Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

CVD synthesis of large-area, highly crystalline MoSe2 atomic layers on diverse substrates and application to photodetectors

Jing Xia a, Xing Huang ab, Ling-Zhi Liu c, Meng Wang a, Lei Wang a, Ben Huang a, Dan-Dan Zhu a, Jun-Jie Li c, Chang-Zhi Gu c and Xiang-Min Meng *a
aKey Laboratory of Photochemical Conversion and Optoelectronic Materials, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing, 100190, P.R. China. E-mail: mengxiangmin@mail.ipc.ac.cn
bDepartment of Inorganic Chemistry, Fritz Haber Institute of the Max Planck Society, Faradayweg 4-6, 14195 Berlin, Germany
cBeijing National Laboratory of Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing, 100190, China

Received 29th April 2014 , Accepted 30th May 2014

First published on 3rd June 2014


Abstract

Synthesis of large-area, atomically thin transition metal dichalcogenides (TMDs) on diverse substrates is of central importance for the large-scale fabrication of flexible devices and heterojunction-based devices. In this work, we successfully synthesized a large area of highly-crystalline MoSe2 atomic layers on SiO2/Si, mica and Si substrates using a simple chemical vapour deposition (CVD) method at atmospheric pressure. Atomic force microscopy (AFM) and Raman spectroscopy reveal that the as-grown ultrathin MoSe2 layers change from a single layer to a few layers. Photoluminescence (PL) spectroscopy demonstrates that while the multi-layer MoSe2 shows weak emission peaks, the monolayer has a much stronger emission peak at ∼1.56 eV, indicating the transition from an indirect to a direct bandgap. Transmission electron microscopy (TEM) analysis confirms the single-crystallinity of MoSe2 layers with a hexagonal structure. In addition, the photoresponse performance of photodetectors based on MoSe2 monolayer was studied for the first time. The devices exhibit a rapid response of ∼60 ms and a good photoresponsivity of ∼13 mA/W (using a 532 nm laser at an intensity of 1 mW mm−2 and a bias of 10 V), suggesting that MoSe2 monolayer is a promising material for photodetection applications.


1. Introduction

The experimental realization of graphene has triggered a worldwide upsurge in research interest on other two-dimensional (2D) layered materials such as VS2, CoS2, GaSe, h-BN and NbSe2.1–7 These 2D crystals exhibit a variety of electrical characteristics including metallic, semimetallic, semiconducting, insulating, and charge density wave behaviour. In particular, atomically-thin group-VIB transition metal dichalcogenides (TMDs) (MX2; M = W, Mo; X = S, Se) are a new class of 2D semiconductors with extraordinary properties and tremendous application potential.8–11 For example, while the bulk of TMDs have an indirect bandgap, the monolayers show a sizable direct bandgap around 1–2 eV, making them appealing materials for ultrathin, lightweight, and flexible device applications such as field effect transistors (FETs), photovoltaic cells, light emitting diodes (LEDs) and photodetectors.8,12–15 Interestingly, the electronic structures of these materials can be effectively modulated by partially substituting the metal or nonmetal atoms with their congeners, forming 2D TMD alloys (Mo1−xWxS2, Mo1−xWxSe2, MoS2(1−x)Se2x, etc.).16–19 The availability of various alloys with tunable bandgaps will not only promote fundamental studies of these materials, but will also allow their practical application in electronics and optics. Recent studies have demonstrated that the breaking of inversion symmetry together with the spin-orbit interaction could result in the coupling of spin and valley physics in monolayer TMDs, making it possible to manipulate the spin and valley degrees of freedom in these novel 2D crystals.20–22

Until now, most studies on group-VIB TMDs have concentrated on MoS2. Other group-VIB TMDs such as MoSe2, however, possess attractive properties. For instance, monolayer MoSe2 has a direct bandgap of ∼1.5 eV, which is close to the optimal bandgap value of single-junction solar cells and photoelectrochemical cells.23,24 Few-layer MoSe2 has nearly degenerate indirect and direct bandgaps, and an increase in temperature can effectively push the system towards the 2D limit.24 In contrast to MoS2, monolayer MoSe2 has a larger spin-splitting energy of ∼180 meV at the top of the valence bands, which makes MoSe2 more applicable than MoS2 in spintronics.25,26 While some methods including exfoliation (chemical or mechanical) from the bulk,27–29 hydrothermal synthesis,27,30 and selenization of metals31,32 have been used to produce MoSe2 nanosheets, none of these approaches are able to synthesize large-area, uniform, and highly-crystalline MoSe2 atomic layers. Although molecular beam epitaxy (MBE) is a facile method to grow large-scale ultrathin MoSe2 layers of high quality,26 it is not suitable for industrial production because of its high cost. Until recently, CVD has been used to successfully grow large-area ultrathin MoSe2 layers on SiO2/Si substrates.33–35

For different applications such as flexible devices36–39 and heterojunction-based devices,40–43 however, the growth of MoSe2 atomic layers on SiO2/Si substrates is not sufficient; different functionalities require the integration of MoSe2 with different kinds of materials. Therefore, the synthesis of large-scale, high-quality, ultrathin MoSe2 layers on diverse substrates is necessary. To this end, we herein report a simple CVD method to directly grow a large area of highly-crystalline MoSe2 atomic layers on SiO2/Si, mica and Si substrates at atmospheric pressure using MoO3 and Se powders as starting materials. Scanning electron microscopy (SEM), optical microscopy (OM), AFM, Raman spectroscopy, PL spectroscopy, and TEM have been used to systematically study the samples. In addition, photodetectors based on monolayer MoSe2 were fabricated and studied for the first time. The devices exhibited a fast response and a good photoresponsivity.

2. Experimental section

Growth of MoSe2 atomic layers

We obtained MoSe2 layers by the selenization of MoO3 powder (99.99%, Beijing Lanyi Chemical Co., Ltd) in a high temperature CVD furnace (GSL-1400X tube furnace, Hefei Kejing Material Technology Co., Ltd). In a typical CVD growth, a ceramic boat containing 10 mg of MoO3 powder located at the centre of the heating zone and another ceramic boat with 0.5 g of Se powder (99.99%, Beijing Lanyi Chemical Co., Ltd) located upstream of the furnace 10–20 cm away from the MoO3 powder were placed in the furnace. Cleaned 300 nm SiO2/Si, mica and the newly-cleaved side-wall of SiO2/Si substrates were placed downstream close to the MoO3 powder along the quartz tube. The furnace chamber was vacuum pumped to expel the air and then filled with high-purity Ar to atmospheric pressure. Next, the centre of the heating zone was heated to 820 °C within 41 min and kept for 15 min before being allowed to cool naturally. During the growth process, high-purity Ar with a flow rate of 10 standard cubic centimeters per minute (sccm) was used as the gas carrier.

Characterization

SEM characterization was carried out on a Hitachi S-4800 with an acceleration voltage of 5–10 kV. Optical images were obtained using an optical microscope (Nikon Inverted Microscope Eclipse Ti-U with a CCD of Nikon Digital Sight). The thickness profile was determined using an atomic force microscope (Veeco Nanoscope V) at a scanning rate of 0.976 Hz with 512 scanning lines. Raman spectra and PL spectra were recorded with a Renishaw InVia Raman microscope with a 50× objective. The Si peak at 520 cm−1 was used as a reference for wavenumber calibration. For the Raman and PL measurements, we used a 532 nm laser and an 1800 I mm−1 grating. The laser was focused on the sample at a power of ∼250 μW to minimize the laser-induced thermal effect. The structure and composition of MoSe2 samples were investigated by a JEOL JEM-2100F at an acceleration voltage of 200 kV and energy dispersive X-ray spectroscopy. For TEM characterization, an ultrathin carbon film supported on a copper grid was used. The method for transferring the MoSe2 sample is similar to that in the previous report on the direct transfer of graphene.44 In our experiments, we didn't wet the sample with isopropanol; instead, 1% HF solution was used directly as the etchant. The detailed procedures are shown in ESI Fig. S2.

Photodetectors based on monolayer MoSe2 were fabricated by evaporating 100 nm Au directly on the top of MoSe2 layers patterned by electron beam lithography (Raith 150). Photocurrent measurement was performed at room temperature under atmospheric conditions. The illuminating 532 nm laser was generated by Spectra-Physics Laser. A laser attenuator was used to obtain different laser intensities, which were confirmed by a laser power meter. The electrical measurement was carried out using a Keithley-4200SCS semiconductor parameter analyser.

3. Results and discussion

In this work, a high-temperature tube furnace was used for the CVD synthesis of MoSe2 atomic layers; the schematic representation of the growth setup is displayed in Fig. 1a. During the CVD growth, the centre of the heating zone was heated to 820 °C under Ar atmosphere (see details in the experimental section). At such a high temperature, MoO3 powder was partially reduced by Se vapour to form volatile suboxide MoO3−x species that were transported downstream by the Ar carrier gas.45 Selenization of vapour-phase MoO3−x could then produce MoSe2 species, which nucleated on the substrate and grew into large MoSe2 atomic layers.45–47 The corresponding schematic is illustrated in Fig. 1b. Recent reports on the growth of MoSe2 atomic layers have used a similar CVD approach.33–35 In their syntheses, the growth was performed at 750 °C or 800 °C under an Ar–H2 gas mixture environment in which H2 gas acted as an indispensable reducing agent. Interestingly, in our method, the CVD growth could take place efficiently without the use of H2.
image file: c4nr02311k-f1.tif
Fig. 1 (a) Growth setup for synthesizing MoSe2 atomic layers. (b) A schematic view illustrating the growth of MoSe2 layers.

Fig. 2a–c show typical SEM images of large-area MoSe2 layers grown on SiO2/Si, mica and Si substrates, respectively. The spatial distribution of MoSe2 layers (darker contrast in Fig. 2a–c) indicates that MoSe2 nucleates randomly on the substrates. For samples grown on SiO2/Si and mica substrates, most of the isolated MoSe2 islands have equilateral triangle morphologies with edge lengths ranging from a few microns to 40 μm (Fig. 2a and b), suggesting the single crystalline nature of these domains.35,48 In contrast, MoSe2 layers synthesized on the cleaved side-wall of SiO2/Si substrates show irregular shapes with sizes larger than 50 μm (Fig. 2c). This result indicates that the substrates have an important influence on the morphology of TMDs.48,49


image file: c4nr02311k-f2.tif
Fig. 2 SEM images of MoSe2 layers synthesized on (a) SiO2/Si, (b) mica and (c) Si substrates. (d) Optical image of MoSe2 layers on a SiO2/Si substrate. The violet area is the bare SiO2/Si substrate; the greyish, light green and pink triangles represent MoSe2. The inset shows a photograph of 2D MoSe2 grown on the SiO2/Si substrate. (e) Optical image of MoSe2 layers on a mica substrate. The inset shows a photograph of 2D MoSe2 grown on the mica substrate. White triangular flakes represent MoSe2; the black area is the mica. (f) AFM image and height profile of an MoSe2 monolayer.

Fig. 2d and e display OM images of MoSe2 layers grown on SiO2/Si and mica substrates, respectively. The sharp colour contrast between the triangular crystallites (greyish triangular flakes in Fig. 2d; white triangular flakes in Fig. 2e) and the substrates (violet area is the bare SiO2/Si substrate in Fig. 2d; black regions represent the mica in Fig. 2e) demonstrates the high uniformity of the MoSe2 domains. The thickness-dependent contrast observed in the OM images reveals single- , bi- , tri- and multi-layered MoSe2 flakes, as marked in Fig. 2d. To determine the thicknesses of these MoSe2 layers, they were characterized with AFM. Fig. 2f shows the AFM image and height profile of a monolayer MoSe2 triangle. The homogeneous colour contrast signifies that this flake has a flat and uniform surface, and its thickness (0.71 nm) confirms that this MoSe2 triangular flake is a monolayer.24,29 AFM images and height profiles are also shown for bilayer and trilayer MoSe2 (Fig. S1).

In order to investigate the crystal structure of the MoSe2 flakes, a direct transfer approach was used to prepare TEM samples. The approach is not only suitable for 2D MoSe2, but also other 2D crystals such as MoS2, WS2 and WSe2 grown on SiO2/Si, mica and Si substrates. Unlike PMMA-assisted transfer technology,35 the direct transfer method is much more rapid and convenient. The detailed procedure can be found in Fig. S2.Fig. 3a displays a bright-field TEM image of a monolayer MoSe2 triangular flake with an edge length of ∼3.5 μm. The lattice fringe measured from the high-resolution TEM (HRTEM) image in Fig. 3b is 0.28 nm, corresponding to the {10[1 with combining macron]0} planes. The selected area electron diffraction (SAED) pattern (inset in Fig. 3b) from the MoSe2 triangle exhibits one set of six-fold symmetry diffraction spots, confirming the single-crystalline nature of this flake with a hexagonal structure. By analysing the HRTEM image in Fig. 3b, we see that the edge of the MoSe2 triangle is perpendicular to the [10[1 with combining macron]0] direction, indicating that the sample has zigzag edges.50,51 The HRTEM phase-contrast image (Fig. 3c) of the MoSe2 triangle shows a honeycomb-like structure, which is consistent with previous reports.47,51


image file: c4nr02311k-f3.tif
Fig. 3 (a) Bright-field TEM image of a monolayer MoSe2 triangle. (b) HRTEM image of the MoSe2 triangle. The inset shows the SAED pattern. (c) HRTEM phase-contrast image of the MoSe2 triangle. The inset shows the hexagonal arrangement of Mo and Se atoms. (d) Bright-field TEM image of a few-layer MoSe2 triangle. (e) HRTEM image and corresponding SAED pattern of the marked region in Fig. 3d. (f) EDX spectrum of the MoSe2 layers.

Few-layer MoSe2 was characterized in addition to monolayer MoSe2 (Fig. 3d and e). The bright-field TEM image in Fig. 3d shows a few-layer MoSe2 triangle with a small triangular island on the top. Intriguingly, the small triangular island is located both at the centre of the bottom triangle and parallel to it, indicating an AB stacking growth mode.33,52 The HRTEM image of the marked region in Fig. 3d is displayed in Fig. 3e. The interface between adjacent layers can be distinguished due to the contrast difference. The corresponding SAED pattern in Fig. 3e demonstrates the single-crystallinity of the few-layer MoSe2 with a hexagonal structure. Furthermore, energy dispersive X-ray spectroscopy (EDX) equipped in TEM was applied to study the composition of the MoSe2 samples. As shown in Fig. 4f, the intense peaks of Mo and Se confirm that MoO3 has been successfully converted into MoSe2.


image file: c4nr02311k-f4.tif
Fig. 4 (a) Raman spectra of MoSe2 layers with different thicknesses. The inset shows schematic representations of the A1g and E2g modes. (b) Room-temperature PL spectra of MoSe2 layers with different thicknesses.

The as-synthesized MoSe2 layers were further investigated by Raman spectroscopy using a 532 nm excitation laser (Fig. 4a). Two typical Raman active modes, i.e., the prominent A1g Raman mode and the weak E2g Raman mode, are observed in the Raman spectra. The A1g mode relates to the out-of-plane vibration of Se atoms, and the E2g mode is associated with the in-plane vibration of Mo and Se atoms (see the inset in Fig. 4a). In general, the location of Raman modes can be used to determine the thickness of 2D materials.6,45,53 In this study, the A1g and E2g modes of single-layer MoSe2 are located at 240.6 cm−1 and 287.5 cm−1, respectively (Fig. 4a). As the layer thickness increases, the A1g Raman mode is blueshifted to 242.3 cm−1 for 2–3L, and to 243.9 cm−1 for thick MoSe2; the E2g Raman mode exhibits a redshift to 285.9 cm−1 for ≥2L. The stiffening of the A1g mode may result from the increasing van der Waals interaction between layers, while the softening of the E2g mode may be caused by the presence of long-range coulomb interactions between layers.54–56 Similar phenomena have been previously reported in other 2D materials including graphene, h-BN, MoS2 and GaSe.5,6,45,53 The frequency difference between the E2g and A1g modes is another important indicator of the thickness of 2D materials, although different substrates may have an effect. In our case, the peak spacing between E2g and A1g modes decreases as the layer number increases. The spacings are 46.9 cm−1, 43.6 cm−1 and 42 cm−1 for MoSe2 monolayers, 2–3L and thick samples, respectively.

Recent studies have demonstrated that the electronic structure of MoSe2 varies with its thickness.8,11 To confirm this behaviour in the CVD-grown samples, PL experiments (532 nm laser) were carried out. Fig. 4b shows the room-temperature PL spectra of MoSe2 layers of different thicknesses. It can be seen that monolayer MoSe2 exhibits a single emission peak at ∼1.56 eV with a very strong PL intensity; which can be ascribed to the direct bandgap at the K high symmetry point of the Brillouin zone.24 With increasing layer number, PL intensity declines sharply. For example, the PL intensity of bilayer MoSe2 is approximately 10-times weaker than that of the monolayer (Fig. 4b). As the thickness increases, the PL peak is also redshifted to ∼1.54 eV for thick MoSe2 layers. The change in PL intensity and bandgap should be attributed to the transition from direct to indirect bandgap, which has been previously reported.24 Generally, Raman spectra and PL spectra demonstrate the high quality of the as-grown MoSe2 atomic layers in this study.

Due to its sizeable direct bandgap of ∼1.56 eV, monolayer MoSe2 could be a promising candidate for photodetection applications. In view of this, photodetectors based on monolayer MoSe2 were fabricated using electron beam lithography and metal evaporation, as illustrated schematically in Fig. 5a. In this study, Au was used as the metal electrode due to its large work function.57,58 In general, 100 nm of Au was deposited on the top of the MoSe2 monolayers by electron beam evaporation. The as-fabricated devices were then annealed at 200 °C for 2 h in an Ar atmosphere to improve the contact.


image file: c4nr02311k-f5.tif
Fig. 5 (a) Three-dimensional schematic view of a monolayer MoSe2 photodetector and the 532 nm laser beam used for illumination. (b) IV characteristics of the device in the dark and in the presence of an illuminating laser with different laser powers. The inset shows the optical image of a monolayer MoSe2 photodetector. (c) Time-resolved photoresponse of the device at a bias voltage of 5 V and a laser power of 0.5 mW mm−2. (d) A single cycle response of laser on and off.

Electrical measurements were performed in the dark and in the presence of an illuminating laser with different powers ranging from 0.1 mW mm−2 to 2 mW mm−2 (Fig. 5b). The quasilinear IV plots of the device (see inset in Fig. 5b) demonstrate good contact between MoSe2 and the Au electrodes. In contrast to the dark current, the current is significantly increased when the device was illuminated. The IV curves also indicate that the current is strongly dependent on the power of the illuminating laser. At a fixed bias, a larger laser power could lead to a larger current due to the increased number of photon-generated carriers. As a consequence, the photocurrent Iph (Iph = IilluminatedIdark) also increases. Photoresponsivity is a critical factor used to evaluate the performance of photodetectors. For the device produced in this study, the effective exposure area is approximately 446 μm2. At a laser power of 1 mW mm−2 and a bias voltage of 10 V, the photoresponsivity is calculated to be approximately 13 mA W−1, which is comparable to the graphene-based photodetector.59

Photoresponse time is another critical parameter used to judge device performance. We investigated the time-resolved photoresponse of the device by switching the laser on and off. Fig. 5c shows the current of the device as a function of time at a bias voltage of 5 V and a laser power of 0.5 mW mm−2. The device exhibits a repeatable and stable response to the laser illumination. As the laser is switched on and off, the device shows a low off-state current of ∼0.09 nA and a high on-state current of ∼1.9 nA, giving a good on/off ratio of ∼20. Fig. 5d displays a single cycle response of laser on and off. The measured rise and decay times are ∼60 ms, indicating the fast response performance of the device. It should be noted that the performance of the device might be further improved by optimizing the contact or using a phototransistor mode;57,58 further study is currently underway.

4. Conclusion

In conclusion, a large area of highly crystalline MoSe2 atomic layers has been directly synthesized on SiO2/Si, mica and Si substrates using a CVD method at atmospheric pressure without the assistance of H2 gas. Characterization by AFM and Raman confirmed that the thickness of the as-grown MoSe2 ranges from a single layer to a few layers. By using a direct transfer method, we successfully obtained TEM samples. HRTEM together with the SAED pattern demonstrated that the ultrathin MoSe2 flakes are single-crystalline and possess a hexagonal lattice structure. Room-temperature PL indicated that monolayer MoSe2 has a direct bandgap of ∼1.56 eV with a strong emission peak. Photodetectors based on monolayer MoSe2 were fabricated and studied for the first time. The devices exhibited a fast response of ∼60 ms and a good photoresponsivity of 13 mA W−1. We believe that this simple CVD approach can be scaled up to produce wafer-scale ultrathin MoSe2 layers on diverse substrates, expanding the applications of this 2D crystal.

Acknowledgements

We gratefully acknowledge financial support from the “Strategic Priority Research Program” of Chinese Academy of Sciences (Grant no. XDA09040203), 973 Project (2012CB932401), National Natural Science Foundation of China (Grant no. 11174362, 91023041, 61390503 and 91323304) and Knowledge Innovation Project of Chinese Academy of Sciences (Grand no. KJCX2-EW-W02).

Notes and references

  1. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666–669 CrossRef CAS PubMed.
  2. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos and A. A. Firsov, Nature, 2005, 438, 197–200 CrossRef CAS PubMed.
  3. J. Feng, X. Sun, C. Wu, L. Peng, C. Lin, S. Hu, J. Yang and Y. Xie, J. Am. Chem. Soc., 2011, 133, 17832–17838 CrossRef CAS PubMed.
  4. T. Shishidou, A. J. Freeman and R. Asahi, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 180401 CrossRef.
  5. S. Lei, L. Ge, Z. Liu, S. Najmaei, G. Shi, G. You, J. Lou, R. Vajtai and P. M. Ajayan, Nano Lett., 2013, 13, 2777–2781 CrossRef CAS PubMed.
  6. Z. Liu, Y. Gong, W. Zhou, L. Ma, J. Yu, J. C. Idrobo, J. Jung, A. H. MacDonald, R. Vajtai, J. Lou and P. M. Ajayan, Nat. Commun., 2013, 4, 2541 Search PubMed.
  7. M. Langer, M. Kisiel, R. Pawlak, F. Pellegrini, G. E. Santoro, R. Buzio, A. Gerbi, G. Balakrishnan, A. Baratoff, E. Tosatti and E. Meyer, Nat. Mater., 2014, 13, 173–177 CrossRef CAS PubMed.
  8. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman and M. S. Strano, Nat. Nanotechnol., 2012, 7, 699–712 CrossRef CAS PubMed.
  9. D. Jariwala, V. K. Sangwan, L. J. Lauhon, T. J. Marks and M. C. Hersam, ACS Nano, 2014, 8, 1102–1120 CrossRef CAS PubMed.
  10. R. Ganatra and Q. Zhang, ACS Nano, 2014, 8, 4074–4099 CrossRef CAS PubMed.
  11. S. Z. Butler, S. M. Hollen, L. Cao, Y. Cui, J. A. Gupta, H. R. Gutiérrez, T. F. Heinz, S. S. Hong, J. Huang, A. F. Ismach, E. Johnston-Halperin, M. Kuno, V. V. Plashnitsa, R. D. Robinson, R. S. Ruoff, S. Salahuddin, J. Shan, L. Shi, M. G. Spencer, M. Terrones, W. Windl and J. E. Goldberger, ACS Nano, 2013, 7, 2898–2926 CrossRef CAS PubMed.
  12. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis, Nat. Nanotechnol., 2011, 6, 147–150 CrossRef CAS PubMed.
  13. M. Bernardi, M. Palummo and J. C. Grossman, Nano Lett., 2013, 13, 3664–3670 CrossRef CAS PubMed.
  14. R. S. Sundaram, M. Engel, A. Lombardo, R. Krupke, A. C. Ferrari, P. Avouris and M. Steiner, Nano Lett., 2013, 13, 1416–1421 CAS.
  15. Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X. Chen and H. Zhang, ACS Nano, 2012, 6, 74–80 CrossRef CAS PubMed.
  16. H. Liu, K. K. A. Antwi, S. Chua and D. Chi, Nanoscale, 2014, 6, 624–629 RSC.
  17. Y. Chen, J. Xi, D. O. Dumcenco, Z. Liu, K. Suenaga, D. Wang, Z. Shuai, Y.-S. Huang and L. Xie, ACS Nano, 2013, 7, 4610–4616 CrossRef CAS PubMed.
  18. S. Tongay, D. S. Narang, J. Kang, W. Fan, C. Ko, A. V. Luce, K. X. Wang, J. Suh, K. D. Patel, V. M. Pathak, J. Li and J. Wu, Appl. Phys. Lett., 2014, 104, 012101 CrossRef PubMed.
  19. Y. Gong, Z. Liu, A. R. Lupini, G. Shi, J. Lin, S. Najmaei, Z. Lin, A. L. Elias, A. Berkdemir, G. You, H. Terrones, M. Terrones, R. Vajtai, S. T. Pantelides, S. J. Pennycook, J. Lou, W. Zhou and P. M. Ajayan, Nano Lett., 2014, 14, 442–449 CrossRef CAS PubMed.
  20. D. Xiao, G. B. Liu, W. X. Feng, X. D. Xu and W. Yao, Phys. Rev. Lett., 2012, 108, 196802 CrossRef.
  21. H. Zeng, J. Dai, W. Yao, D. Xiao and X. Cui, Nat. Nanotechnol., 2012, 7, 490–493 CrossRef CAS PubMed.
  22. K. F. Mak, K. He, J. Shan and T. F. Heinz, Nat. Nanotechnol., 2012, 7, 494–498 CrossRef CAS PubMed.
  23. J. J. Loferski, J. Appl. Phys., 1956, 27, 777–784 CrossRef CAS PubMed.
  24. S. Tongay, J. Zhou, C. Ataca, K. Lo, T. S. Matthews, J. Li, J. C. Grossman and J. Wu, Nano Lett., 2012, 12, 5576–5580 CrossRef CAS PubMed.
  25. Z. Y. Zhu, Y. C. Cheng and U. Schwingenschlögl, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 153402 CrossRef.
  26. Y. Zhang, T.-R. Chang, B. Zhou, Y.-T. Cui, H. Yan, Z. Liu, F. Schmitt, J. Lee, R. Moore, Y. Chen, H. Lin, H.-T. Jeng, S.-K. Mo, Z. Hussain, A. Bansil and Z.-X. Shen, Nat. Nanotechnol., 2014, 9, 111–115 CrossRef CAS PubMed.
  27. H. S. S. R. Matte, B. Plowman, R. Datta and C. N. R. Rao, Dalton Trans., 2011, 40, 10322–10325 RSC.
  28. S. Larentis, B. Fallahazad and E. Tutuc, Appl. Phys. Lett., 2012, 101, 223104 CrossRef PubMed.
  29. J. S. Ross, S. Wu, H. Yu, N. J. Ghimire, A. M. Jones, G. Aivazian, J. Yan, D. G. Mandrus, D. Xiao, W. Yao and X. Xu, Nat. Commun., 2013, 4, 1474 CrossRef PubMed.
  30. Y. Y. Peng, Z. Y. Meng, C. Zhong, J. Lu, W. C. Yu, Y. B. Jia and Y. T. Qian, Chem. Lett., 2001, 772–773 CrossRef CAS.
  31. J. Pouzet and J. C. Bernede, Rev. Phys. Appl., 1990, 25, 807–815 CrossRef CAS.
  32. D. Kong, H. Wang, J. J. Cha, M. Pasta, K. J. Koski, J. Yao and Y. Cui, Nano Lett., 2013, 13, 1341–1347 CrossRef CAS PubMed.
  33. J. Shaw, H. Zhou, Y. Chen, N. Weiss, Y. Liu, Y. Huang and X. Duan, Nano Res., 2014, 1–7,  DOI:10.1007/s12274-014-0417-z.
  34. X. Lu, M. I. B. Utama, J. Lin, X. Gong, J. Zhang, Y. Zhao, S. T. Pantelides, J. Wang, Z. Dong, Z. Liu, W. Zhou and Q. Xiong, Nano Lett., 2014, 14, 2419–2425 CrossRef CAS PubMed.
  35. X. Wang, Y. Gong, G. Shi, W. L. Chow, K. Keyshar, G. Ye, R. Vajtai, J. Lou, Z. Liu, E. Ringe, B. K. Tay and P. M. Ajayan, ACS Nano, 2014, 8, 5125–5131 CrossRef CAS PubMed.
  36. J. Yoon, W. Park, G.-Y. Bae, Y. Kim, H. S. Jang, Y. Hyun, S. K. Lim, Y. H. Kahng, W.-K. Hong, B. H. Lee and H. C. Ko, Small, 2013, 9, 3295–3300 CAS.
  37. J. Liu, Z. Zeng, X. Cao, G. Lu, L.-H. Wang, Q.-L. Fan, W. Huang and H. Zhang, Small, 2012, 8, 3517–3522 CrossRef CAS PubMed.
  38. Y. Zhou, Y. Nie, Y. Liu, K. Yan, J. Hong, C. Jin, Y. Zhou, J. Yin, Z. Liu and H. Peng, ACS Nano, 2014, 8, 1485–1490 CrossRef CAS PubMed.
  39. H.-Y. Chang, S. Yang, J. Lee, L. Tao, W.-S. Hwang, D. Jena, N. Lu and D. Akinwande, ACS Nano, 2013, 7, 5446–5452 CrossRef CAS PubMed.
  40. O. Lopez-Sanchez, E. Alarcon Llado, V. Koman, A. Fontcuberta I Morral, A. Radenovic and A. Kis, ACS Nano, 2014, 8, 3042–3048 CrossRef CAS PubMed.
  41. L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle, A. Mishchenko, Y. J. Kim, R. V. Gorbachev, T. Georgiou, S. V. Morozov, A. N. Grigorenko, A. K. Geim, C. Casiraghi, A. H. C. Neto and K. S. Novoselov, Science, 2013, 340, 1311–1314 CrossRef CAS PubMed.
  42. M. R. Esmaeili-Rad and S. Salahuddin, Sci. Rep., 2013, 3, 2345 Search PubMed.
  43. W. Zhang, C.-P. Chuu, J.-K. Huang, C.-H. Chen, M.-L. Tsai, Y.-H. Chang, C.-T. Liang, Y.-Z. Chen, Y.-L. Chueh, J.-H. He, M.-Y. Chou and L.-J. Li, Sci. Rep., 2014, 4, 3826 Search PubMed.
  44. W. Regan, N. Alem, B. Alemán, B. Geng, Ç. Girit, L. Maserati, F. Wang, M. Crommie and A. Zettl, Appl. Phys. Lett., 2010, 96, 113102 CrossRef PubMed.
  45. Y.-H. Lee, X.-Q. Zhang, W. Zhang, M.-T. Chang, C.-T. Lin, K.-D. Chang, Y.-C. Yu, J. T.-W. Wang, C.-S. Chang, L.-J. Li and T.-W. Lin, Adv. Mater., 2012, 24, 2320–2325 CrossRef CAS PubMed.
  46. J.-K. Huang, J. Pu, C.-L. Hsu, M.-H. Chiu, Z.-Y. Juang, Y.-H. Chang, W.-H. Chang, Y. Iwasa, T. Takenobu and L.-J. Li, ACS Nano, 2014, 8, 923–930 CrossRef CAS PubMed.
  47. Y. Zhang, Y. Zhang, Q. Ji, J. Ju, H. Yuan, J. Shi, T. Gao, D. Ma, M. Liu, Y. Chen, X. Song, H. Y. Hwang, Y. Cui and Z. Liu, ACS Nano, 2013, 7, 8963–8971 CrossRef CAS PubMed.
  48. S. Wu, C. Huang, G. Aivazian, J. S. Ross, D. H. Cobden and X. Xu, ACS Nano, 2013, 7, 2768–2772 CrossRef CAS PubMed.
  49. Y.-H. Lee, L. Yu, H. Wang, W. Fang, X. Ling, Y. Shi, C.-T. Lin, J.-K. Huang, M.-T. Chang, C.-S. Chang, M. Dresselhaus, T. Palacios, L.-J. Li and J. Kong, Nano Lett., 2013, 13, 1852–1857 CAS.
  50. J. V. Lauritsen, J. Kibsgaard, S. Helveg, H. Topsoe, B. S. Clausen, E. Laegsgaard and F. Besenbacher, Nat. Nanotechnol., 2007, 2, 53–58 CrossRef CAS PubMed.
  51. H. R. Gutierrez, N. Perea-Lopez, A. L. Elias, A. Berkdemir, B. Wang, R. Lv, F. Lopez-Urias, V. H. Crespi, H. Terrones and M. Terrones, Nano Lett., 2013, 13, 3447–3454 CrossRef CAS PubMed.
  52. T. Liang, W. G. Sawyer, S. S. Perry, S. B. Sinnott and S. R. Phillpot, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 77, 104105 CrossRef.
  53. A. C. Ferrari and D. M. Basko, Nat. Nanotechnol., 2013, 8, 235–246 CrossRef CAS PubMed.
  54. P. N. Ghosh and C. R. Maiti, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 28, 2237–2239 CrossRef CAS.
  55. P. A. Bertrand, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 44, 5745–5749 CrossRef CAS.
  56. C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone and S. Ryu, ACS Nano, 2010, 4, 2695–2700 CrossRef CAS PubMed.
  57. S. Das, H.-Y. Chen, A. V. Penumatcha and J. Appenzeller, Nano Lett., 2012, 13, 100–105 CrossRef PubMed.
  58. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic and A. Kis, Nat. Nanotechnol., 2013, 8, 497–501 CrossRef CAS PubMed.
  59. T. Mueller, F. Xia and P. Avouris, Nat. Photonics, 2010, 4, 297–301 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: AFM images and height profile of bilayer and trilayer MoSe2 flakes; A direct transfer method used for preparation of MoSe2 TEM samples. See DOI: 10.1039/c4nr02311k

This journal is © The Royal Society of Chemistry 2014