Next Article in Journal
Growth by the μ-PD Method and Visible Laser Operation of a Single-Crystal Fiber of Pr3+:KY3F10
Next Article in Special Issue
Advanced Scanning Probe Microscopy of Graphene and Other 2D Materials
Previous Article in Journal
An Overview of the Top Ten Detergents Used for Membrane Protein Crystallization
Previous Article in Special Issue
Role of the Potential Barrier in the Electrical Performance of the Graphene/SiC Interface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis Methods of Two-Dimensional MoS2: A Brief Review

1
Key Laboratory of Optoelectronics Technology, College of Microelectronics, Beijing University of Technology, Beijing 100124, China
2
Quantum Device Physics Laboratory, Department of Microtechnology and Nanoscience, Chalmers University of Technology, Göteborg 41296, Sweden
3
High-Frequency High-Voltage Device and Integrated Circuits Center, Institute of Microelectronics, Chinese Academy of Sciences, Beijing 10029, China
4
Electronics Material and Systems Laboratory, Department of Microtechnology and Nanoscience, Chalmers University of Technology, Göteborg 41296, Sweden
*
Authors to whom correspondence should be addressed.
Submission received: 19 May 2017 / Revised: 20 June 2017 / Accepted: 28 June 2017 / Published: 1 July 2017
(This article belongs to the Special Issue Integration of 2D Materials for Electronics Applications)

Abstract

:
Molybdenum disulfide (MoS2) is one of the most important two-dimensional materials after graphene. Monolayer MoS2 has a direct bandgap (1.9 eV) and is potentially suitable for post-silicon electronics. Among all atomically thin semiconductors, MoS2’s synthesis techniques are more developed. Here, we review the recent developments in the synthesis of hexagonal MoS2, where they are categorized into top-down and bottom-up approaches. Micromechanical exfoliation is convenient for beginners and basic research. Liquid phase exfoliation and solutions for chemical processes are cheap and suitable for large-scale production; yielding materials mostly in powders with different shapes, sizes and layer numbers. MoS2 films on a substrate targeting high-end nanoelectronic applications can be produced by chemical vapor deposition, compatible with the semiconductor industry. Usually, metal catalysts are unnecessary. Unlike graphene, the transfer of atomic layers is omitted. We especially emphasize the recent advances in metalorganic chemical vapor deposition and atomic layer deposition, where gaseous precursors are used. These processes grow MoS2 with the smallest building-blocks, naturally promising higher quality and controllability. Most likely, this will be an important direction in the field. Nevertheless, today none of those methods reproducibly produces MoS2 with competitive quality. There is a long way to go for MoS2 in real-life electronic device applications.

1. Introduction

Recently, graphene has received considerable attention in scientific communities. It is recognized as a new wonder material (e.g., record carrier mobility, thermal conductivity, mechanical strength and flexibility), and applications in nanocomposites, paints, transparent electrodes, heat spreaders, etc. are being explored [1]. However, one should not forget that the original motivation for studying graphene was in its field effect [2], with a hope that it might play a key role in nanoelectronics, rather than serving as a passive element such as electrodes [3]. Nevertheless, at this stage, due to the difficulties in opening an energy bandgap in graphene, it is not clear whether graphene will become a key material in electronics at all. Therefore, the search for other two-dimensional (2D) materials (materials with one or few atomic layers) with an inherent energy bandgap that is suitable to make transistors is essential. Black phosphorus [4], transition metal dichalcogenide (TMD) [5], gallium nitride [6], etc. are extensively explored. For any type of application, the material production is always the first experimental step. Among these 2D semiconductors, molybdenum disulfide (MoS2) is relatively more mature in terms of its synthesis technologies. Hence, in this paper, we provide a short literature survey of recent progresses in the synthesis of MoS2. The paper focuses only on the production techniques rather than serving as a complete review on MoS2-related physics, materials and electronics. Also, we put special emphasis on the recent achievements in metalorganic chemical vapor deposition (MOCVD) and atomic layer deposition (ALD), where gaseous precursors are used to grow MoS2. It has a high potential for producing MoS2 with better quality and controllability for nanoelectronic device applications.

2. Synthesis Methods of MoS2

MoS2 has several polymorphs, where the 2H MoS2 has a layered hexagonal structure with lattice parameters of a = 0.316 nm and c = 1.229 nm [7], yielding an inter-layer spacing of 0.615 nm [8], as schematically shown in Figure 1. Bulk 2H MoS2 is an indirect bandgap semiconductor (~1.29 eV). With decreasing thickness, the bandgap increases. When it comes to monolayer MoS2, it becomes a direct bandgap material, and the energy gap can be as large as ~1.90 eV [9], suitable for fabricating most electronic devices. The synthesis methods of MoS2 can be divided into top-down and bottom-up approaches. We hereby describe them in turn.

2.1. Top-down Approaches

2.1.1. Micromechanical Exfoliation

Similar to the micromechanical exfoliation of graphene (a vivid lab demonstration can be found in [10]), MoS2 flakes can be produced on a substrate using sticky tapes. The starting material is bulk MoS2, where some parts are peeled off with the tape and pressed into the substrate. Upon the release of the tape, some parts stay with the substrate rather than the tape due to the van der Waals force to the substrate. Repeating the process may produce flakes of MoS2 with random shapes, sizes and numbers of layers. This method offers 2D materials of the highest quality, enabling to study the pristine properties and ultimate device performances. For example, Kis et al. have used micro-exfoliation to produce MoS2 monolayers that are suitable for ultrasensitive photodetectors [11], analogue [12] and digital circuits [13], and observation of interesting physics such as mobility engineering and a metal–insulator transition [14]. However, it was found that the van der Waals adhesion of many TMDs to SiO2, a typical substrate, is much weaker than that of graphene. Therefore, the produced flakes are very small (typically <10 µm) in their lateral size [15]. These results notwithstanding, large area MoS2 flakes can be exfoliated by slightly modifying the technique. Magda et al. obtained mono MoS2 layers with lateral size of several hundreds of microns by improving the MoS2-substrate adhesion, as shown in Figure 2 [15]. The mechanism is that sulfur atoms can bind more strongly to gold than to SiO2. The drawback is that such as-exfoliated MoS2 flakes have to be transferred to another new substrate, since in most applications the material is used on insulating substrates. In addition, like their graphene counterpart, the tape-assisted micromechanical exfoliation method offers very low yield, and it’s not possible to scale up for large volume production, therefore MoS2 synthesized by this method is limited to fundamental study at lab scale.

2.1.2. Liquid Phase Exfoliation

Liquid phase exfoliation also starts from bulk MoS2, producing flakes with random shapes, sizes and number of layers. The quantities, however, are much larger, and the quality lower. Roughly, there are 2 routes to exfoliate MoS2 in solution. The first is exfoliation by mechanical means such as sonication, shearing, stirring, grinding and bubbling [16,17,18,19,20,21]. The essence is purely physical, although some chemistry may still be involved [22,23,24]. For example, surfactants such as sodium deoxycholate (SDC) bile salt [22] and chitosan [23] may be added to the solution to prevent the exfoliated flakes from recombination. Electrolysis is employed to generate bubbles in some cases [24]. It is known that fine bubbles can squeeze into material interfaces and exfoliate them [25]. The yield of this method is dramatically improved from tape-assisted exfoliation, but the efficiency is still low for industrial applications. Therefore, a second route by atomic intercalation via solution chemistry [26] or electrochemistry [27] has been proposed. Lithium is typically used to intercalate between the MoS2 layers and enlarge the interlayer spacing, easing the following exfoliation by mechanical treatment (e.g., sonication). Fan et al. developed a process to exfoliate MoS2 nanosheets very efficiently using sonication assisted Li intercalation. They found that the complete Li intercalation with butylithium can be effected within 1.5 hours [28]. An example of the MoS2 flakes synthesized by this method is shown in Figure 3. However, this method results in the loss of semiconducting properties due to the structural changes during Li intercalation. It is reported that above an annealing temperature of 300 °C, these MoS2 flakes can be somewhat recovered [26].
Liquid phase exfoliation is a simple technology with low cost, producing large quantity 2D nanosheets with relatively high quality. Therefore, it is regarded as the most suitable method for the large-scale industrial production of one- to few-layer 2D material flakes at low expense [29]. It is worth noting that liquid phase exfoliation can produce not only wet suspensions of flakes or dry powders, but macroscopically continuous thin films as well [30,31,32]. On substrates, through drop casting of the suspension and drying, MoS2 thin films are prepared by Walia et al. [30,31,32], where optical and electrical properties can be studied.

2.2. Bottom-up Approaches

2.2.1. Physical Vapor Deposition

Although molecular beam epitaxy (MBE) is an advanced technology for growing single crystal semiconductor thin films, its application in producing 2D materials needs further development, and the grain size of the 2D material is not as high as expected [33]. In fact, even ordinary physical vapor deposition is rare. There is a report on growing MoS2-Ti composite by vacuum sputtering, using Ti and MoS2 as the targets [34]. However, the MoS2 is an amorphous structure in this case.

2.2.2. Solution Chemical Process

There are various methods for producing MoS2 by solution chemistry. Hydrothermal synthesis [35] and solvothermal synthesis [36] typically use molybdate to react with sulfide or just sulfur in a stainless steel autoclave, where a series of physicochemical reactions take place under relatively high temperature (e.g., 200 °C) and high pressure for several hours or longer. The resultant is MoS2 powders of different shapes. The size of individual particles can be adjusted to some extent. Very frequently, the powders are high-temperature post-annealed, to improve their crystalline quality and purity. The only difference between hydrothermal and solvothermal synthesis is that the precursor solution in the latter case is usually not aqueous. Other solution chemical processes start at almost room temperature and atmospheric pressure, where post-annealing is often used anyway [37,38,39,40,41]. The products can be either a powder or thin film, depending on the preparation details. The most commonly-used precursor is (NH4)2MoS4 (ammonium tetrathiomolybdate), or something similar. (NH4)2MoS4 decomposes to form MoO3 at 120–360 °C, which can be further converted into MoS2. In some cases, electrochemical deposition [42,43,44] and photo-assisted deposition are used [45]. For instance, Li et al. fabricated MoS2 nanowires and nanoribbons using MoOx as the precursor [43]. The MoOx precursor is then exposed to H2S at high temperature (500–800 °C) to synthesize MoS2. Depending on the temperature for processing, both 2H and 3R phase of MoS2 can be synthesized. The MoS2 nanoribbons can be 50 µm in length and are aligned in parallel arrays which are preferable for device fabrication, as shown in Figure 4.

2.2.3. Chemical Vapor Deposition (CVD)

Among all the synthesis methods, CVD is the most compatible to existing semiconductor technology. Typically, it forms the film on a substrate, where chemical reactions are involved in the growth mechanism. The large-area growth of highly uniform MoS2 will enable the batch fabrication of atomically thin devices and circuitry. Currently, most works start with solid state precursors of molybdenum, such as MoO3 [46,47] or Mo [48]. If the Mo precursors are not vaporized in the process, often the technique is also called sulfurization [46,47,48]. Typical S precursors are H2S gas [49] or vaporized S [46,47,48,50]. It is argued that using MoCl5 precursors are easier to achieve monolayer MoS2 over large area [50]. Earlier, the MoS2 could be noncontinuous [47], which is not a big problem nowadays. The usual growth temperatures are 700–1000 °C, and the inclusion of a metal catalyst such as Au is favorable for the film quality [49]. Typical reactions taking place in the CVD are one of these:
MoO3 + 2H2S + H2 = MoS2 + 3H2O
8MoO3 + 24H2S = 8MoS2 + 24H2O+S8
16MoO3 + 7S8 = 16MoS2 + 24SO2
With plasma-enhanced chemical vapor deposition (PECVD), the growth temperature of MoS2 can be lowered to 150–300 °C, making it possible to directly deposit MoS2 even on plastic substrates [51].
Metal-organic CVD is a special case of CVD where organometallic precursors are used. It is used in the modern semiconductor industry to produce single crystal expitaxial films. Recently, a few reports have appeared on the application of this method in the growth of MoS2 thin films [49,52,53,54]. Kang et al. [52] have grown MoS2 and WS2 on 4-inch oxidized silicon wafers by MOCVD. The MoS2 is grown with gaseous precursors of Mo(CO)6 (boiling point 156 °C) and (C2H5)2S. At 550 °C, monolayer MoS2 with full coverage of the substrate (typically SiO2/Si and fused silica) is deposited for t0~26 h. Figure 5a shows scanning electron microscopy (SEM) images of the samples, demonstrating the effect of adding H2 in the gas mixture. It is argued that H2 enhances the decomposition of (C2H5)2S (increasing nucleation due to hydrogenolysis) and etches MoS2 (preventing intergrain connection) [52]. It seems that a low H2 rate is beneficial for growing continuous monolayer MoS2. Nevertheless, H2 is necessary for removing the carbonaceous species generated during the MOCVD. Therefore, the H2 flow should be optimized. To have successful growth, dehydrating the growth environment is also necessary. Figure 5b indicates that the growth mode is layer-by-layer under this specific condition. When t < t0 there is almost no nucleation of the second MoS2 layer, which takes place mainly at the grain boundaries of the first layer when t > t0. Figure 5c shows a top gated (VTG) MoS2 transistor with a field-effect mobility of 29 cm2 V−1 s−1 and transconductance 2 µSµm, determined from the transfer properties. The current saturation occurs at relatively low VSD (see the inset). These features compete well with the-state-of-the-art MoS2 transistors. Figure 5d–f demonstrates the optical absorption, Raman and photoluminescence spectra measured from the MOCVD grown MoS2 (and WS2) films, respectively, where the features agree well with the characteristics of exfoliated monolayer counterparts (denoted by diamonds).
At this stage, the material quality in [53] is still too low to be competitive. The large-area deposition of monolayer MoS2 as well as other TMDs with spatial homogeneity and high performance is still challenging. MoS2 films today are polycrystalline with the grain size typically ranging from 1 to 10 micrometers. This notwithstanding, the MOCVD synthesis of 2D semiconductors indeed opens a pathway towards future exciting nanoelectronic materials. The most important reason for this is that, in MOCVD, ultrapure gaseous precursors are injected into the reactor and finely dosed to deposit a very thin layer of atoms onto the substrate. Even if a solid or liquid precursor is used, it will be vaporized in situ in the chamber [53]. Unlike solid and liquid precursors, here the building-blocks for growing the film are much finer, naturally leading to better quality and controllability. The heated metalorganic precursor molecules decompose via pyrolysis, leaving the atoms on the substrate. The atoms bond to the surface, and an ultrathin MoS2 layer is grown.
Based on this argument, one can easily think of another promising synthesis method: ALD [55], which can be seen as a subclass of CVD. In ALD, the film is grown by exposing the substrate surface to alternating gas precursors. In contrast to ordinary CVD, the precursors are never simultaneously present in the chamber. Instead, they are introduced as sequential and non-overlapping pulses, separated by purges in between. In each pulse, molecules react with the surface in a self-limiting way. Hence, it is possible to grow materials with high-thickness precision (layer by layer). Indeed, very recently, people have used ALD to grow MoS2, such as in [56]. Using MoCl5 (vaporized) and H2S precursors, monolayer to few-layer MoS2 on large SiO2/Si or quartz substrates is obtained [56]. In the photos of two 150 mm quartz wafers before and after the ALD, the substrate changes to dark yellow uniformly (~2 layers of MoS2) [56]. It is found that photoresist residues on the surface lead to more nucleation sites for MoS2, but this mechanism remains to be confirmed. Figure 6a plots Raman signals for 50 cycle ALD MoS2, where the E12g and A1g characteristic peaks are sharper and higher for 475 °C deposition temperature, indicating its higher material quality. After the ALD, some samples are annealed in a sulfur environment (H2S or S atmosphere). Figure 6b shows that after sulfur annealing at increasingly high temperature, the characteristic peaks get narrower, stronger and more symmetric. After annealing at 920 °C, the full width at half maximum of the E12g peak is close to that of exfoliated MoS2 monolayer (4.2 cm−1 vs. 3.7 cm−1). By analyzing the separation Δ between E12g and A1g peaks, one can tentatively determine the number of layers in MoS2 films [56]. In Figure 6c, the sample grown at 475 °C and annealed at 920 °C seems to be 1–2 monolayers. Future electronic measurements are needed to further examine the film quality.

3. Conclusions

In this paper, we have reviewed the synthesis methods of MoS2 in terms of top-down and bottom-up approaches. Micromechanical exfoliation offers the highest material quality but is limited by the yield so it is mainly for basic research, whereas liquid phase exfoliation and solution chemical process are of low cost, with decent material quality and suitable for large-scale industrial production. MoS2 thin films aiming for high-end electronic applications are produced by CVD. In this review, we have put a special emphasis on the MOCVD and ALD methods, because they are more sophisticated technology with gas phase precursors and will most likely be the dominant techniques in producing MoS2 films oriented towards real electronic device applications in the future. In most cases the MoS2 films are transfer-free, which is vital for industrial applications that require a high process controllability and reproducibility [57]. At the moment, production of MoS2 as a high quality electronic material is still challenging. This notwithstanding, most of the reviewed processes can be extended to other 2D materials (with some modification), and with future technological advances, the successful production of advanced 2D nanoelectronic materials can be envisioned.

Acknowledgments

We would like to thank National Natural Science Foundation of China (11674016), National key R & D program (2017YFB0403100), Beijing Natural Science Foundation (4152003), Beijing Municipal Commission of Education (PXM2016_014204_500018), and Swedish Foundation for International Cooperation in Research and Higher Education (CH2015-6202).

Author Contributions

J. Sun, X. Li and Y. Fu wrote the paper, led the discussion and summarized the literature; W. Guo, M. Zhao, C. Xu and J. Deng participated in the discussion; X. Fan and Y. Dong helped to summarize the literature.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferrari, A.C.; Bonaccorso, F.; Fal'Ko, V.; Novoselov, K.S.; Roche, S.; Bøggild, P.; Borini, S.; Koppens, F.H.; Palermo, V.; Pugno, N.; et al. Science and technology roadmap for graphene, related two-dimensional crystals, and hybrid systems. Nanoscale 2015, 7, 4598–4810. [Google Scholar] [CrossRef] [PubMed]
  2. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, K.; Xu, C.; Deng, J.; Zhu, Y.; Guo, W.; Mao, M.; Zheng, L.; Sun, J. Graphene transparent electrodes grown by rapid chemical vapor deposition with ultrathin indium tin oxide contact layers for GaN light emitting diodes. Appl. Phys. Lett. 2013, 102, 162102. [Google Scholar]
  4. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black phosphorus field-effect transistors. Nature Nanotechnol. 2014, 9, 372–377. [Google Scholar] [CrossRef] [PubMed]
  5. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nature Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  6. Al Balushi, Z.Y.; Wang, K.; Ghosh, R.K.; Vilá, R.A.; Eichfeld, S.M.; Caldwell, J.D.; Qin, X.; Lin, Y.C.; DeSario, P.A.; Stone, G.; et al. Two-dimensional gallium nitride realized via graphene encapsulation. Nat. Mater. 2016, 15, 1166–1171. [Google Scholar] [CrossRef] [PubMed]
  7. Stewart, J.A.; Spearot, D.E. Atomistic simulations of nanoindentation on the basal plane of crystalline molybdenum disulfide (MoS2). Model. Simul. Mater. Sci. Eng. 2013, 21, 045003. [Google Scholar] [CrossRef]
  8. Ai, K.; Ruan, C.; Shen, M.; Lu, L. MoS2 nanosheets with widened interlayer spacing for high-efficiency removal of mercury in aquatic systems. Adv. Funct. Mater. 2016, 26, 5542–5549. [Google Scholar] [CrossRef]
  9. 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] [PubMed]
  10. Vivid Lab Demonstration. Available online: https://www.youtube.com/watch?v=9l5d0YLZgec&feature=youtu.be (accessed on 30 June 2017).
  11. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nature Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef] [PubMed]
  12. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Small-signal amplifier based on single-layer MoS2. Appl. Phys. Lett. 2012, 101, 043103. [Google Scholar] [CrossRef]
  13. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Integrated circuits and logic operations based on single-layer MoS2. ACS Nano 2011, 5, 9934–9938. [Google Scholar] [CrossRef] [PubMed]
  14. Radisavljevic, B.; Kis, A. Mobility engineering and a metal–insulator transition in monolayer MoS2. Nature Mater. 2013, 12, 815–820. [Google Scholar] [CrossRef] [PubMed]
  15. Magda, G.; Petö, J.; Dobrik, G.; Hwang, C.; Biró, L.; Tapasztó, L. Exfoliation of large-area transition metal chalcogenide single layers. Sci. Rep. 2015, 5, 14714. [Google Scholar] [CrossRef] [PubMed]
  16. Forsberg, V.; Zhang, R.; Bäckström, J.; Dahlström, C.; Andres, B.; Norgren, M.; Andersson, M.; Hummelgård, M.; Olin, H. Exfoliated MoS2 in Water without Additives. PLoS ONE 2016, 11, 0154522. [Google Scholar] [CrossRef] [PubMed]
  17. Gupta, A.; Arunachalam, V.; Vasudevan, S. Liquid-phase exfoliation of MoS2 nanosheets: The critical role of trace water. J. Phys. Chem. Lett. 2016, 7, 4884–4890. [Google Scholar] [CrossRef] [PubMed]
  18. Yao, Y.; Lin, Z.; Li, Z.; Song, X.; Moon, K.-S.; Wong, C.-p. Large-scale production of two-dimensional nanosheets. J. Mater. Chem. 2012, 22, 13494–13499. [Google Scholar] [CrossRef]
  19. Yu, Y.; Jiang, S.; Zhou, W.; Miao, X.; Zeng, Y.; Zhang, G.; Liu, S. Room temperature rubbing for few-layer two-dimensional thin flakes directly on flexible polymer substrates. Sci. Rep. 2013, 3, 2697. [Google Scholar] [CrossRef] [PubMed]
  20. Varrla, E.; Backes, C.; Paton, K.; Harvey, A.; Gholamvand, Z.; McCauley, J.; Coleman, J. Large-scale production of size-controlled MoS2 nanosheets by shear exfoliation. Chem. Mater. 2015, 27, 1129–1139. [Google Scholar] [CrossRef]
  21. Paton, K.; Varrla, E.; Backes, C.; Smith, R.J.; Khan, U.; O’Neill, A.; Boland, C.; Lotya, M.; Istrate, O.M.; King, P.; et al. Scalable production of large quantities of defect-free few-layer graphene by shear exfoliation in liquids. Nature Mater. 2014, 13, 624–630. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, M.; Howe, R.; Woodward, R.; Kelleher, E.; Torrisi, F.; Hu, G.; Popov, S.; Taylor, J.; Hasan, T. Solution processed MoS2-PVA composite for sub-bandgap mode-locking of a wideband tunable ultrafast Er:fiber laser. Nano Res. 2015, 8, 1522–1534. [Google Scholar] [CrossRef]
  23. Zhang, W.; Wang, Y.; Zhang, D.; Yu, S.; Zhu, W.; Wang, J.; Zheng, F.; Wang, S.; Wang, J. A one-step approach to the large-scale synthesis of functionalized MoS2 nanosheets by ionic liquid assisted grinding. Nanoscale 2015, 7, 10210–10217. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, N.; Kim, P.; Kim, J.; Ye, J.; Kim, S.; Lee, C. Large-area atomically thin MoS2 nanosheets prepared using electrochemical exfoliation. ACS Nano 2014, 8, 6902–6910. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, L.; Liu, X.; Zhan, Z.; Guo, W.; Xu, C.; Deng, J.; Chakarov, D.; Hyldgaard, P.; Schröder, E.; Yurgens, A.; Sun, J. A mechanism for highly efficient electrochemical bubbling delamination of CVD-grown graphene from metal substrates. Adv. Mater. Interf. 2016, 3, 1500492. [Google Scholar] [CrossRef]
  26. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef] [PubMed]
  27. Zeng, Z.; Yin, Z.; Huang, X.; Li, H.; He, Q.; Lu, G.; Boey, F.; Zhang, H. Single layer semiconducting nanosheets: High-yield preparation and device fabrication. Angew. Chem. Int. Ed. 2011, 50, 11093–11097. [Google Scholar] [CrossRef] [PubMed]
  28. Fan, X.; Xu, P.; Zhou, D.; Sun, Y.; Li, Y.; Nguyen, M.; Terrones, M.; Mallouk, T. Fast and efficient preparation of exfoliated 2H MoS2 nanosheets by sonication-assisted lithium intercalation and infrared laser-induced 1T to 2H phase reversion. Nano Lett. 2015, 15, 5956–5960. [Google Scholar] [CrossRef] [PubMed]
  29. Coleman, J.; Lotya, M.; O’Neill, A.; Bergin, S.D.; King, P.J.; Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R.J.; et al. Two-dimensional nanosheets produced by liquid exfoliation of layered materials. Science 2011, 331, 568–571. [Google Scholar] [CrossRef] [PubMed]
  30. Rezk, A.R.; Walia, S.; Ramanathan, R.; Nili, H.; Ou, J.Z.; Bansal, V.; Friend, J.R.; Bhaskaran, M.; Yeo, L.Y.; Sriram, S. Acoustic–excitonic coupling for dynamic photoluminescence manipulation of quasi-2D MoS2 nanoflakes. Adv. Opt. Mater. 2015, 3, 888–894. [Google Scholar] [CrossRef]
  31. Balendhran, S.; Walia, S.; Nili, H.; Ou, J.Z.; Zhuiykov, S.; Kaner, R.B.; Sriram, S.; Bhaskaran, M.; Kalantar-zadeh, K. Two-dimensional molybdenum trioxide and dichalcogenides. Adv. Funct. Mater. 2013, 23, 3952–3970. [Google Scholar] [CrossRef]
  32. Walia, S.; Balendhran, S.; Wang, Y.; Kadir, R.A.; Zoolfakar, A.S.; Atkin, P.; Ou, J.Z.; Sriram, S.; Kalantar-zadeh, K.; Bhaskaran, M. Characterization of metal contacts for two-dimensional MoS2 nanoflakes. Appl. Phys. Lett. 2013, 103, 232105. [Google Scholar] [CrossRef]
  33. Vishwanath, S.; Liu, X.; Rouvimov, S.; Mende, P.; Azcatl, A.; McDonnell, S.; Wallace, R.; Feenstra, R.; Furdyna, J.; Jena, D.; Xing, H. Comprehensive structural and optical characterization of MBE grown MoSe2 on graphite, CaF2 and graphene. 2D Mater. 2015, 2, 024007. [Google Scholar] [CrossRef]
  34. Qin, X.; Ke, P.; Wang, A.; Kim, K. Microstructure, mechanical and tribological behaviors of MoS2-Ti composite coatings deposited by a hybrid HIPIMS method. Surf. Coat. Technol. 2013, 228, 275–281. [Google Scholar] [CrossRef]
  35. Zhou, X.; Xu, B.; Lin, Z.; Shu, D.; Ma, L. Hydrothermal synthesis of flower-like MoS2 nanospheres for electrochemical supercapacitors. J. Nanosci. Nanotechnol. 2014, 14, 7250–7254. [Google Scholar] [CrossRef] [PubMed]
  36. Feng, X.; Tang, Q.; Zhou, J.; Fang, J.; Ding, P.; Sun, L.; Shi, L. Novel mixed–solvothermal synthesis of MoS2 nanosheets with controllable morphologies. Cryst. Res. Technol. 2013, 48, 363–368. [Google Scholar] [CrossRef]
  37. Liao, H.; Wang, Y.; Zhang, S.; Qian, Y. A solution low-temperature route to MoS2 fiber. Chem. Mater. 2011, 13, 6–8. [Google Scholar] [CrossRef]
  38. Li, X.; Zhang, W.; Wu, Y.; Min, C.; Fang, J. Solution-processed MoSx as an efficient anode buffer layer in organic solar cells. ACS Appl. Mater. Interf. 2013, 5, 8823–8827. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, K.K.; Zhang, W.; Lee, Y.-H.; Lin, Y.-C.; Chang, M.-T.; Su, C.-Y.; Chang, C.-S.; Li, H.; Shi, Y.; Zhang, H.; Lai, C.-S.; Li, L.-J. Growth of large-area and highly crystalline MoS2 thin layers on insulating substrates. Nano Lett. 2012, 12, 1538–1544. [Google Scholar] [CrossRef] [PubMed]
  40. Bezverkhy, I.; Afanasiev, P.; Lacroix, M. Aqueous preparation of highly dispersed molybdenum sulfide. Inorg. Chem. 2000, 39, 5416–5417. [Google Scholar] [CrossRef] [PubMed]
  41. Afanasiev, P.; Geantet, C.; Thomozeau, C.; Jouget, B. Molybdenum polysulfide hololow microtubules grown at room temperature from solution. Chem. Commun. 2000, 12, 1001–1002. [Google Scholar] [CrossRef]
  42. Maijenburg, A.; Regis, M.; Hattori, A.; Tanaka, H.; Choi, K.-S.; ten Elshof, J. MoS2 nanocube structures as catalysts for electrochemical H2 evolution from acidic aqueous solutions. ACS Appl. Mater. Interf. 2014, 6, 2003–2010. [Google Scholar] [CrossRef] [PubMed]
  43. Li, Q.; Walter, E.; van der Veer, W.; Murray, B.; Newberg, J.; Bohannan, E.; Switzer, J.; Hemminger, J.; Penner, R. Molybdenum disulfide nanowires and nanoribbons by electrochemical/chemical synthesis. J. Phys. Chem. B 2005, 109, 3169–3182. [Google Scholar] [CrossRef] [PubMed]
  44. Kibsgaard, J.; Chen, Z.; Reinecke, B.; Jaramillo, T. Engineering the surface structure of MoS2 to preferentially expose active edge sites for electrocatalysis. Nature Mater. 2012, 11, 963–969. [Google Scholar] [CrossRef] [PubMed]
  45. Nguyen, M.; Tran, P.; Pramana, S.; Lee, R.; Batabyal, S.; Mathews, N.; Wong, L.; Graetzel, M. In situ photo-assisted deposition of MoS2 electrocatalyst onto zinc cadmium sulphide nanoparticle surfaces to construct an efficient photocatalyst for hydrogen generation. Nanoscale 2013, 5, 1479–1482. [Google Scholar] [CrossRef] [PubMed]
  46. Lin, Y.-C.; Zhang, W.; Huang, J.-K.; Liu, K.-K.; Lee, Y.-H.; Liang, C.-T.; Chu, C.-W.; Li, L.-J. Wafer-scale MoS2 thin layers prepared by MoO3 sulfurization. Nanoscale 2012, 4, 6637–6641. [Google Scholar] [CrossRef] [PubMed]
  47. Cai, J.; Jian, J.; Chen, X.; Lei, M.; Wang, W. Regular hexagonal MoS2 microflakes grown from MoO3 precursor. Appl. Phys. A 2007, 89, 783–788. [Google Scholar] [CrossRef]
  48. Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P.; Lou, J. Large-area vapor-phase growth and characterization of MoS2 atomic layers on a SiO2 substrate. Small 2014, 8, 966–971. [Google Scholar] [CrossRef] [PubMed]
  49. Song, I.; Park, C.; Hong, M.; Baik, J.; Shin, H.-J.; Choi, H. Patternable large-scale molybdenium disulfide atomic layers grown by gold-assisted chemical vapor deposition. Angew. Chem. Int. Ed. 2014, 53, 1266–1269. [Google Scholar] [CrossRef] [PubMed]
  50. Yu, Y.; Li, C.; Liu, Y.; Su, L.; Zhang, Y.; Cao, L. Controlled scalable synthesis of uniform, high-quality monolayer and few-layer MoS2 Films. Sci. Rep. 2013, 3, 1866. [Google Scholar] [CrossRef] [PubMed]
  51. Ahn, C.; Lee, J.; Kim, H.-U.; Bark, H.; Jeon, M.; Ryu, G.; Lee, Z.; Yeom, G.; Kim, K.; Jung, J.; Kim, Y.; Lee, C.; Kim, T. Low-temperature synthesis of large-scale molybdenum disulfide thin films directly on a plastic substrate using plasma-enhanced chemical vapor deposition. Adv. Mater. 2015, 27, 5223–5229. [Google Scholar] [CrossRef] [PubMed]
  52. Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P.; Mak, K.; Kim, C.-J.; Muller, D.; Park, J. High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity. Nature 2015, 520, 656–660. [Google Scholar] [CrossRef] [PubMed]
  53. Olofinjana, B.; Egharevba, G.; Taleatu, B.; Akinwunmi, O.; Ajayi, E. MOCVD of molybdenum sulphide thin film via single solid source precursor bis-(morpholinodithioato-s,s’)-Mo. J. Mod. Phys. 2011, 2, 341–349. [Google Scholar] [CrossRef]
  54. Kumar, V.; Dhar, S.; Choudhury, T.; Shivashankar, S.; Raghavan, S. A predictive approach to CVD of crystalline layers of TMDs: The case of MoS2. Nanoscale 2015, 7, 7802–7810. [Google Scholar] [CrossRef] [PubMed]
  55. Sun, J.; Larsson, M.; Maximov, I.; Hardtdegen, H.; Xu, H. Gate-defined quantum-dot devices realized in InGaAs/InP by incorporating a HfO2 layer as gate dielectric. Appl. Phys. Lett. 2009, 94, 042114. [Google Scholar] [CrossRef]
  56. Valdivia, A.; Tweet, D.; Conley, J., Jr. Atomic layer deposition of two dimensional MoS2 on 150 mm substrates. J. Vac. Sci. Technol. A 2016, 34, 021515. [Google Scholar] [CrossRef]
  57. Sun, J.; Cole, M.; Lindvall, N.; Teo, K.; Yurgens, A. Noncatalytic chemical vapor deposition of graphene on high-temperature substrates for transparent electrodes. Appl. Phys. Lett. 2012, 100, 022102. [Google Scholar]
Figure 1. Distances between different layers in 2H MoS2.
Figure 1. Distances between different layers in 2H MoS2.
Crystals 07 00198 g001
Figure 2. Large area MoS2 flakes exfoliated with chemically enhanced adhesion to the substrate [15].
Figure 2. Large area MoS2 flakes exfoliated with chemically enhanced adhesion to the substrate [15].
Crystals 07 00198 g002
Figure 3. TEM, SEM and AFM images of MoS2 flakes exfoliated by Li intercalation and sonication [28].
Figure 3. TEM, SEM and AFM images of MoS2 flakes exfoliated by Li intercalation and sonication [28].
Crystals 07 00198 g003
Figure 4. MoS2 nanowires/nanoribbons synthesized from MoOx precursor [43].
Figure 4. MoS2 nanowires/nanoribbons synthesized from MoOx precursor [43].
Crystals 07 00198 g004
Figure 5. (a) Grain size evolution of monolayer MoS2 as a function of H2 flow. From left to right, 5, 20 and 200 sccm H2. (b) Optical images of MoS2 films as a function of growth time. t0 is the optimal time for full monolayer coverage (~26 h). (c) Transfer and output properties of the field-effect transistor fabricated in the MOCVD grown MoS2 with top gate configuration. σ denotes sheet conductance. Scale bar: 10 µm. (df) Optical absorption, Raman and normalized photoluminescence spectra of as-grown monolayer MoS2 (red) and WS2 (orange) films. The peak positions in (df) are consistent with exfoliated flakes (denoted by diamonds) [52].
Figure 5. (a) Grain size evolution of monolayer MoS2 as a function of H2 flow. From left to right, 5, 20 and 200 sccm H2. (b) Optical images of MoS2 films as a function of growth time. t0 is the optimal time for full monolayer coverage (~26 h). (c) Transfer and output properties of the field-effect transistor fabricated in the MOCVD grown MoS2 with top gate configuration. σ denotes sheet conductance. Scale bar: 10 µm. (df) Optical absorption, Raman and normalized photoluminescence spectra of as-grown monolayer MoS2 (red) and WS2 (orange) films. The peak positions in (df) are consistent with exfoliated flakes (denoted by diamonds) [52].
Crystals 07 00198 g005
Figure 6. (a) Raman spectroscopy measurements of ALD MoS2 at 375 or 475 °C, where the latter shows higher quality. (b) Raman spectra of 475 °C deposited MoS2 films (some are post annealed in sulfur environment), where 920 °C annealing leads to the best quality. The curve for bulk MoS2 is for reference. (c) E12g to A1g peak separation Δ plotted against the cycle number in ALD. Δ~19 cm−1 tentatively indicates MoS2 monolayer, whereas bulk (>5 layers) material has Δ~25 cm−1 [56].
Figure 6. (a) Raman spectroscopy measurements of ALD MoS2 at 375 or 475 °C, where the latter shows higher quality. (b) Raman spectra of 475 °C deposited MoS2 films (some are post annealed in sulfur environment), where 920 °C annealing leads to the best quality. The curve for bulk MoS2 is for reference. (c) E12g to A1g peak separation Δ plotted against the cycle number in ALD. Δ~19 cm−1 tentatively indicates MoS2 monolayer, whereas bulk (>5 layers) material has Δ~25 cm−1 [56].
Crystals 07 00198 g006

Share and Cite

MDPI and ACS Style

Sun, J.; Li, X.; Guo, W.; Zhao, M.; Fan, X.; Dong, Y.; Xu, C.; Deng, J.; Fu, Y. Synthesis Methods of Two-Dimensional MoS2: A Brief Review. Crystals 2017, 7, 198. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst7070198

AMA Style

Sun J, Li X, Guo W, Zhao M, Fan X, Dong Y, Xu C, Deng J, Fu Y. Synthesis Methods of Two-Dimensional MoS2: A Brief Review. Crystals. 2017; 7(7):198. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst7070198

Chicago/Turabian Style

Sun, Jie, Xuejian Li, Weiling Guo, Miao Zhao, Xing Fan, Yibo Dong, Chen Xu, Jun Deng, and Yifeng Fu. 2017. "Synthesis Methods of Two-Dimensional MoS2: A Brief Review" Crystals 7, no. 7: 198. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst7070198

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