Chalcogenide Letters Vol. 13, No. 8, August 2016, p.
351 - 357
MORPHOLOGICAL EVOLUTION OF MoS2 NANOSHEETS BY CHEMICAL
VAPOR DEPOSITION
X. WANG, Y. P. ZHANG*, Z. Q. CHEN
Faculty of Materials and Energy, Southwest University, Chongqing 400715,
China
MoS2 nanosheets are synthesized using a low-pressure chemical vapor deposition (CVD)
system at elevated temperature on Si substrates with S and MoO3 power as its precursors.
SEM and XRD measurements indicate that the morphology and structure of the deposited
MoS2 nanosheets are highly dependent on the spatial location of silicon substrate, with the
preferred crystalline orientation varied from (108) to (201) as the sample further from the
sulphur source. XPS shows that the MoO2 constituent increases slightly as the distance
further away from the sulphur source. The nanostructured MoS2 film exhibits strong
photoluminescence in the visible range. This paper may provide new insight into our
understanding the kinetic dynamics related to the MoS2 growth.
(Received June 11, 2016; Accepted August 17, 2016)
Keywords: Transition Metal Dichalcogenides, MoS2, Nanosheets, Chemical Vapor
Deposition
1. Introduction
Layered materials have attracted extensive attention and found promising applications for
sensors, catalysis, and energy storage because of their unique structure with high epecific surface
area and exotic physical and chemical properties.1-2 Besides the zero-bandgap graphene, transition
metal dichalcogenides (TMDs) and transition metal oxides (TMOs) are two important materials
with the two-dimensional layered structure.3-4 Researchers have made every endeavors in
developing reliable and flexible approaches for obtaining thin layer MoS2, and explored a variety
of techniques in synthesizing MoS2 nanosheets, such as exfoliation,5-6 hydrothermal synthesis,7
physical vapor deposition (PVD) and chemical vapor deposition (CVD). 8-13 Among these methods,
chemical vapor deposition has been one of the most promising methods of producing large-area
and high-quality MoS2 thin films and nanomaterials. The precursors and deposition parameters
used in CVD have great affects on the structural mophology of MoS 2 nanosheets. Bulk MoS2 has
an indirect bandgap of 1.2 eV, but for layered MoS2, its bandgap increases as the layer number
decreases, and the monolayer MoS2 has a direct bandgap of 1.9 eV.1-4 Besides bulk and monolayer
MoS2, nanoscale MoS2 has ample property variations by tuning the structural changes and find
new potential applications. Therefore, it is of great importance to achieve the controlled growth of
MoS2 nanosheets and get deeper insight into the reaction mechanism since the process is complex
and related to many factors.
In this paper, we report the effect of the substrate positions on the structural mophologies
of MoS2 nanosheets in a CVD system. Large-area of MoS2 nanosheets were obtained using
low-pressure CVD with the precursors of sulphur and MoO3 powder and characterized by
field-emission scanning electron microscopy (SEM), X-ray diffraction (XRD), X-ray
photoelectron spectroscopy (XPS), Photoluminescence (PL) and Raman spectroscopy.
________________________________________
*
Corresponding author: [email protected]
352
2. Experimental methods
The experimental setup for the synthesis of MoS2 nanosheets are illustrated in Fig. 1. The
quartz tube with smaller diameter (D=20 mm) was used to ensure sufficient reaction of the vapor
of S and MoO3. Two grams of sulphur powder was put in a ceramic boat and placed at the inlet of
CVD furnance 30 cm away from the center, while 50 mg of MoO3 powder was put in another
ceramic boat and located at the centre of furnance. The cleaned four Si substrates (sonication in
acetone, alcohol for 10 min each and subsequent dried by air-blow) were faced up and placed at
different position in the small quartz tube. The insulating plug placed between the MoO3 and the
sulfur can prevent S powder from evaporating too quickly. During the synthesis of MoS2
nanosheets, the temperature of the reaction chamber was slowly heated to 750℃ (20℃/min) and
kept for 60 min and then the furnace was naturally cooled down to room temperature. The process
was carried with an argon environment (100 sccm) at a chamber pressure of 100 Pa.
Fig. 1. Schematic illustration of low-pressure CVD system and the sample position.
The structural mophology of resulting products was observed using field-emission
scanning electron microscopy (FESEM) (JMS-7800F, JEOL Ltd, Japan) and X-ray diffraction
(XRD) (XRD-7000, Shimadzu). The chemical composition and oxidation state were studied using
X-ray photoeletron spectroscopy (XPS) (ESCALAB250Xi, ThermoFisher Ltd). The pass energy
for survey scan is 100 eV, and for Mo 3d and S 2p is 20 eV. The Raman and PL spectra were
obtained by using a Renishaw inVia 2000 micro-Raman system at room temperature in an ambient
air. The excitation laser line is 532 nm.
3. Results and discussion
Figure 2 shows the SEM images of MoS2 nanosheets deposited on the Si(100) substrates
located at different positions. The deposited material, as shown in Fig. 2(a)-(d), is composed of
numerous densely grown, well-faceted, well-distributed, semi-vertically and interleaving lamellar
nanosheets with rough edges. As the distance from the sulphur source increases, the thicknesses of
MoS2 nanosheets gradually decrease from 100-200 nm to 40-60 nm and the lateral sizes from
around 6 m to 1m. The surface edges become rougher with sample further from the sulphur
source. As the insertion shown in Fig. 2(a), the deposition of MoS2 nanosheets is continuous on the
entire Si surface (1 cm2) which shows a high material yield and a controllable nanosheet
morphology.
353
Fig. 2.The SEM images of the MoS2 nanosheets of 4 samples placed at different positions.
(a) Sample 1, (b) Sample 2, (c) Sample 3, and (d) Sample 4 .
XRD was employed to characterize the lattice structure of rusulting products, which is
shown in Fig. 3. Figure 3(a) shows the patterns of MoS2 nanosheets in the range of 10-80°.
Because the the peak intensities differed too widely, Figure 3(b) and (c) give the enlarged
diffraction peaks in the range of 10° to 65° and 65° to 75°, respectively. The XRD patterns show
that the samples are mainly composed of MoS2 and a small amount of MoO2 crystals. The
diffraction peaks (2) positioned at 14.5, 68.8, 69.4, and 69.7 are ascribed to the lattice planes
(002), (200), (108) and (201) of the crystalline 2H-MoS2 (JCPDS card no.75-1539).14 The peaks
(2) at 18.5, 26.1, 37.4, and 57.6 are attributed to the lattice planes (100), (110), (111), and
(300) of MoO2 (JCPDS card no.76-1807). As the distance of the samples from the sulphur source
increases, the peak intensity for MoO2(111) increases slightly. Meanwhile, the peak for MoS2(108)
decreases, and the peak for MoS2(201) increases as the distance from sulphur source increases.
Fig. 3. XRD patterns of the MoS2 nanosheets for Samples 1-4 for the diffraction
angle (2) in the range of (a) 10-80, (b) 10-65, and (c) 65-75.
354
Fig. 4. XPS spectra of the MoS2 nanosheets of Sample 1-4. (a) survey scan,
and the spectra of (b) Mo 3d, (c) S 2p and (d) O 1s.
XPS is a powful technique to provide the composition of the surface region, and
distinguishes the different chemical states of one element. Figure 4 shows the XPS spectra of the
four samples placed at different positions. The survey scan in Fig. 4(a) only shows the peaks
derived from the elements of Mo, S, O, and C. The C 1s signal represents the adventitious carbon.
The ratio of O/Mo increases as the distance of samples from sulphur source increases, which is
consistent with the XRD results. The Mo 3d spectra are shown in Fig. 4(b), in which the small
peaks at binding energy of 226.7 eV are due to the electrons of S 2s corresponding to MoS2.15-16
The peak at 235.8 eV of Mo 3d3/2 for Sample 4 is ascribed to the existence of Mo-O bond of Mo6+,
it is due to the distance of sample 4 from the sulfur source is the greatest and surface oxidation.
The main peaks at binding energy of 229.5 eV and 232.6 eV are attributed to Mo 3d doublet peaks
of 3d5/2 and 3d3/2 which are consistent with Mo4+ in MoS2.15-16 Additionally, the peaks at 162.3 eV
and 163.5 eV in the S 2p spectra in Fig. 4(c) are ascribed to S 2p doublets of 2p3/2 and 2p1/2. In Fig.
4(d), the O 1s core level spectra of the four samples, the weak peaks located at 532.2 eV presence
in a very small portion which is most likely related to physically adsorbed oxygen molecules. The
peaks at the bingding energy of 530.1 eV are caused by O2- oxidation state in MoO2, meanwhile,
as the distance of the samples from the sulphur source increases, the peak intensity for O 1s expect
adsorbed O increases slightly. Moreover, if the sensitive factors for Mo 3d and S 2p are selected as
2.867 and 0.57, respectively, the calculated stoichiometric ratio of S/Mo is approximately 1.76,
which is comparable to the reported values.17-18
355
In the chemical vapor reaction process, MoO3 presumably changes to MoO2 at the sulphur
envrionment. Afterwards, MoO2 is further sulphurized to MoS2 by reacting with the sulphur
vapor.18 Our experiments show the morphology and composition of the resulting MoS2 nanosheets
are related to the S/Mo atomic ration in the reaction atomsphere. As the sample positioned further
to the sulphur source, the MoO2 constituent increases slightly. The sulphur vapor concentration
alone the small quatz tube drops as the sample distance further from the sulphur source. Therefore,
the MoO2 in Sample 1 is fully sulphurized into MoS2, while the MoO2 in Sample 4 is only partially
sulphurized.
These MoS2 nanosheets were further characterized by Raman spectroscopy and
photoluminescence (PL), shown in Fig. 5. In Raman spectra, both A1g and E12g vibrational modes
are associated with out-of-plane vibration of sulfur atoms and in-plane vibration of Mo and S
atoms, respectively.19 The fitting results for Raman spectra in Fig. 5(a) show that these two modes
are located at 382 and 407 cm-1 for the 2H-MoS2 nanosheets in Sample 1, and the frequency
difference is about 25 cm-1 which is confirmed by XRD measurements as well. As for Sample 2,
these two modes are located at 384 and 406 cm-1, respectively, and the frequency difference is
about 22 cm-1. The frequency difference between the two prominent Raman modes (A1g and E12g)
has been known closely related with the layer number of MoS2, therefore, our Raman resultes
indicating the nanosheets consists of many MoS2 layers stacked together. The peak at 365 cm-1 is
attributed to the stretching vibration of molybdyl (Mo=O),20 and its intensity is stronger than that
of Sample 1. That is consistent with the results of XPS.
Fig. 5. Raman (a) and Photoluminescence (b) spectra of the MoS 2 nanosheets
of the two samples placed at different positions.
PL spectra for the MoS2 nanosheets are shown in Fig. 5(b), in which a distinct peak at 671
nm (1.85 eV) is observed. Compared with Sample 1, the PL peak for Sample 2 is about 667
nm.The PL intensity could depend on many parameters such as height, strain, thickness, the
number of layers and defects of nanosheets. It has been reported that the indirect gap bulk MoS2
and MoS2 nanosheets with thickness greater than 5 nm do not exhibit photoluminescence.21
Surprisingly, the as-synthesized MoS2 nanosheets still exhibit PL signal though their thickness are
about 100 nm from the SEM observarion. The PL intensity could depend on many parameters
including thickness, strain, defects, and height of nanosheets. The thickness of the synthesized
MoS2 nanosheets is about 100 nm from the SEM observation, and surprisingly these nanosheets
still exhibit PL signal. The observed PL signal for thick nanosheets may be due to the structural
discontinuity at the nanosheet edges, which induces variation of the electronic structure of MoS2
nanosheets.22
4. Conclusions
Large-area MoS2 nanosheets were successfully synthesized by low-pressure CVD using
sulphur and MoO3 powder as precursors at elevated temperature. The whole sample was covered
with high-quality crystalline MoS2 nanosheets and the stoichiometric ratio of MoS2. The further
356
the distance of sample from the sulphur source, the smaller the width and size of the MoS2
nanosheets. As the distance from sulphur inceases, the preferred crystalline orientation changes to
(210) from (108), and the MoO2 constituent increases slightly. The nanostructured MoS2 film
exhibits strong photoluminescence in the visible range due to its sharp edges. This paper may
provide new insight into our understanding the kinetic dynamics on MoS2 growth.
Acknowledgement
Project supported by the National Natural Sience Foundation of China (Grant No.
21173170) and the Fundamental Research Funds for the Central Universites, China (Grant No.
XDJK2015D001).
References
[1] B. Radisavljevic, A. Radenovic, J. Brivio, et al. Nat. Nanotechnol. 6, 147 (2011).
[2] K. F. Mak, C. G. Lee, J. Hone, et al, Phys. Rev. Lett. 105, 136805 (2010).
[3] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, et al. Nat. Nanotechnol. 7, 699 (2012).
[4] Y. H. Lee, X. Q. Zhang, W. J. Zhang, et al. Adv. Mater. 24, 2320 (2012).
[5] J. N. Coleman, M. Lotya, A. O’Neill, et al. Science 331, 568 (2011).
[6] R. J. Smith, P. J. King, M. Lotya, et al. Adv. Mater. 23, 3944 (2011).
[7] L. J. Ye, H. Y. Xu, D. K. Zhang, et al. Mater. Res. Bull. 55, 221 (2014).
[8] H. Liu, M. W. Si, S. Najmaei, et al. Nano Lett. 13, 2640 (2013).
[9] X. S. Wang, H. B. Feng, Y. M. Wu, et al. J. Am. Chem. Soc. 135, 5304 (2013).
[10] Y. J. Zhan, Z. Liu, S. Najmaei, et al. Small 8, 966 (2012).
[11] Y. F. Yu, C. Li, Y. Liu, et al. Sci. Rep. 3, 1866 (2013).
[12] J. Mann, Q. Ma, P. M. Odenthal, et al. Adv. Mater. 26, 1399 (2014).
[13] X. Ling, Y. H. Lee, Y. X. Lin, et al. Nano Lett. 14, 464 (2014).
[14] W. X. Gu, J. Y. Shen, X. Y. Ma, Nanoscale Res. Lett. 9, 100 (2014).
[15] B. Cho, A. R. Kim, Y. Park, et al. ACS Appl. Mater. Interfaces 7, 2952 (2015).
[16] S. McDonnell, R. Addou, C. Buie, et al. ACS Nano 3, 2880 (2014).
[17] H. T. Lin, X. Y. Chen, H. L. Li, et al. Mater. Lett. 64, 1748 (2010).
[18] D. Q. Gao, M. S. Si, J. Y. Li, et al. Nanoscale Res. Lett. 8, 129 (2013).
[19] X. L. Li, Y. D. Li, Chem. Eur. J. 9, 2726 (2003).
[20] X. Ling, Y. H. Lee, Y. Lin, et al. Nano Lett. 14, 464 (2014).
[21] G. Eda, H. Yamaguchi, D. Voiry, et al. Nano Lett. 11, 5111 (2011).
[22] G. Deokar, D. Vignaud, R. Arenal, P. Louette, J. F. Colomer, Nanotechnology 27, 075604
(2016).