 μm were readily formed for each batch of the experiment, indicating that the distribution of the solid-sate precursor was controlled well by the drop-casting method. Recently, centimeter scale continuous WS2 films with GBs and triangular monolayer WS2 flakes hundreds of micrometers (~300 μm) in dimension were grown by direct sulfurization of powdered WO3 on SiO2/Si substrates, however, the growth time for previous studies was more than 10 min or even more [28,36,57]. In this study, triangular monolayer WS2 flakes with an edge length of 250 to 370 μm were effectively formed only within 5 min CVD growth. Therefore, the transportation of precursors could be limited and thus more precursors were able to participate in the synthesis of WS2 when a small one-end sealed quartz tube was employed for holding the precursors and substrates simultaneously.

4. Conclusions

We demonstrated a facile and controllable method for the synthesis of large-area monolayer WS2 flakes by direct sulfurization of powdered WO3 on SiO2/Si substrates in a horizontal single-zone CVD system. A series of monolayer WS2 flakes with an edge length of 250 to 370 μm on SiO2/Si substrates were achieved by five independent growth experiments. The morphology, thickness, atomic structure, and light emission of the as-grown WS2 samples were characterized using various tools. It was found that the as-grown monolayer WS2 flakes exhibit homogeneous crystallinity and size-dependent FL emission. For the WS2, triangles with edge lengths less than 95 μm exhibited uniform FL emission while that with larger edge lengths showed heterogeneous FL emission. This heterogeneity of FL emission in large monolayer WS2 flakes can be attributed to the heterogeneous release of intrinsic tensile strain after growth. We believe our method could offer an opportunity that opens a new window for the growth of other TMDCs with large areas.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2079-4991/9/4/578/s1, Figure S1: Optical image of powdered WO3 on SiO2/Si substrate after the evaporation of ethanol. Figure S2: Raman spectra of the WO3 powder (a), sulfur powder (b) and particulate WS2 (c) marked with white arrow in Figure S2(d). Representative optical image of the as-grown WS2 sample on the bottom SiO2/Si substrate that was drop-casted by WO3-ethanol solution (d). Figure S3: Optical images of as-grown WS2 flakes on SiO2/Si substrate with various morphologies: Truncated-triangle (a), Hexagon (b, Butterfly (c). Figure S4: Optical image of the WS2 flake with edge length of ~171 μm on SiO2/Si substrate (a). Raman mappings of the specific WS2 sample: frequency difference between A 1 g ( Г ) and E 2 g 1 ( Г ) modes (b); frequency (c), full width of half maximum (FWHM) (d) and normalized intensity (e) of A 1 g ( Г ) mode; frequency of E 2 g 1 ( Г ) mode (f).

Author Contributions

Conceptualization, B.S. and D.W.; Data curation, K.D.; Formal analysis, S.F. (Shaoxi Fang), S.F. (Shuanglong Feng) and H.Z.; Funding acquisition, D.Z.; Investigation, B.S. and D.Z.; Project administration, D.W.; Supervision, C.T. and D.W.; Writing—original draft, B.S.; Writing—review & editing, D.Z. and C.T. B.S. and D.Z. contributed equally to this work.

Funding

This work was supported by National Natural Science Foundation of China (Grant No. 61701474), Natural Science Foundation of Chongqing, China (Grant No. cstc2017jcyjAX0320) and West Light Foundation of CAS.

Acknowledgments

The authors would like to thank Yuzhi Li for PL measurements, Risheng Qiu for TEM measurements and Xilin Tan for polishing writing.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ji, Q.; Zhang, Y.; Zhang, Y.; Liu, Z. Chemical vapour deposition of group-VIB metal dichalcogenide monolayers: Engineered substrates from amorphous to single crystalline. Chem. Soc. Rev. 2015, 44, 2587–2602. [Google Scholar] [CrossRef] [PubMed]
  2. Shi, J.P.; Ji, Q.Q.; Liu, Z.F.; Zhang, Y.F. Recent advances in controlling syntheses and energy related applications of MX2 and MX2/graphene heterostructures. Adv. Energy Mater. 2016, 6, 24. [Google Scholar] [CrossRef]
  3. Brent, J.R.; Savjani, N.; O’Brien, P. Synthetic approaches to two-dimensional transition metal dichalcogenide nanosheets. Prog. Mater. Sci. 2017, 89, 411–478. [Google Scholar] [CrossRef]
  4. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Najmaei, S.; Liu, Z.; Zhou, W.; Zou, X.; Shi, G.; Lei, S.; Yakobson, B.I.; Idrobo, J.-C.; Ajayan, P.M.; Lou, J. Vapour phase growth and grain boundary structure of molybdenum disulphide atomic layers. Nat. Mater. 2013, 12, 754–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. van der Zande, A.M.; Huang, P.Y.; Chenet, D.A.; Berkelbach, T.C.; You, Y.; Lee, G.-H.; Heinz, T.F.; Reichman, D.R.; Muller, D.A.; Hone, J.C. Grains and grain boundaries in highly crystalline monolayer molybdenum disulphide. Nat. Mater. 2013, 12, 554–561. [Google Scholar] [CrossRef] [PubMed]
  7. Cai, L.; He, J.; Liu, Q.; Yao, T.; Chen, L.; Yan, W.; Hu, F.; Jiang, Y.; Zhao, Y.; Hu, T.; et al. Vacancy-induced ferromagnetism of MoS2 nanosheets. J. Am. Chem. Soc. 2015, 137, 2622–2627. [Google Scholar] [CrossRef] [PubMed]
  8. Taube, A.; Judek, J.; Jastrzebski, C.; Duzynska, A.; Switkowski, K.; Zdrojek, M. Temperature-dependent nonlinear phonon shifts in a supported MoS2 monolayer. ACS Appl. Mater. Interfaces 2014, 6, 8959–8963. [Google Scholar] [CrossRef] [PubMed]
  9. Lanzillo, N.A.; Birdwell, A.G.; Amani, M.; Crowne, F.J.; Shah, P.B.; Najmaei, S.; Liu, Z.; Ajayan, P.M.; Lou, J.; Dubey, M.; et al. Temperature-dependent phonon shifts in monolayer MoS2. Appl. Phys. Lett. 2013, 103, 093102. [Google Scholar] [CrossRef]
  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. Zhao, H.-Q.; Mao, X.; Zhou, D.; Feng, S.; Shi, X.; Ma, Y.; Wei, X.; Mao, Y. Bandgap modulation of MoS2 monolayer by thermal annealing and quick cooling. Nanoscale 2016, 8, 18995–19003. [Google Scholar] [CrossRef] [PubMed]
  12. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  13. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.; Asefa, T.; Shenoy, V.B.; Eda, G.; Chhowalla, M. Conducting MoS2 Nanosheets as Catalysts for Hydrogen evolution reaction. Nano Lett. 2013, 13, 6222–6227. [Google Scholar] [CrossRef] [PubMed]
  14. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef]
  15. Zhang, W.; Huang, J.-K.; Chen, C.-H.; Chang, Y.-H.; Cheng, Y.-J.; Li, L.-J. High-gain phototransistors based on a CVD MoS2 monolayer. Adv. Mater. 2013, 25, 3456–3461. [Google Scholar] [CrossRef] [PubMed]
  16. Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.H.; Tok, A.I.Y.; Zhang, Q.; Zhang, H. Fabrication of single- and multilayer MoS2 film-based field-effect transistors for sensing NO at room temperature. Small 2012, 8, 63–67. [Google Scholar] [CrossRef] [PubMed]
  17. Late, D.J.; Huang, Y.-K.; Liu, B.; Acharya, J.; Shirodkar, S.N.; Luo, J.; Yan, A.; Charles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing behavior of atomically thin-layered MoS2 transistors. ACS Nano 2013, 7, 4879–4891. [Google Scholar] [CrossRef] [PubMed]
  18. Peimyoo, N.; Shang, J.; Cong, C.; Shen, X.; Wu, X.; Yeow, E.K.L.; Yu, T. Nonblinking, intense two-dimensional light emitter: Mono layer WS2 triangles. ACS Nano 2013, 7, 10985–10994. [Google Scholar] [CrossRef] [PubMed]
  19. Zhao, W.; Ghorannevis, Z.; Chu, L.; Toh, M.; Kloc, C.; Tan, P.-H.; Eda, G. Evolution of electronic structure in atomically thin sheets of WS2 and WSe2. ACS Nano 2013, 7, 791–797. [Google Scholar] [CrossRef] [PubMed]
  20. Zhu, Z.Y.; Cheng, Y.C.; Schwingenschloegl, U. Giant spin-orbit-induced spin splitting in two-dimensional transition-metal dichalcogenide semiconductors. Phys. Rev. B 2011, 84, 153402. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Elias, A.L.; Perea-Lopez, N.; Castro-Beltran, A.; Berkdemir, A.; Lv, R.T.; Feng, S.M.; 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]
  23. Orofeo, C.M.; Suzuki, S.; Sekine, Y.; Hibino, H. Scalable synthesis of layer-controlled WS2 and MoS2 sheets by sulfurization of thin metal films. Appl. Phys. Lett. 2014, 105. [Google Scholar] [CrossRef]
  24. Jung, Y.; Shen, J.; Liu, Y.; Woods, J.M.; Sun, Y.; Cha, J.J. Metal Seed Layer Thickness-induced transition from vertical to horizontal growth of MoS2 and WS2. Nano Lett. 2014, 14, 6842–6849. [Google Scholar] [CrossRef] [PubMed]
  25. Park, J.; Kim, M.S.; Cha, E.; Kim, J.; Choi, W. Synthesis of uniform single layer WS2 for tunable photoluminescence. Sci. Rep. 2017, 7, 16121. [Google Scholar] [CrossRef]
  26. Chen, Y.; Gan, L.; Li, H.; Ma, Y.; Zhai, T. Achieving uniform monolayer transition metal dichalcogenides film on silicon wafer via silanization treatment: A typical study on WS2. Adv. Mater. 2017, 29, 1603550. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, Y.S.; Shi, J.P.; Han, G.F.; Li, M.J.; Ji, Q.Q.; Ma, D.L.; Zhang, Y.; Li, C.; Lang, X.Y.; Zhang, Y.F.; et al. Chemical vapor deposition of monolayer WS2 nanosheets on Au foils toward direct application in hydrogen evolution. Nano Res. 2015, 8, 2881–2890. [Google Scholar] [CrossRef]
  28. Jia, Z.; Hu, W.; Xiang, J.; Wen, F.; Nie, A.; Mu, C.; Zhao, Z.; Xu, B.; Tian, Y.; Liu, Z. Grain wall boundaries in centimeter-scale continuous monolayer WS2 film grown by chemical vapor deposition. Nanotechnology 2018, 29, 255705. [Google Scholar] [CrossRef] [PubMed]
  29. Okada, M.; Sawazaki, T.; Watanabe, K.; Taniguch, T.; Hibino, H.; Shinohara, H.; Kitaura, R. Direct chemical vapor deposition growth of WS2 atomic layers on hexagonal boron nitride. ACS Nano 2014, 8, 8273–8277. [Google Scholar] [CrossRef] [PubMed]
  30. Xu, Z.Q.; Zhang, Y.P.; Lin, S.H.; Zheng, C.X.; Zhong, Y.L.; Xia, X.; Li, Z.P.; Sophia, P.J.; Fuhrer, M.S.; Cheng, Y.B.; et al. Synthesis and transfer of large-area monolayer WS2 crystals: Moving toward the recyclable use of sapphire substrates. ACS Nano 2015, 9, 6178–6187. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, Y.; Zhang, Y.F.; Ji, Q.Q.; Ju, J.; Yuan, H.T.; Shi, J.P.; Gao, T.; Ma, D.L.; Liu, M.X.; Chen, Y.B.; 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]
  32. Gao, Y.; Liu, Z.B.; Sun, D.M.; Huang, L.; Ma, L.P.; Yin, L.C.; Ma, T.; Zhang, Z.Y.; Ma, X.L.; Peng, L.M.; et al. Large-area synthesis of high-quality and uniform monolayer WS2 on reusable Au foils. Nat. Commun. 2015, 6, 8569. [Google Scholar] [CrossRef]
  33. Cong, C.X.; Shang, J.Z.; Wu, X.; Cao, B.C.; Peimyoo, N.; Qiu, C.; Sun, L.T.; Yu, T. Synthesis and optical properties of large-area single-crystalline 2D semiconductor WS2 monolayer from chemical vapor deposition. Adv. Opt. Mater. 2014, 2, 131–136. [Google Scholar] [CrossRef]
  34. Lan, F.; Yang, R.; Xu, Y.; Qian, S.; Zhang, S.; Cheng, H.; Zhang, Y. Synthesis of large-scale single-crystalline monolayer WS2 using a semi-sealed Method. Nanomaterials 2018, 8, 100. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, Y.H.; Yu, L.L.; Wang, H.; Fang, W.J.; Ling, X.; Shi, Y.M.; Lin, C.T.; Huang, J.K.; Chang, M.T.; Chang, C.S.; et al. Synthesis and transfer of single-layer transition metal disulfides on diverse surfaces. Nano Lett. 2013, 13, 1852–1857. [Google Scholar] [CrossRef] [PubMed]
  36. Yue, Y.; Chen, J.; Zhang, Y.; Ding, S.; Zhao, F.; Wang, Y.; Zhang, D.; Li, R.; Dong, H.; Hu, W.; et al. Two-dimensional high-quality monolayered triangular WS2 flakes for field-effect transistors. ACS Appl. Mater. Interfaces 2018, 10, 22435–22444. [Google Scholar] [CrossRef] [PubMed]
  37. Zhou, J.; Lin, J.; Huang, X.; Zhou, Y.; Chen, Y.; Xia, J.; Wang, H.; Xie, Y.; Yu, H.; Lei, J.; et al. A library of atomically thin metal chalcogenides. Nature 2018, 556, 355–361. [Google Scholar] [CrossRef] [PubMed]
  38. McCreary, K.M.; Hanbicki, A.T.; Jernigan, G.G.; Culbertson, J.C.; Jonker, B.T. Synthesis of large-area WS2 monolayers with exceptional photoluminescence. Sci. Rep. 2016, 6, 19159. [Google Scholar] [CrossRef] [PubMed]
  39. Fu, Q.; Wang, W.H.; Yang, L.; Huang, J.; Zhang, J.Y.; Xiang, B. Controllable synthesis of high quality monolayer WS2 on a SiO2/Si substrate by chemical vapor deposition. RSC Adv. 2015, 5, 15795–15799. [Google Scholar] [CrossRef]
  40. Liu, P.; Luo, T.; Xing, J.; Xu, H.; Hao, H.; Liu, H.; Dong, J. Large-area WS2 film with big single domains grown by chemical vapor deposition. Nanoscale Res. Lett. 2017, 12, 558. [Google Scholar] [CrossRef]
  41. Rong, Y.M.; Fan, Y.; Koh, A.L.; Robertson, A.W.; He, K.; Wang, S.S.; Tan, H.J.; Sinclair, R.; Warner, J.H. Controlling sulphur precursor addition for large single crystal domains of WS2. Nanoscale 2014, 6, 12096–12103. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, S.; Rong, Y.; Fan, Y.; Pacios, M.; Bhaskaran, H.; He, K.; Warner, J.H. Shape evolution of monolayer MoS2 crystals grown by chemical vapor deposition. Chem. Mater. 2014, 26, 6371–6379. [Google Scholar] [CrossRef]
  43. Thangaraja, A.; Shinde, S.M.; Kalita, G.; Tanemura, M. An effective approach to synthesize monolayer tungsten disulphide crystals using tungsten halide precursor. Appl. Phys. Lett. 2016, 108, 053104. [Google Scholar] [CrossRef]
  44. McCreary, K.M.; Hanbicki, A.T.; Singh, S.; Kawakami, R.K.; Jernigan, G.G.; Ishigami, M.; Ng, A.; Brintlinger, T.H.; Stroud, R.M.; Jonker, B.T. The Effect of preparation conditions on Raman and photoluminescence of monolayer WS2. Sci. Rep. 2016, 6, 35154. [Google Scholar] [CrossRef] [PubMed]
  45. Berkdemir, A.; Gutierrez, H.R.; Botello-Mendez, A.R.; Perea-Lopez, N.; Elias, A.L.; Chia, C.I.; Wang, B.; Crespi, V.H.; Lopez-Urias, 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]
  46. 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, 9677–9683. [Google Scholar] [CrossRef] [PubMed]
  47. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous lattice vibrations of single- and few-layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [PubMed]
  48. Chakraborty, B.; Bera, A.; Muthu, D.V.S.; Bhowmick, S.; Waghmare, U.V.; Sood, A.K. Symmetry-dependent phonon renormalization in monolayer MoS2 transistor. Phys. Rev. B 2012, 85, 161403. [Google Scholar] [CrossRef]
  49. McCreary, A.; Berkdemir, A.; Wang, J.J.; Nguyen, M.A.; Elias, A.L.; Perea-Lopez, N.; Fujisawa, K.; Kabius, B.; Carozo, V.; Cullen, D.A.; et al. Distinct photoluminescence and Raman spectroscopy signatures for identifying highly crystalline WS2 monolayers produced by different growth methods. J. Mater. Res. 2016, 31, 931–944. [Google Scholar] [CrossRef]
  50. Mignuzzi, S.; Pollard, A.J.; Bonini, N.; Brennan, B.; Gilmore, I.S.; Pimenta, M.A.; Richards, D.; Roy, D. Effect of disorder on Raman scattering of single-layer MoS2. Phys. Rev. B 2015, 91, 195411. [Google Scholar] [CrossRef]
  51. Wang, Y.; Cong, C.; Yang, W.; Shang, J.; Peimyoo, N.; Chen, Y.; Kang, J.; Wang, J.; Huang, W.; Yu, T. Strain-induced direct-indirect bandgap transition and phonon modulation in monolayer WS2. Nano Res. 2015, 8, 2562–2572. [Google Scholar] [CrossRef]
  52. Wang, Y.; Cong, C.; Qiu, C.; Yu, T. Raman spectroscopy study of lattice vibration and crystallographic orientation of monolayer MoS2 under uniaxial strain. Small 2013, 9, 2857–2861. [Google Scholar] [CrossRef] [PubMed]
  53. Molina-Sanchez, A.; Wirtz, L. Phonons in single-layer and few-layer MoS2 and WS2. Phys. Rev. B 2011, 84, 155413. [Google Scholar] [CrossRef]
  54. Reale, F.; Palczynski, P.; Amit, I.; Jones, G.F.; Mehew, J.D.; Bacon, A.; Ni, N.; Sherrell, P.C.; Agnoli, S.; Craciun, M.F.; et al. High-mobility and high-optical quality atomically thin WS2. Sci. Rep. 2017, 7, 14911. [Google Scholar] [CrossRef] [PubMed]
  55. Gutierrez, H.R.; Perea-Lopez, N.; Elias, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; Lopez-Urias, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary room-temperature photoluminescence in triangular WS2 monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef]
  56. Liu, H.; Lu, J.; Ho, K.; Hu, Z.; Dang, Z.; Carvalho, A.; Tan, H.R.; Tok, E.S.; Sow, C.H. Fluorescence concentric triangles: A case of chemical heterogeneity in WS2 atomic monolayer. Nano Lett. 2016, 16, 5559–5567. [Google Scholar] [CrossRef]
  57. Feng, S.; Yang, R.; Jia, Z.; Xiang, J.; Wen, F.; Mu, C.; Nie, A.; Zhao, Z.; Xu, B.; Tao, C.; et al. Strain release induced novel fluorescence variation in CVD-grown monolayer WS2 crystals. ACS Appl. Mater. Interfaces 2017, 9, 34071–34077. [Google Scholar] [CrossRef]
  58. Liu, Z.; Amani, M.; Najmaei, S.; Xu, Q.; Zou, X.; Zhou, W.; Yu, T.; Qiu, C.; Birdwell, A.G.; Crowne, F.J.; et al. Strain and structure heterogeneity in MoS2 atomic layers grown by chemical vapour deposition. Nat. Commun. 2014, 5, 5246. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of the horizontal single-zone furnace employed for the synthesis of WS2 on SiO2/Si substrates. (b) Schematic illustration for the drop-casting of WO3-ethanol suspension solution onto the SiO2/Si substrate. (c) Temperature profile adopted for the synthesis of WS2 flakes at 950 °C for 5 min.
Figure 1. (a) Schematic diagram of the horizontal single-zone furnace employed for the synthesis of WS2 on SiO2/Si substrates. (b) Schematic illustration for the drop-casting of WO3-ethanol suspension solution onto the SiO2/Si substrate. (c) Temperature profile adopted for the synthesis of WS2 flakes at 950 °C for 5 min.
Nanomaterials 09 00578 g001
Figure 2. Representative optical images of the as-grown WS2 flakes on the covering (a) and bottom (b) SiO2/Si substrates, respectively. AFM image (c) and its corresponding height profiles (d) of an equilateral triangular WS2 flake on SiO2/Si substrate. High-resolution TEM image (e) and its corresponding SAED pattern (f) of the as-grown triangular WS2 transferred on a holy carbon-coated copper TEM grid.
Figure 2. Representative optical images of the as-grown WS2 flakes on the covering (a) and bottom (b) SiO2/Si substrates, respectively. AFM image (c) and its corresponding height profiles (d) of an equilateral triangular WS2 flake on SiO2/Si substrate. High-resolution TEM image (e) and its corresponding SAED pattern (f) of the as-grown triangular WS2 transferred on a holy carbon-coated copper TEM grid.
Nanomaterials 09 00578 g002
Figure 3. Room temperature Raman (a) and PL (b) spectra collected from an equilateral triangular WS2 flake on SiO2/Si substrate with 532 nm excitation. The inset plotted in Figure 3a shows the schematic diagram of the atomic vibrations of A 1 g ( Г ) and E 2 g 1 ( Г ) modes in WS2. It is noted that the black curve is raw data while the color ones are obtained by Lorentz fitting as shown in Figure 3a.
Figure 3. Room temperature Raman (a) and PL (b) spectra collected from an equilateral triangular WS2 flake on SiO2/Si substrate with 532 nm excitation. The inset plotted in Figure 3a shows the schematic diagram of the atomic vibrations of A 1 g ( Г ) and E 2 g 1 ( Г ) modes in WS2. It is noted that the black curve is raw data while the color ones are obtained by Lorentz fitting as shown in Figure 3a.
Nanomaterials 09 00578 g003
Figure 4. (a) FL image of the as-grown WS2 flake on SiO2/Si substrate, its optical image is shown in Figure 2a. (bf) Raman mappings of the specific WS2 sample: frequency difference between A 1 g ( Г ) and E 2 g 1 ( Г ) modes (b), frequency of A 1 g ( Г ) mode (c), full width of half maximum (FWHM) of A 1 g ( Г ) mode (d), normalized intensity of A 1 g ( Г ) mode (e), and frequency of E 2 g 1 ( Г ) mode (f).
Figure 4. (a) FL image of the as-grown WS2 flake on SiO2/Si substrate, its optical image is shown in Figure 2a. (bf) Raman mappings of the specific WS2 sample: frequency difference between A 1 g ( Г ) and E 2 g 1 ( Г ) modes (b), frequency of A 1 g ( Г ) mode (c), full width of half maximum (FWHM) of A 1 g ( Г ) mode (d), normalized intensity of A 1 g ( Г ) mode (e), and frequency of E 2 g 1 ( Г ) mode (f).
Nanomaterials 09 00578 g004
Figure 5. FL images of the as-grown WS2 flakes with different edge lengths: ~171 μm (a), ~95 μm (b), and less than 32 μm (c).
Figure 5. FL images of the as-grown WS2 flakes with different edge lengths: ~171 μm (a), ~95 μm (b), and less than 32 μm (c).
Nanomaterials 09 00578 g005
Figure 6. (a–d) Representative optical images of the as-grown WS2 flakes on the covering SiO2/Si substrates obtained from different batches of experiments under the same conditions (950 °C for 5 min.).
Figure 6. (a–d) Representative optical images of the as-grown WS2 flakes on the covering SiO2/Si substrates obtained from different batches of experiments under the same conditions (950 °C for 5 min.).
Nanomaterials 09 00578 g006

Share and Cite

MDPI and ACS Style

Shi, B.; Zhou, D.; Fang, S.; Djebbi, K.; Feng, S.; Zhao, H.; Tlili, C.; Wang, D. Facile and Controllable Synthesis of Large-Area Monolayer WS2 Flakes Based on WO3 Precursor Drop-Casted Substrates by Chemical Vapor Deposition. Nanomaterials 2019, 9, 578. https://0-doi-org.brum.beds.ac.uk/10.3390/nano9040578

AMA Style

Shi B, Zhou D, Fang S, Djebbi K, Feng S, Zhao H, Tlili C, Wang D. Facile and Controllable Synthesis of Large-Area Monolayer WS2 Flakes Based on WO3 Precursor Drop-Casted Substrates by Chemical Vapor Deposition. Nanomaterials. 2019; 9(4):578. https://0-doi-org.brum.beds.ac.uk/10.3390/nano9040578

Chicago/Turabian Style

Shi, Biao, Daming Zhou, Shaoxi Fang, Khouloud Djebbi, Shuanglong Feng, Hongquan Zhao, Chaker Tlili, and Deqiang Wang. 2019. "Facile and Controllable Synthesis of Large-Area Monolayer WS2 Flakes Based on WO3 Precursor Drop-Casted Substrates by Chemical Vapor Deposition" Nanomaterials 9, no. 4: 578. https://0-doi-org.brum.beds.ac.uk/10.3390/nano9040578

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