Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

View Article Online

View Journal

PCCP
Physical Chemistry Chemical Physics
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: D. Cho, W. Lee, J.
Wi, W. S. Han, S. J. Yun, B. Shin and Y. Chung, Phys. Chem. Chem. Phys., 2018, DOI:
10.1039/C8CP02390E.

Volume 18 Number 1 7 January 2016 Pages 1–636 This is an Accepted Manuscript, which has been through the
Royal Society of Chemistry peer review process and has been
accepted for publication.
PCCP
Physical Chemistry Chemical Physics Accepted Manuscripts are published online shortly after
www.rsc.org/pccp

acceptance, before technical editing, formatting and proof reading.


Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the


author guidelines.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
ISSN 1463-9076 standard Terms & Conditions and the ethical guidelines, outlined
PERSPECTIVE
in our author and reviewer resource centre, still apply. In no
Darya Radziuk and Helmuth Möhwald
Ultrasonically treated liquid interfaces for progress in cleaning and
separation processes
event shall the Royal Society of Chemistry be held responsible
for any errors or omissions in this Accepted Manuscript or any
consequences arising from the use of any information it contains.

rsc.li/pccp
Page 1 of 9 Physical Chemistry Chemical Physics
View Article Online
DOI: 10.1039/C8CP02390E

PCCP

Physical Chemistry Chemical Physics Accepted Manuscript


PAPER

Enhanced Sulfurization Reaction of Molybdenum Using a Thermal


Cracker for Forming Two-Dimensional MoS2 Layer
Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Received 00th January 20xx,


Accepted 00th January 20xx
Dae-Hyung Cho,*a,b Woo-Jung Lee,a Jae-Hyung Wi,a Won Seok Han,a Sun Jin Yun,a,c Byungha Shin b
DOI: 10.1039/x0xx00000x a,c
and Yong-Duck Chung
www.rsc.org/

We propose a method to fabricate two-dimensional (2D) molybdenum disulfide (MoS2) layers for overcoming issues in
typical fabrication processes by promoting the sulfurization reaction of molybdenum (Mo). A thin sputtered-Mo layer was
sulfurized using a sulfur (S) thermal cracker for forming 2D MoS2 layers. The effects of key process parameters such as
cracking-zone temperature (TC-zone), thickness of sputtered-Mo layer, and Ar pressure during deposition of Mo layer were
systematically investigated. A degree of thermal treatment of evaporated S vapor is controlled by varying the TC-zone. The
higher TC-zone enabled to easily form thin MoS2 layers at low substrate temperature of 250 °C due to the greatly enhanced
sulfurization reaction. The thickness of the final MoS2 layers was controlled by changing the initial thickness of the
sputtered-Mo film. Ultra-thin MoS2 film about 2-layer-thick was obtained by sulfurizing a 2-Å-thick Mo film. The chemical
state of the MoS2 layers largely depended on the Ar pressure during the sputtering process of the initial Mo. Lower Ar
pressure enhanced the MoS2 formation due to more efficient substituting MoS2 phase for MoO3 phase. By using the S
thermal cracker, we demonstrate the method to easily fabricate the 2D MoS2 layer, excluding some problematic issues
such as toxic and expensive reactants, non-vacuum condition susceptible to contamination, and high substrate
temperature.

controllability of flakes, despite of excellent quality. On the


1. Introduction other hand, the typical vapor phase growth, especially that
involves with sulfurization process, employs non- or low-
Nowadays, interest in molybdenum disulfide (MoS2) has been
vacuum equipment (e.g. furnace),27,28 highly toxic H2S gas,29,30
drastically increasing for applications of electronic and
and very extremely high temperatures (over 800 °C).25,31 High
optoelectronic devices because there are indirect-to-direct band-
vacuum process can improve reproducibility of the MoS2
gap transition and extreme mobility improvement by thinning
quality by avoiding contamination species existing in lower
the MoS2 toward monolayer. The monolayer MoS2 transistors
vacuum level. H2S gas is very toxic and poses safety concerns
have been reported showing extremely high carrier mobility.1-5
limiting its use in not only research laboratories but also
The two-dimensional (2D) MoS2 has been also studied for
industrial production. Alternatively thermally evaporated
various applications of optoelectronic devices, such as
elemental source with a carrier gas can be used but the
phototransistors,6-9 sensors,10-12 photovoltaic cells,13,14 and
sulfurization method typically requires the use of high S
water splitting devices.15-17
background pressure, therefore, consuming a huge amount of S
There are several ways of obtaining the 2D MoS2 layers source and carrier gas such as argon or nitrogen. Also, the
including a mechanical exfoliation,18-20 a chemical elevation of substrate temperature raises the manufacturing cost
exfoliation,21-24 and a vapor phase growth.25,26 However, these and induces the performance degradation of some fragile
conventional MoS2 fabrication methods are not without some devices. These drawbacks of the conventional methods are
drawbacks. The exfoliation method is very difficult to apply for mainly from the fact that the molybdenum (Mo) precursors
large-scale fabrication due to limited size and poor thickness (Mo,29,32 MoO3,31,33 MoO2,34 MoCl5,35 or Mo-Au alloy36) tend
to be chemically stable in the presence of S vapor. Therefore,
the formation of MoS2 requires substantial thermal energy (high
temperature), high process pressure, and/or highly reactive gas
such as H2S instead of S. Therefore, a new method to form 2D
MoS2 films that is vacuum compatible, safe, and cost-effective
is highly desirable.

This journal is © The Royal Society of Chemistry 2018 Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 | 1
Physical Chemistry Chemical Physics Page 2 of 9
View Article Online
DOI: 10.1039/C8CP02390E
PAPER PCCP

Physical Chemistry Chemical Physics Accepted Manuscript


Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Figure 1. Schematic illustration of the MoS2 fabrication process. Firstly, the thin Mo films
are sputtered directly onto substrates. Then the Mo films are sulfurized using the S
thermal cracker. The hot cracking zone induces a chemical cracking of the vaporized S
molecules into more activated smaller molecules.

In this work, we suggest a method to form 2D MoS2 layers were < 1.0×10-5 Torr, < 1.0×10-4 Torr, and 10 min, respectively. The
by sulfurizing a thin Mo layer using a thermally cracked S cracker consisted of a reservoir zone (R-zone) and a cracking zone
source. By using highly reactive cracked S species, the method (C-zone) in series. In the R-zone, a typical Knudsen cell, the solid S
alleviates a high thermal budget needed to form MoS2 and also granules (Materion, 99.9%) evaporated to a vapor phase of S at
removes the need of using toxic H2S gas. We analyzed the temperature of 150 °C. The evaporation flux of S can be controlled
effect of the S vapor cracking temperature on the formation of by changing the temperature of R-zone. The evaporated vapor S
MoS2 thin layers at low and medium substrate temperatures went through the hot C-zone heated up to 1000 °C. The C-zone is a
(Tsub) of 250 °C and 550 °C. Also, the dependence of MoS2 quartz tube with a length of about 20 cm surrounded by spiral
properties upon deposition condition of the initial Mo layer was tungsten filaments. The highest TC-zone of 1000 °C was found to be
investigated. The present work offers an interesting finding that the most efficient to make S reactive to metal, based on our previous
controlling the temperature of S vapor heating and the works.37,38 The Tsub of 250 °C and 550 °C were used in this work.
properties of initial Mo layer were important factors for The Tsub were measured using a calibrated thermocouple wafer of
improving the quality of MoS2 thin layers. Mo-coated soda-lime glass (SLG).

2.3. Characterization
2. Experimental The Raman spectroscopic measurement (514 nm, Horiba Jobin
2.1. Mo sputtering Yvon, ARAMIS) revealed the layer number of MoS2 films. The
The 2D MoS2 layers were fabricated in a two-step process including value of peak (E12g)-to-peak (A1g) distance accurately indicated the
Mo sputter-deposition and sulfurization. Firstly, angstrom-scale- tMoS2.39 The average excitation power and the spectral resolution
thick Mo layers were deposited directly on SiO2-terminated were 1.25 mW and 0.9 cm-1, respectively. The optical absorbance
substrates (soda-lime glass or SiO2/Si) using a magnetron DC and reflectance measurement (UV-VIS spectrophotometer, Hitachi
sputtering system (SNTEK). An extremely low deposition rate to U-4001) were carried out to observe the exciton energies40 of the
control the film thickness in a few angstroms was achieved with a fabricated MoS2 and the optical properties of Mo films. The
very low DC power density of 0.6 W/cm2 (50 W was applied to morphology of the MoS2 layer was obtained by atomic force
10.16 cm-diameter target). The Mo thickness was controlled using microscopy (AFM) measurement (Park Systems, XE-100) with the
different deposition times. The 2-, 5-, 10-, 20-, and 30-Å-thick Mo lateral resolution of 30 nm. The thickness, crystallinity, and
films were obtained using the deposition times of 6, 15, 30, 60, and elemental distribution of the MoS2 layers were observed using the
90 sec. In the Mo sputtering process, the base pressure and substrate high-resolution transmission electron microscopy (HRTEM) with an
temperature were < 1.0×10-7 Torr and ambient temperature, energy-dispersive X-ray spectroscopy (EDS) (FEI, Tecnai G2 F30 S-
respectively, using a Mo sputter target (99.95%, Materion, USA). Twin). The chemical bonding states of the MoS2 thin layers and their
phase evolution after the sulfurization were revealed by X-ray
2.2. Sulfurization using S cracker photoelectron spectroscopy (XPS) measurement (ULVAC-PHI, PHI-
The prepared Mo thin layers were sulfurized using a downward S 5000 Versaprobe) with an X-ray beam size of 100 µm2 and a take-
thermal cracker (JMON, SCS-500D) in a vacuum chamber. In the off angle of 90 ° using an electron-neutralizing gun. The sheet
sulfurization process, base pressure, working pressure, and duration resistance of the Mo initial films were measured by a 4-point probe
system (CHANG MIN CO. LTD, CMT-SR200N).

2 | Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 This journal is © The Royal Society of Chemistry 2018
Page 3 of 9 Physical Chemistry Chemical Physics
View Article Online
DOI: 10.1039/C8CP02390E
PCCP PAPER

different C-zone temperatures (TC-zone) of 400 °C or 1000 °C and at


two different Tsub of 250 °C or 550 °C. The higher Tsub presents a
situation where higher thermal energy is directly provided to the Mo
layer during the sulfurization, while the higher TC-zone provides a
larger number of reactive S species onto the Mo surface. Both high

Physical Chemistry Chemical Physics Accepted Manuscript


Tsub and high TC-zone are likely to enhance the Mo-S reaction to form
MoS2 layers. However, since high Tsub may cause the degradation of
underlying materials, lowering Tsub is desired. We investigated if
increasing TC-zone can assist to lower the Tsub. In our previous study,37
the elevation of TC-zone had successfully enhanced the Zn-S reaction
while keeping Tsub the same.
Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Figure 2 shows the Raman spectra of the MoS2 films prepared


Figure 2. The Raman spectra of the MoS2 films prepared using under different Tsub and TC-zone. The thickness of Mo (tMo) was 30 Å.
different Tsub and TC-zone. The Tsub of 250 °C and 550 °C and the The signals of in-plane (E12g) and out-of-plane (A1g) Raman
TC-zone of 400 °C and 1000 °C were used. The peak intensities of vibration modes were clearly observed except the sample fabricated
E12g and A1g largely depended on the Tsub and TC-zone. The E12g at Tsub = 250 °C and TC-zone = 400 °C. The E12g and A1g peaks
and A1g peaks emerged with the elevated TC-zone of 1000 °C when emerged with the elevated TC-zone of 1000 °C even at the lower Tsub
Tsub = 250 °C. At the higher Tsub of 550 °C, the intensity increased
when increasing the TC-zone from 400 °C to 1000 °C. of 250 °C. At the higher Tsub of 550 °C, the Raman intensity
enhanced when the TC-zone increased from 400 °C to 1000 °C, as
anticipated by the enhanced activity of the S. The differences
between the E12g and A1g peaks in few monolayers of MoS2 are
3. Results and discussion known to be under 24 cm-1.39 The measured differences were about
25 cm-1, which indicated that the fabricated MoS2 was in the bulk
The S cracker dissociates the vaporized S molecules to chemically
state. The broad and asymmetric peaks at 460 cm-1 indicate a
active smaller S molecules. The large S molecules such as S8, S7,
superposition of the second-order 2LA(M) vibration mode and first
and S6 are dominant species in S vapor at a relatively low
order optical phonon mode A2u.43,44
temperature of approximately 150 °C; however, the number of
smaller molecules, such as S2, S3, and S4, rapidly increases at
To achieve thinner MoS2 with 2D-scale, the tMo was varied from
temperatures above 500 °C.41,42 Thus, the S vapor, which passes
30 to 2 Å by reducing the Mo sputter deposition time from 90 to 6
through the C-zone, is believed to become very reactive species. In
sec. The values of tMo were derived from the deposition rate
our previous works, the S thermal cracker of the same design
measured from a deposition of thick Mo films. The Mo films on
successfully enhanced the S reaction with a metallic Zn film by
SLG substrates were sulfurized at Tsub = 550 °C and TC-zone = 1000
increasing the temperature of the C-zone.37,38 Based on the same
°C with consideration for the highest Raman intensity in Fig.2. As
technique, sputtered thin Mo layers were sulfurized by a thermal S
shown in Figure 3(a), the Raman spectra with different tMo illustrate
cracker to form MoS2 layers. The schematic of the MoS2 fabrication
that the difference between E12g and A1g peaks was decreased by
process and the S cracker is shown in Figure 1.
decreasing the tMo. The peak frequencies of E12g and A1g and their
differences are shown in Figure 3(b). As tMo decreased from 30 Å to
In order to verify the effect of the S cracking on the MoS2
2 Å, the peak frequency of E12g increased from 383.3 to 386.1 cm-1
formation, the sulfurization processes were carried out at two

Figure 3. The effect of tMo observed by a Raman spectroscopic measurement. The tMo were varied from 30 Å to 2 Å. (a) The Raman spectra
of MoS2 layers with different tMo. (b) The peak positions of E12g and A1g, and their difference value of E12g - A1g as a function of tMo. (c)
The estimated MoS2 thickness at different tMo.

This journal is © The Royal Society of Chemistry 2018 Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 | 3
Physical Chemistry Chemical Physics Page 4 of 9
View Article Online
DOI: 10.1039/C8CP02390E
PAPER PCCP

Physical Chemistry Chemical Physics Accepted Manuscript


Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Figure 4. The effect of tMo observed by an optical absorption measurement. (a) The absorption spectra of
MoS2/SLG samples prepared from different tMo. The absorbance increased as tMo increased. The exciton
peaks of A, B, and C were clearly observed. (b) The differentiated absorbance data for clear assignment
of exciton peak positions of A and B. (c) The corresponding picture of the fabricated MoS2/SLG samples
with different tMo and tMoS2. The MoS2 layers are yellowish and much transparent in thinner MoS2
samples.
while the peak frequency of A1g decreased from 408.4 to 407.4 cm-1. becomes smaller, the band-gap widens because of more dominant
Also the frequency difference decreased from 25.1 to 21.3 cm-1, role of the quantum confinement effect.26,50 The positions of A and B
which indicates that the MoS2 thickness (tMoS2) reduced. Lee et al. exciton peaks were derived by differentiating the absorbance spectra,
reported the correlation between the frequency difference and the as shown in Figure 4(b). The local minima in the wavelengths of the
number of MoS2 layers.39 By comparing our frequency difference differentiated spectra were assumed to be the position of A and B
values to the reported data,39 the estimated numbers of MoS2 layers exciton peaks in the original absorbance spectra. Similar to C
were plotted in Figure 3(c). The thinnest MoS2 of about 2-layers was exciton peak, A and B peaks also exhibited blue shift. When the
achieved using the tMo of 2 Å. By increasing tMo, the tMoS2 linearly tMoS2 reduced from the bulk to 4-layers, the peak position shifted
increased and became almost bulk over tMo of 20 Å. The ratio about ~ 40 and ~ 10 meV in A and B exciton peaks, respectively.
between MoS2 layer numbers and tMo was not constant because the This larger shift of A peaks than B peaks is attributed to that the shift
Mo roughness was believed to change with tMo. of the A exciton peak originated only from the increase of the first
exciton level while the valence band splitting energy caused the B
UV-Visible spectrophotometric analysis was carried out to exciton to barely change.40 Although the A and B exciton peaks were
observe the absorption spectra of the MoS2 films with various very weak for tMoS2 of 2-layers and 3-layers, both peaks seemed to
thicknesses, as shown in Figure 4(a). The tMoS2 estimated from shift to a higher energy (lower wavelength) than the thicker MoS2.
Figure 3(c) is denoted in the legends of Figure 4(a) and (b). The The results of the peak shift around 1.9 eV matched well with the
exciton resonance peaks of A, B, and C appeared at around 660 nm, shift of photoluminescence of A peaks in a previous work; the
620 nm, and 420 nm, respectively.45 The overall absorbance measured absorption resonances correlated with the direct-gap hot
decreased in the entire wavelength range as the tMoS2 decreased. The luminescence.51 Figure 4(c) presents photographs of MoS2/SLG
C exciton peak showed blue shift from 430 nm (~ 2.88 eV) to 390 samples with different tMo. The samples become more transparent
nm (~ 3.18 eV) with decreasing tMoS2. The C peaks at around 2.8 eV with thinner tMo due to the absorbance reduction as shown in Figure
are known to be due to the transition in nesting bands between the Γ 4(a).
and Λ positions of the Brillouin zone.40,45,46 Although the C peaks
locate at slightly higher energy compared to previous works,45 the To improve the quality of MoS2 layers, the Ar pressure (PAr,Mo)
tendency of blue-shift with decreasing tMoS2 matches well with other during the sputtering process of the initial Mo metal was varied from
reports.47,48 The slightly higher energies of C peaks than previous 40 to 4 mTorr. The PAr,Mo used in the samples of Figures 2 – 4 was
works seemed to be originated from the imperfect film quality due to 40 mTorr. Upon varying PAr,Mo, the deposition time was adjusted to
the residual structural disorder with poor crystallinity as compared to result in the same tMo because the deposition rate varies with PAr,Mo.
the pristine mechanically exfoliated monolayers.47,49 When the tMoS2 The deposition time of Mo were about 15, 13, and 11 sec at PAr,Mo of

4 | Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 This journal is © The Royal Society of Chemistry 2018
Page 5 of 9 Physical Chemistry Chemical Physics
View Article Online
DOI: 10.1039/C8CP02390E
PCCP PAPER

Physical Chemistry Chemical Physics Accepted Manuscript


Figure 5. (a) The Raman spectra of the MoS2 layers prepared with different PAr,Mo. The higher PAr,Mo of 4 and 10 mTorr greatly enhanced the
Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Raman intensity than the PAr,Mo of 40 mTorr. (b) The AFM image of MoS2/SiO2/Si sample prepared with PAr,Mo of 4 mTorr. The line profile
across the boundary with SiO2 (left dark region)- and MoS2 (right bright region)-terminated surfaces is shown in inset. (c, d) The cross-
sectional HRTEM image of MoS2 layer prepared at 4 mTorr. (e) The corresponding EDS elemental mapping image of Mo Kα (red) and S
Kα (green) dots.

40, 10, and 4 mTorr, respectively. The TC-zone and Tsub were 1000 °C the MoO2 at ~230 eV), (ii) Mo5+ (assigned to Mo2O5 at ~231.5 eV),
and 550 °C, respectively, during the sulfurization. Figure 5(a) and (iii) Mo6+ (derived from MoO3 at ~232.9 eV).52-54 The chemical
displays the Raman spectra of MoS2 layers formed with different states of Mo layer are comprised of the most Mo6+ states of
PAr,Mo of 40, 10, and 4 mTorr. The peak intensities of 4 and 10 approximately 99%; i.e., the MoO3 phase seemed to be natively
mTorr samples were greatly higher than that of 40 mTorr, formed after the Mo deposition due to the air exposure prior to the
maintaining peak positions of E12g and A1g frequencies. It means that sulfurization reaction as proposed in equation (1), while the peaks of
the decrease of PAr,Mo dramatically improves the formation of 2D- MoO2 and Mo2O5 states were negligible.55
MoS2 layer. An AFM image of the MoS2 layer with PAr,Mo of 4
mTorr is shown in Figure 5(b). The MoS2-coated area was easily 2Mo(s) + 3O2(g) → 2MoO3(s) (1)
distinguished from the edge area on which Mo metal was not
deposited. The line profile across the both area indicated the tMoS2 After sulfurization, the chemical states in MoS2 layer depending
was about 23 Å which corresponded to about 3 layers.39 The on PAr,Mo were investigated by fitting the Mo 3d peak as shown in
thickness measured by AFM agrees well with the result of the Figure 6(b). There are three different Mo states in MoS2 layer: Mo-I
Raman peak difference of about 22.7 cm-1 which also corresponded (~ 229.6 eV), Mo-II (~ 231.0 eV), and Mo-III (~ 232.8 eV),
to the thickness about 3 layers. The AFM image and the line profile corresponding to the MoS2, MoS3, and MoO3, respectively.56-58 With
indicate the MoS2 films consists of some grain or island structures decreasing PAr,Mo, the MoS2 state (Mo-I) was dominantly formed
with an insufficient lateral homogeneity. This was likely attributed to while the MoO3 (Mo-III) and MoS3 (Mo-II) states became lesser. As
the fact that the initial thin Mo was grown in island structures rather compared to the Mo 3d of Mo layer, the three 3d5/2 peak positions in
than in layer-by-layer structures due to the surface energy difference MoS2 layer shifted toward lower binding energy due to the lower
between the Mo and the underlying SiO2 substrate. The cross- bonding strength of Mo-S than Mo-O,52 which clearly demonstrates
sectional images of the MoS2 layer were obtained by a HRTEM as that the sulfurization process successfully converted the MoO3 phase
shown in Figure 5(c) and (d). The 2.5-nm-thick continuous MoS2 into the MoS2 phase especially at lower PAr,Mo of 4 mTorr. Notable
atomic layer parallel to the SiO2 substrate was observed. The fine thing is the presence of two S states of S-I (~ 226.5 eV) and S-II (~
crystalline microstructure was hardly seen which was attributed to 229.5 eV) arisen from the S in MoS2 and elemental S, respectively.52
the relatively poor crystallinity of the MoS2 layers as discussed the The peaks of S-I and S-II were derived from the S 2s orbital. With
absorption results. The well-defined MoS2 layer was verified by the decreasing PAr,Mo, both states of S-I and S-II were increased. It is
EDS elemental color mapping of Mo and S. The Mo signal was expected result of the generation of the S-I state with the formation
relatively weak because the Mo Kα was used due to the same of MoS2, whereas the S-II state has been rarely reported in MoS2
characteristic X-ray energies of Mo Lα and S Kα. layers. For more detailed analysis, the Mo and S chemical states are
quantified using relative concentration values obtained from the
To investigate the effect of PAr,Mo on formation of MoS2 layer, the fitting results, as shown in Figure 6(c). The concentration of the
chemical states of the bare Mo layer and MoS2 layers grown at MoS2 states drastically increased from 17.5% to 85.7% whereas the
different PAr,Mo of 40, 10, and 4 mTorr were identified by XPS. MoO3 phase decreased from 81.9% to 14.2%, as decreasing PAr,Mo
Figure 6(a) shows the fitted Mo 3d spectra of the Mo layer before from 40 mTorr to 4 mTorr. The MoS3 concentration showed a very
sulfurization. In 3d orbital, clearly distinguishable two peaks are low values near zero with little change. Based on these fitting
detected with 3d5/2 and 3d3/2 due to the spin orbit splitting; the peak results, we proposed a possible processes as follows in equation (2)
intensity ratio of 3d5/2-to-3d3/2 is about 3/2 and peak separation – (4).
energy of 3d5/2-to-3d3/2 is approximately 3.2 eV.52 In the fitting 2MoO3(s) + S(g) → 2MoO2(s) + SO2(g) (2)
results of Mo 3d spectra, there was no elemental Mo 3d5/2 metal peak
MoO2(s) + 2S(g) → MoS2(s) + O2(g) (3)
around 228 eV but three different Mo oxide states were observed
(binding energy was read from the 3d5/2 peak); (i) Mo4+ (related to MoO2(s) + 3S(g) → MoS2(s) + SO2(g) (4)

This journal is © The Royal Society of Chemistry 2018 Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 | 5
Physical Chemistry Chemical Physics Page 6 of 9
View Article Online
DOI: 10.1039/C8CP02390E
PAPER PCCP

Physical Chemistry Chemical Physics Accepted Manuscript


Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Figure 6. The XPS spectra of bare Mo layer before sulfurization and MoS2 layers
prepared with different PAr,Mo. (a) The fitted Mo 3d spectra of pure Mo layer consisted of
Mo4+, Mo5+ and Mo6+. (b) The variation of Mo 3d spectra of MoS2 layer as a function of
PAr,Mo, fitted with different states of Mo (Mo-I: MoS2, Mo-II: MoS3, Mo-III: MoO3) and S
(S-I: S in MoS2, S-II: elemental S). (c) Relative concentration of various Mo (solid line)
and S (dashed line) phases distributed in MoS2 layer as a function of PAr,Mo..

The MoO2 phase was formed first by reduction of MoO3 during the 1T-phase MoS2 exhibits dominant J1 peak (~ 153 cm-1) in Raman
the sulfurization (2).34,59 Then the MoO2 was likely to be converted spectra.61-63 In the Raman spectra, however, the J1 peak was not
to MoS2 by a single displacement reaction (3)34 or a double observed (not shown). The formation of 2H-phase-dominant MoS2
displacement reaction (4).30,59 The highly reactive S vapor promoted layers were likely attributed to that the 2H phase is more
the S molecules to substitute the oxygen in MoO3 for the MoS2 thermodynamically stable than the 1T or 3R phases at the
formation. Interesting findings is the dependence of chemical temperature used in this proposed process.47,64,65
reaction on PAr,Mo of Mo layer during the sulfurization using the S To observe the optical and electrical properties of initial Mo layer
thermal cracker. As PAr,Mo is decreased, the S atoms seems to be depending on PAr,Mo, thin Mo films were deposited on the SiO2/Si
more actively incorporated into the Mo layer, resulting in the substrates. Because the angstrom-scale-thick Mo used for the MoS2
transformation from MoO3 into the MoS2. At higher PAr,Mo of 40 fabrication was too thin to characterize and distinguish their
mTorr, however, the diffusion of S molecules into the Mo layer is properties, the thicker 20-nm-thick Mo films were prepared for clear
impeded, thus an amount of unreacted S atoms exists with the high property difference between each sample. Figure 7(a) shows the
intensity of S-II states. Moreover, considering the substantial S-II reflectance of the Mo films formed by different PAr,Mo and figure
states at every PAr,Mo, we assumed that the supplied S molecules 7(b) illustrates their sheet resistance (Rsheet) and roughness kurtosis
were too abundant to react Mo atoms due to the very thin Mo layer, (Rku). The reflectance decreased as increasing PAr,Mo in the full range
and thus, unreacted S atoms remained in the MoS2 layer. of wavelength. The Rsheet, measured by a 4-point probe, dramatically
The most MoS2 thin layers are known to show trigonal prismatic 2H increased as increasing PAr,Mo; about 10 times increased from 4
or octahedral 1T phases depending on their fabrication method and mTorr to 10 mTorr and then about 10,000 times increased from 10
post-treatment. We concluded that there are much dominant 2H mTorr to 40 mTorr. Moreover, the Rku, observed by an AFM,
phase in the fabricated MoS2 considering the XPS and Raman results. increased from 3.3 to 3.9 as increasing PAr,Mo from 4 mTorr to 40
As shown in Fig. 6(b), the binding energy of the fitted Mo4+ 3d5/2 mTorr. The Rku describes the sharpness of peaks on the film
peaks of MoS2 phases are about 229 eV and the peaks are sharp morphologies: the higher Rku indicates that the Mo surface is
without any shoulder at lower binding energy. The Mo4+ 3d5/2 is spikier.66 The measured Rku represent that the higher PAr,Mo induced
generally observed at 229 eV in a 2H phase, whereas the Mo4+ 3d5/2 spikier and sharper shapes of Mo film surfaces. Thus, the Mo films
binding energy of the 1T phase is known to be about 228 eV which produced at higher PAr,Mo were believed to reduce the reflected light
is about 1 eV lower than the 2H phase. If there are some mixed in direction of perpendicular to the surface (lower reflectance) and
phase of 2H and 1T, the peak shows doublet peaks.47,60 Moreover, much lengthen the electrical paths (higher Rsheet). These results are

6 | Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 This journal is © The Royal Society of Chemistry 2018
Page 7 of 9 Physical Chemistry Chemical Physics
View Article Online
DOI: 10.1039/C8CP02390E
PCCP PAPER

Physical Chemistry Chemical Physics Accepted Manuscript


Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

Figure 7. (a) The reflectance spectra of the Mo/SiO2/Si samples with different PAr,Mo
when the incident light was perpendicular to the Mo surfaces. (b) The sheet resistance
(Rsheet) and the roughness kurtosis (Rku) of the Mo/SiO2/Si samples as a function of PAr,Mo,
and (c) the corresponding pictures of the samples (c).

likely to be due to the fact that the Mo films at lower PAr,Mo


contained fewer voids and are formed more densely with larger
crystallite onto substrates because the kinetic energy, hence surface Acknowledgements
mobility of Mo adatoms increases at a lower working pressure.67,68 This work was supported by Institute for Information &
Thus, the higher quality Mo film sputtered at lower PAr,Mo is seemed communications Technology Promotion (IITP) grant funded by
to enhance the MoS2 formation with remaining less MoO3 and the Korea government (MSIP) (2016-0-00576, Fundamental
elemental S, as shown in figure 6(c). Figure 7(c) displays the picture technologies of two-dimensional materials and devices for the
of the corresponding samples with different colors due to the platform of new-functional smart devices).
reflectance variation.

References
4. Conclusions 1 X. Cui, G.-H. Lee, Y. D. Kim, G. Arefe, P. Y. Huang, C.-H.
The proposed sulfurization method successfully enabled to Lee, D. A. Chenet, X. Zhang, L. Wang, F. Ye, F.
form MoS2 thin layers. The sputtered Mo films were sulfurized Pizzocchero, B. S. Jessen, K. Watanabe, T. Taniguchi, D.
using the thermal S cracker. The 2D MoS2 layers were easily A. Muller, T. Low, P. Kim and J. Hone, Nat Nano, 2015,
10, 534.
formed at relatively low substrate temperatures of 250°C – 550
2 D. Krasnozhon, D. Lembke, C. Nyffeler, Y. Leblebici and
°C using the cracking-zone temperature of 1000 °C. As A. Kis, Nano Lett., 2014, 14, 5905.
increasing the cracking-zone temperature, the Raman intensity 3 S. Das, H.-Y. Chen, A. V. Penumatcha and J. Appenzeller,
increased at the same substrate temperature. The layer number Nano Lett., 2013, 13, 100.
of MoS2 was able to be controlled by changing the initial Mo 4 W. Bao, X. Cai, D. Kim, K. Sridhara and M. S. Fuhrer,
film thickness. The 2-layers of MoS2 was obtained using 2-Å- Appl. Phys. Lett., 2013, 102, 042104.
5 H. Liu and P. D. Ye, IEEE Electron Device Lett., 2012, 33,
thick Mo film. Moreover, the quality of the MoS2 layers largely
546.
depended on the argon working pressure in the Mo sputter- 6 W. Zhang, J.-K. Huang, C.-H. Chen, Y.-H. Chang, Y.-J.
deposition. Lower Mo working pressure enhanced the MoS2 Cheng and L.-J. Li, Adv. Mater., 2013, 25, 3456.
formation via more efficient substituting MoS2 phase for MoO3 7 O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic
phase. Studied MoS2 fabrication process is production- and A. Kis, Nat Nano, 2013, 8, 497.
favorable because it includes non-toxic and cheap materials, 8 Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q.
low substrate temperature, and vacuum condition with little Zhang, X. Chen and H. Zhang, ACS Nano, 2012, 6, 74.
9 H. S. Lee, S.-W. Min, Y.-G. Chang, M. K. Park, T. Nam,
contamination. We found a possibility to improve the MoS2 H. Kim, J. H. Kim, S. Ryu and S. Im, Nano Lett., 2012, 12,
quality by investigating the dependence of Mo properties and 3695.
also the S vapor cracking temperature. The proposed method is 10 H. Li, Z. Yin, Q. He, H. Li, X. Huang, G. Lu, D. W. H.
expected to be easily adapted to other 2D transition metal Fam, A. I. Y. Tok, Q. Zhang and H. Zhang, Small, 2012, 8,
dichalcogenides by varying the sputtered transition metals and 63.
evaporated chalcogen vapors. 11 D. Sarkar, W. Liu, X. Xie, A. C. Anselmo, S. Mitragotri
and K. Banerjee, ACS Nano, 2014, 8, 3992.
12 D. J. Late, Y.-K. Huang, B. Liu, J. Acharya, S. N.

This journal is © The Royal Society of Chemistry 2018 Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 | 7
Physical Chemistry Chemical Physics Page 8 of 9
View Article Online
DOI: 10.1039/C8CP02390E
PAPER PCCP

Shirodkar, J. Luo, A. Yan, D. Charles, U. V. Waghmare, V. Reports, 2013, 3, 1866.


P. Dravid and C. N. R. Rao, ACS Nano, 2013, 7, 4879. 36 I. Song, C. Park, M. Hong, J. Baik, H.-J. Shin and H. C.
13 M.-L. Tsai, S.-H. Su, J.-K. Chang, D.-S. Tsai, C.-H. Chen, Choi, Angew. Chem. Int. Ed., 2014, 53, 1266.
C.-I. Wu, L.-J. Li, L.-J. Chen and J.-H. He, ACS Nano, 37 D.-H. Cho, W.-J. Lee, S.-W. Park, J.-H. Wi, W. S. Han, J.
2014, 8, 8317. Kim, M.-H. Cho, D. Kim and Y.-D. Chung, J. Mater.

Physical Chemistry Chemical Physics Accepted Manuscript


14 X. Gu, W. Cui, H. Li, Z. Wu, Z. Zeng, S.-T. Lee, H. Zhang Chem. A, 2014, 2, 14593.
and B. Sun, Advanced Energy Materials, 2013, 3, 1262. 38 D.-H. Cho, W.-J. Lee, J.-H. Wi, W. S. Han, T. G. Kim, J.
15 Y.-J. Yuan, J.-R. Tu, Z.-J. Ye, D.-Q. Chen, B. Hu, Y.-W. W. Kim and Y.-D. Chung, ETRI J., 2016, 38, 265.
Huang, T.-T. Chen, D.-P. Cao, Z.-T. Yu and Z.-G. Zou, 39 C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone and S.
Applied Catalysis B: Environmental, 2016, 188, 13. Ryu, ACS Nano, 2010, 4, 2695.
16 Y.-J. Yuan, D. Chen, J. Zhong, L.-X. Yang, J. Wang, M.-J. 40 K. P. Dhakal, D. L. Duong, J. Lee, H. Nam, M. Kim, M.
Liu, W.-G. Tu, Z.-T. Yu and Z.-G. Zou, J. Mater. Chem. A, Kan, Y. H. Lee and J. Kim, Nanoscale, 2014, 6, 13028.
Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

2017, 5, 15771. 41 H. Rau, T. R. N. Kutty and J. R. F. Guedes De Carvalho, J.


17 Y.-J. Yuan, Z.-J. Ye, H.-W. Lu, B. Hu, Y.-H. Li, D.-Q. Chem. Thermodynamics, 1973, 5, 833.
Chen, J.-S. Zhong, Z.-T. Yu and Z.-G. Zou, ACS Catalysis, 42 S.-J. Park, Y.-D. Chung, W.-J. Lee, D.-H. Cho, J.-H. Wi,
2016, 6, 532. W.-S. Han, Y. Cho and J.-m. Yoon, J. Korean Phys. Soc.,
18 K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. 2015, 66, 76.
Khotkevich, S. V. Morozov and A. K. Geim, Proc. Natl. 43 B. R. Carvalho, Y. Wang, S. Mignuzzi, D. Roy, M.
Acad. Sci. U. S. A., 2005, 102, 10451. Terrones, C. Fantini, V. H. Crespi, L. M. Malard and M. A.
19 M. M. Benameur, B. Radisavljevic, J. S. Héron, S. Sahoo, Pimenta, Nature Communications, 2017, 8
H. Berger and A. Kis, Nanotechnology, 2011, 22, 125706. 44 F. Bozheyev, D. Valiev and R. Nemkayeva, J. Lumin.,
20 A. Castellanos-Gomez, M. Barkelid, A. M. Goossens, V. E. 2017, 188
Calado, H. S. J. van der Zant and G. A. Steele, Nano Lett., 45 D. Kozawa, R. Kumar, A. Carvalho, K. Kumar Amara, W.
2012, 12, 3187. Zhao, S. Wang, M. Toh, R. M. Ribeiro, A. H. Castro Neto,
21 J. N. Coleman, M. Lotya, A. O’Neill, S. D. Bergin, P. J. K. Matsuda and G. Eda, Nature Communications, 2014, 5,
King, U. Khan, K. Young, A. Gaucher, S. De, R. J. Smith, 4543.
I. V. Shvets, S. K. Arora, G. Stanton, H.-Y. Kim, K. Lee, 46 A. Carvalho, R. M. Ribeiro and A. H. Castro Neto, Phys.
G. T. Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Rev. B, 2013, 88, 115205.
Wang, J. F. Donegan, J. C. Grunlan, G. Moriarty, A. 47 G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen and
Shmeliov, R. J. Nicholls, J. M. Perkins, E. M. Grieveson, M. Chhowalla, Nano Lett., 2011, 11, 5111.
K. Theuwissen, D. W. McComb, P. D. Nellist and V. 48 Y. Yu, Y. Yu, Y. Cai, W. Li, A. Gurarslan, H. Peelaers, D. E.
Nicolosi, Science, 2011, 331, 568. Aspnes, C. G. Van de Walle, N. V. Nguyen, Y.-W. Zhang
22 A. Jawaid, D. Nepal, K. Park, M. Jespersen, A. Qualley, P. and L. Cao, Scientific Reports, 2015, 5, 16996.
Mirau, L. F. Drummy and R. A. Vaia, Chem. Mater., 2016, 49 S. Jiménez Sandoval, D. Yang, R. F. Frindt and J. C. Irwin,
28 Phys. Rev. B, 1991, 44
23 N. Liu, P. Kim, J. H. Kim, J. H. Ye, S. Kim and C. J. Lee, 50 A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim,
ACS Nano, 2014, 8, 6902. G. Galli and F. Wang, Nano Lett., 2010, 10, 1271.
24 E. Varrla, C. Backes, K. R. Paton, A. Harvey, Z. 51 K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Phys.
Gholamvand, J. McCauley and J. N. Coleman, Chem. Rev. Lett., 2010, 105, 136805.
Mater., 2015, 27 52 H. W. Wang, P. Skeldon and G. E. Thompson, Surf. Coat.
25 C.-R. Wu, X.-R. Chang, T.-W. Chu, H.-A. Chen, C.-H. Wu Technol., 1997, 91
and S.-Y. Lin, Nano Lett., 2016, 16, 7093. 53 A. Borgschulte, O. Sambalova, R. Delmelle, S. Jenatsch,
26 L. Hongfei, K. K. A. Antwi, Y. Jifeng, C. Soojin and C. R. Hany and F. Nüesch, Scientific Reports, 2017, 7
Dongzhi, Nanotechnology, 2014, 25, 405702. 54 Irfan, F. So and Y. Gao, International Journal of
27 S. Wu, C. Huang, G. Aivazian, J. S. Ross, D. H. Cobden Photoenergy, 2011, 2011
and X. Xu, ACS Nano, 2013, 7, 2768. 55 NIST X-Ray Photoelectron Spectroscopy Database,
28 M. H. Heyne, D. Chiappe, J. Meersschaut, T. Nuytten, T. http://srdata.nist.gov/xps/, (accessed April 2018).
Conard, H. Bender, C. Huyghebaert, I. P. Radu, M. 56 M. A. Baker, R. Gilmore, C. Lenardi and W. Gissler, Appl.
Caymax, J. F. de Marneffe, E. C. Neyts and S. De Gendt, Surf. Sci., 1999, 150
Journal of Materials Chemistry C, 2016, 4, 1295. 57 J. G. Choi and L. T. Thompson, Appl. Surf. Sci., 1996, 93
29 Y. Lee, J. Lee, H. Bark, I.-K. Oh, G. H. Ryu, Z. Lee, H. 58 W. C. Moshier, G. D. Davis and G. O. Cote, J.
Kim, J. H. Cho, J.-H. Ahn and C. Lee, Nanoscale, 2014, 6, Electrochem. Soc., 1989, 136
2821. 59 X. L. Li and Y. D. Li, Chemistry – A European Journal,
30 P. Kumar, M. Singh, R. K. Sharma and G. B. Reddy, J. 2003, 9, 2726.
Alloys Compd., 2016, 671, 440. 60 A. Ambrosi, Z. Sofer and M. Pumera, Chem. Commun.,
31 Y.-C. Lin, W. Zhang, J.-K. Huang, K.-K. Liu, Y.-H. Lee, 2015, 51, 8450.
C.-T. Liang, C.-W. Chu and L.-J. Li, Nanoscale, 2012, 4, 61 D. Wang, X. Zhang, S. Bao, Z. Zhang, H. Fei and Z. Wu,
6637. J. Mater. Chem. A, 2017, 5, 2681.
32 Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan and J. Lou, 62 C. Guo, J. Pan, H. Li, T. Lin, P. Liu, C. Song, D. Wang, G.
Small, 2012, 8, 966. Mu, X. Lai, H. Zhang, W. Zhou, M. Chen and F. Huang, J.
33 Y.-H. Lee, X.-Q. Zhang, W. Zhang, M.-T. Chang, C.-T. Mater. Chem. C, 2017, 5, 10855.
Lin, K.-D. Chang, Y.-C. Yu, J. T.-W. Wang, C.-S. Chang, 63 H. Wang, C. Zhang and F. Rana, Nano Lett., 2015, 15, 339.
L.-J. Li and T.-W. Lin, Adv. Mater., 2012, 24, 2320. 64 C. Kun, H. Xiao, P. Hong, Z. Huabin, S. Li, L. Guigao, L.
34 X. Wang, H. Feng, Y. Wu and L. Jiao, J. Am. Chem. Soc., Huimin, Z. Guixia, L. Mu and Y. Jinhua, Adv. Mater.,
2013, 135, 5304. 2016, 28, 10033.
35 Y. Yu, C. Li, Y. Liu, L. Su, Y. Zhang and L. Cao, Scientific 65 J. A. Wilson and A. D. Yoffe, Adv. Phys., 1969, 18, 193.

8 | Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 This journal is © The Royal Society of Chemistry 2018
Page 9 of 9 Physical Chemistry Chemical Physics
View Article Online
DOI: 10.1039/C8CP02390E
PCCP PAPER

66 B. Rajesh Kumar and T. Subba Rao, Digest Journal of


Nanomaterials and Biostructures, 2012, 7, 1881.
67 J.-H. Yoon, S. Cho, W. M. Kim, J.-K. Park, Y.-J. Baik, T.
S. Lee, T.-Y. Seong and J.-h. Jeong, Sol. Energy Mater.
Sol. Cells, 2011, 95, 2959.

Physical Chemistry Chemical Physics Accepted Manuscript


68 H. Khatri and S. Marsillac, J. Phys.: Condens. Matter,
2008, 20, 055206.
Published on 23 May 2018. Downloaded by Kings College London on 25/05/2018 12:29:46.

This journal is © The Royal Society of Chemistry 2018 Phys. Chem. Chem. Phys., D.-H. Cho et al., 2018, 00, 1-3 | 9

You might also like