Monolayer Behaviour in Bulk Res Due To Electronic and Vibrational Decoupling
Monolayer Behaviour in Bulk Res Due To Electronic and Vibrational Decoupling
Monolayer Behaviour in Bulk Res Due To Electronic and Vibrational Decoupling
Received 27 Aug 2013 | Accepted 13 Jan 2014 | Published 6 Feb 2014 DOI: 10.1038/ncomms4252
1 StateKey Laboratory of Superlattices and Microstructures, Institute of Semiconductors, Chinese Academy of Sciences, P.O. Box 912, Beijing 100083,
People’s Republic of China. 2 Department of Materials Science and Engineering, University of California, Berkeley, California 94720, USA. 3 Department of
Physics, University of Antwerp, Groenenborgerlaan 171, Antwerpen B-2020, Belgium. 4 Department of Thermal Science and Energy Engineering, University of
Science and Technology of China, Anhui 230027, China. 5 Department of Electronic Engineering, National Taiwan University of Science and Technology, Taipei
106, Taiwan. 6 Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. 7 Molecular Foundry, Lawrence Berkeley
National Laboratory, Berkeley, California 94720, USA. 8 Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.
Correspondence and requests for materials should be addressed to J.W. (email: [email protected]).
A
tomically thin monolayer transition metal dichalcogen- DFT calculations show that the hexagonal (H) phase is unstable,
ides (sTMDs) are a new class of two-dimensional (2D) whereas the octahedral (T) phase goes through the Peierls
materials with the chemical formula MX2, where M is a distortion, resulting in buckled S layers and zigzag Re chains along
transition metal (Mo, W and Re) and X is a chalcogen (S, Se and one of the lattice vectors in the plane (Fig. 1a bottom,
Te) element. Monolayer sTMDs are promising functional Supplementary Fig. 1). The resultant lattice in Fig. 1a can be
materials for next-generation flexible optoelectronics and photo- regarded as a distorted 1T structure (Supplementary Notes 1 and 2).
voltaics applications owing to their mechanical flexibility, Owing to the reduced crystal symmetry, ReS2 displays a more
chemically and environmentally stability, optical properties and complex Raman spectrum than conventional sTMDs, as will be
low operating voltages in various device configurations1–3. One of discussed later. DFT calculations reveal that bulk and monolayer
the landmark features of sTMDs is that they undergo a crossover ReS2 have nearly identical band structures (Fig. 1b), both being
from indirect bandgap in the bulk to direct bandgap in direct-bandgap semiconductors with a predicted generalized gra-
monolayers4, and as a result, monolayer sTMDs become direct- dient approximation (GGA) bandgap at Egap ¼ 1.35 eV (bulk) and
gap semiconductors that absorb and emit light efficiently. 1.43 eV (monolayer), respectively (Supplementary Note 3). This is
Currently, this crossover has been observed in all members of in stark contrast to conventional sTMDs, where the band structure
the sTMDs family known in the field, and is attributed to a strong is strongly dependent on the number of layers4,5. Our DFT
interlayer coupling and explained as a confinement effect in the calculations (Fig. 1c) also show that adjacent layers in ReS2 are
thickness direction. indeed only weakly coupled to each other, with a coupling energy
Here we present a new member of the sTMDs family, rhenium ofB18 meV per unit cell, o8% of that of MoS2 (460 meV for the
disulphide (ReS2), in which the band renormalization is absent corresponding 2 2 conventional cell).
and the bulk behaves as electronically and vibrationally decoupled
monolayers. Such unique response originates from lack of
interlayer registry and weak interlayer coupling arising from HRTEM measurements and determination of crystal structure.
Peierls distortion of the 1T structure of ReS2, as confirmed by To study the structure and define the stacking order of ReS2,
high-resolution transmission electron microscopy (HRTEM), we deployed extensive HRTEM and electron diffractometry
electron diffraction and density functional theory (DFT) calcula- measurements. In Fig. 2a, we display the HRTEM image taken
tions. Consequently, from bulk to monolayers, ReS2 remains a from two ReS2 flakes, one with the layers perpendicular (red
direct-bandgap semiconductor, its photoluminescence (PL) square) and the other parallel (green square) to the electron
intensity increases, whereas its Raman spectrum remains beam, confirming high crystallinity of the exfoliated ReS2. This
unchanged with increasing number of layers. Even after further arrangement is convenient as it provides a view of the crystalline
modulating the interlayer distance (coupling) by external structure from two orthogonal directions. Fast Fourier transform
hydrostatic pressure, the optical absorption and Raman spectrum images taken perpendicular to the planes displays quasi-
remain unchanged, implying that the interlayer interaction is hexagonal pattern formed by (100), (001) and (110) reflections
indeed rather weak. The presented results establish ReS2 as an with the diffraction spot width defined by the beam convergence
unusual, new member of the sTMDs family where the bulk acts as (B0.1 mrad) (Fig. 2b top), consistent with a well-defined
if it is electronically and vibrationally monolayers. crystalline structure (simulated kinematic diffraction patterns in
Fig. 2c top). To provide insight into the layer stacking order, we
Results have oriented the crystal with the electron beam pointing at the
DFT calculations and calculated crystal structure. Figure 1a [210] axis. The diffraction pattern consists of (120) and (002)
shows a comparison of the structure of a single layer of ReS2 with reflections (Fig. 2b,c bottom). The lack of odd ((001), (003), (005)
the conventional 1H structure of sTMDs such as MoS2. For ReS2, and so on.) reflections suggests lack of specific (Bernal) stacking,
MoS2
1.5 ReS2
Monolayer 0.1
0.0 Egap
MoS2
– 0.5 0.2
460 meV
–1.0 0.3
ReS2 –1.5
0.4 MoS2
–2.0
L
K′ ′ M′ 2 3 4 5 6 7 8 9
Inter-layer distance (Å)
Figure 1 | Different crystal structure and band structure of ReS2 from conventional sTMDs. (a) Side (top two panels) and top view (bottom two panels)
of ReS2 with distorted 1T crystal structure compared with the 1H structure of conventional sTMDs. The Re atoms dimerize as a result of the Peierls
distortion forming a Re chain denoted by the red zigzag line. (b) DFT calculated electronic band structure of bulk (orange solid curves) and monolayer
(purple dashed curves) ReS2. Both are predicted to be a direct bandgap semiconductor with nearly identical bandgap value at the G point. (c) The
calculated total energy of the system as a function of interlayer separation. The significantly shallower depth of the well in ReS2 implies much weaker
interlayer coupling energy in ReS2 as compared with MoS2.
ZA [001]
(010)
(100) (110)
ZA [210]
(120) (122)
(002)
21nm
(004)
(220)
Figure 2 | Structural characterization of ReS2. (a) HRTEM images taken from ReS2 along in- and out-of-plane directions. There are two flakes: one
flake’s basal plane is perpendicular to the electron beam (red square box) and the other flake’s basal plane is oriented parallel to the beam (green square
box). Scale bar, 20 nm. (b) Fast Fourier transform images from the denoted red/green windows. (c) Simulated kinematic diffraction patterns, and
(d) nano-beam electron diffraction pattern. Inset: Zoom-in image to the nano-beam electron diffraction pattern.
unlike in 2H-MoS2 that possesses Bernal stacking between the MoSe2, WSe2 and WS2) were also prepared by mechanical
layers. The (0,0,2l) reflections are inhomogeneously broadened exfoliation from bulk crystals.
both in the 001 and (120) directions. Nano-beam electron dif- All ReS2 flakes exhibit mPL at room temperature with a peak in
fraction pattern in Fig. 2d and inset allow direct comparison of the range of 1.5–1.6 eV. The peak energy increases slightly (by
the broadening because of the particular sample geometry. For o10%) for an isolated monolayer (Fig. 3a). The integrated PL
simplicity, we attribute all the inhomogeneous broadening of the intensity increases with the number of layers, and starts to
0,0,2l reflections to fluctuation in interlayer distance. From this saturate for flakes thicker than six layers, as shown in Fig. 3b. This
analysis, the interlayer spacing is determined to vary between 6.0 is in stark contrast to the behaviour observed in conventional
and 6.9 Å. This reduced degree of crystallinity is not turbostratic sTMDs such as MoS2 (ref. 4), MoSe2 (ref. 7), WS2 and WSe2
in nature; rather, it is a consequence of weak interlayer bonding (ref. 8), where the monolayer PL intensity is enhanced by orders
with small in-plane displacements from equilibrium position of magnitude as a result of the crossover from an indirect
(Supplementary Notes 1,2 and 7). bandgap in the bulk to a direct bandgap in monolayers. In those
Mo and W systems, the PL in the indirect-bandgap multilayers
Comparison between PL behaviour of ReS2 versus other arises from hot luminescence4 across the wider direct bandgap,
sTMDs. Considering such contrast in the crystal structure, we and is exponentially weaker in thicker layers. In addition, the PL
next focus experimentally on the physical properties of ReS2 peak position of these conventional sTMDs increases drastically
flakes. Mono- and multilayer ReS2 were mechanically exfoliated by 30–60% for thinner layers (Fig. 3c), which can be explained by
from synthetic bulk crystals (see Methods section for crystal stronger out-of-plane quantum confinement with decreasing
growth) onto Si substrates with 90 nm thermal oxide6. These thickness4. In contrast, the PL peak of ReS2 is nearly independent
ReS2 flakes were characterized using atomic force microscopy of the number of layers, indicating that thinning down the flake
(AFM), Raman spectroscopy, micro-PL (mPL), and nano-Auger does not enhance the quantum confinement of electrons in the
electron spectroscopy (AES). Nano-AES elemental composition system, and that neighbouring monolayers in the flake are already
analysis revealed an S/Re ratio of 1.98±0.10 (Supplementary largely electronically decoupled.
Fig. 8, Supplementary Note 6). Tapping-mode AFM scan typically
yielded 0.7 nm thickness for the monolayers. For comparison, Electronic decoupling in ReS2. To further test the interlayer
monolayers and multilayers of conventional sTMDs (MoS2, decoupling in ReS2, we modulate its interlayer distance by
PL intensity normalized
ReS2
5-Layer 100 least 11 modes in the 100–400 cm 1 range, which is significantly
PL signal (a.u.)
4-Layer more than for sTMDs with higher crystal symmetries. These
10–1 Raman peaks are mostly caused by the low crystal symmetry and
2-Layer
are associated with fundamental Raman modes (A1g, E2g, and E1g)
coupled to each other and to acoustic phonons. The two most
10–2 WSe2 prominent Raman peaks are at 163 and 213 cm 1, respectively,
Monolayer MoSe2
MoS2 WS2
and correspond to the in-plane (E2g) and mostly out-of-plane
10–3 (A1g-like) vibration modes (Supplementary Fig. 6, Supplementary
1.3 1.4 1.5 1.6 1.7 1.8 1.9 2 1 2 3 4 5 6 7
E (eV) Number of layers Note 3). Comparison between the Raman spectra of monolayer
and bulk ReS2 (Fig. 4c) does not show detectable changes within
the 0.3 cm 1 resolution. Again in contrast, the Raman spectrum
2.0 is known to be highly sensitive to the number of layers in MoS2
PR ReS2 (Fig. 4d) and other sTMDs12–14. For example, going from the bulk
Optical signal (a.u.)
1.6
amount. As ReS2 is vibrationally decoupled, its Raman spectrum is
ReS2
PL expected to be also less sensitive to hydrostatic pressure compared
1.4
MoS2 with other sTMDs. In Fig. 5a, we present the pressure induced
7.0G Pa
1.2
WSe2
200cm–1 change in the most prominent Raman peaks for bulk ReS2 and
MoSe2 Absorption
MoS2. As the interlayer coupling directly affects the out-of-plane
0.5G Pa
1.0 vibration modes, we focus on the pressure behaviour of the Raman
0 2 4 6 8 10 12 14 1.2 1.3 1.4 1.5 1.6 1.7 1.8
peaks associated with the out-of-plane vibration. The primarily
Number of layers E (eV)
out-of-plane, Ag-like Raman peak in ReS2 located at 213 cm 1
Figure 3 | Optical properties of ReS2 in comparison with conventional stiffens by only B1 cm 1 per GPa of hydrostatic pressure,
sTMDs. (a) PL spectrum of ReS2 flakes with different number of layers. whereas it is 42.7 cm 1 per GPa for the Ag mode in MoS2,
(b) Integrated PL intensity as a function of number of layers (normalized to 2.2 cm 1 per GPa in MoSe2 and 2.5 cm 1 per GPa in MoTe2
that of monolayer) in ReS2, MoS2, MoSe2, WS2 and WSe2. (c) Change in (ref. 15). In Fig. 5b, these numbers are compared against each
the PL peak position as a function of number of layers in ReS2, MoS2, MoSe2 other. Because the 213 cm 1 vibration mode in ReS2 is not purely
and WSe2. (d) Absorption coefficient of a bulk ReS2 flake (thickness Ag, but mixed with the more pressure-sensitive, in-plane Eg mode
B10 mm) at hydrostatic pressures ranging from 0.5 to 7.0 GPa. Also shown (see Supplementary Note 3), the intrinsic pressure coefficient of
is the PL and photo-modulated reflectance spectra of bulk ReS2 taken at the Ag mode is expected to be even lower in ReS2. Moreover,
ambient condition. It can be seen that a direct bandgap exists at considering the B50 times lower elastic modulus of ReS2 than
1.55±0.05 eV and is insensitive to the pressure. MoS2, the deformation coefficient (dAg/dz, where z is interlayer
distance) of out-of-plane vibration in ReS2 is estimated to be
orders of magnitude lower than MoS2. Related to the observed
applying hydrostatic pressure using a diamond anvil cell (DAC). electronic and vibrational decoupling in ReS2, the out-of-plane
The hydrostatic pressure is applied via a liquid medium onto a thermal conductivity is also calculated and found to be much
bulk flake (B10 mm), where the pressure is calibrated with the smaller than that of MoS2 (k>MoS2/k>ReS2E42), whereas the in-
standard Ruby luminescence9 (Supplementary Note 5). The plane thermal conductivity of ReS2 and MoS2 are on the same
transparent DAC window allows experiments for micro-Raman order of magnitude (Supplementary Note 4).
and optical micro-absorption. Comparing the absorption curves
with the PL and photo-modulated reflectance (see Supplementary
Fig. 7, Supplementary Note 5) spectra in Fig. 3d, the bandgap of Discussion
ReS2 is determined to be Egap ¼ 1.55 eV, close to the DFT On the basis of the distorted crystal structure, we discuss the
calculated Egap. In Fig. 3d, the optical absorption curve shows physical origin of the vanishing interlayer coupling in ReS2. We
nearly no change under the application of hydrostatic pressure up first note that each Re atom has seven valence electrons and
to 7 GPa. The out-of-plane linear elastic moduli of ReS2 and MoS2 therefore has one extra electron in the 1T or H structure. DFT
are estimated to be 0.4 and 20 GPa, respectively, calculated from calculations show that when this electron contributes to the
the second derivative of the interlayer coupling energy in Fig. 1c. formation of Re–Re bonds, the total energy of the system is
Therefore, a pressure of 7 GPa would significantly reduce their lowered, the Eigen values of the acoustic phonons become real
interlayer distance. In bulk MoS2 where adjacent layers are and the lattice becomes a distorted 1T structure as shown in
coupled, the pressure dependence of the bandgap is dEgap/ Fig. 1a (Supplementary Notes 1 and 2). Upon transition from 1T
dPB0.02 eV per GPa (refs 10,11), corresponding to a 0.14-eV (1H) to the distorted 1T structure, the energy gain is 1.1 (0.9) eV
shift of the absorption edge for the pressure range in Fig. 3d, per ReS2 molecule. In the distorted 1T structure, the Re atoms in
whereas ReS2 shows a total shift of 0.04 eV at the most. Therefore, the layer dimerize in such a way that they form a zigzag Re–Re
the overlap of electron wavefunctions from adjacent layers is chain extending along one of the lattice vectors16,17. Interestingly,
much weaker in ReS2 than in MoS2, such that a modulation of the by geometry consideration, the presence of the Re chains
interlayer distance cannot renormalize the band structure in the eliminates the energy benefit for the ReS2 layers to order in
former. Bernal (AB) or rhombohedral (ABC) stacking that occurs in other
sTMDs. Our DFT calculations show that while sliding one MoS2
monolayer over the other, one yields variation in total energy of
Vibrational decoupling in ReS2. The much weaker interlayer B200 meV; for ReS2 the energy difference is just on the order of
interaction energy shown in Fig. 1c also leads to a decoupling of 5 meV (Supplementary Note 2). The lack of ordered stacking
400
Eg Ag-like ReS2
Eg-like 2nd order
Frequency (cm–1)
100
L L
M′ K′ Raman (a.u.)
350 375 400 425
Raman shift (cm–1)
Figure 4 | Vibrational properties of ReS2 and Raman spectra. (a) Calculated phonon dispersion for monolayer ReS2. The phonon dispersion of bulk ReS2
(Supplementary Fig. 6) remains nearly identical to that of monolayer. (b) Raman spectrum measured on a monolayer ReS2. (c) Raman spectrum taken on
bulk (blue) and monolayer (red) ReS2. (d) Same on MoS2. The Raman spectrum of ReS2 shows no thickness dependence, whereas the MoS2 in-plane (E2g)
and out-of-plane (A1g) peaks are highly sensitive to the number of layers.
MoS2 : E2g
ReS2: Eg-like the sTMDs explored to date.
300
ReS2: Ag-like
Methods
Micro-Raman and lPL experiments. Raman and PL measurements were
performed using a 100 objective in a Renishaw micro-Raman/PL system. The
200 1 excitation laser wavelength is 488 nm, with laser power ranging from 100 to
1,000 mW power on a B6-mm2 spot.
ReS2: Eg
100
DAC measurements. Two symmetric DACs with type II diamonds of 300 and
MoS2 : E1g 500 mm flat culets were utilized for Raman and optical absorption experiments,
respectively. The pressure volume of 120 mm in height and 70 mm in diameter was
0 0 machined from a 250-mm-thick stainless steel gaskets. A methanol–ethanol mix-
0 5 10 15 20
M 2
2
e
e
2
2
oS
ture of 4:1 volume ratio was used as a pressure transmitting medium. Hydrostatic
eS
oT
oS
M
pressure applied inside the chamber hole was calibrated by measuring the R1 PL
line shift of ruby. A series of Raman spectra were taken on a ReS2 single-crystal
Figure 5 | Pressure dependence of Raman peaks of ReS2 compared with flake at various pressures up to 26 GPa using a 50 magnification, 18-mm
MoS2. (a) Variation of the most prominent Raman peaks of bulk ReS2 and working distance objective, on the Renishaw micro-Raman/PL system. For optical
MoS2 as a function of applied hydrostatic pressure. Empty blue points absorption measurements with the DAC, a home-built optics system equipped with
a 0.5-m single-grating monochromator and a tungsten–halogen lamp was used.
are taken from ref. 15, and filled red and blue points are from this work. The The sample flake was magnified and imaged on the focal plane, where an apertured
pressure dependence of MoS2 Ag and E2g peaks is confirmed by the silicon photodiode detects the light intensity with a lock-in amplifier.
good agreement between this work and ref. 15. The straight lines are linear
fit to the experimental data. (b) Comparison of pressure coefficient ReS2 growth. Single crystals of ReS2 were grown by the chemical vapour transport
(dAg /dP) of the out-of-plane (Ag) Raman peaks of MoS2, MoSe2, MoTe2 method, using Br2 as a transport agent, leading to n-type conductivity. Before the
and ReS2. It can be seen that the ReS2 out-of-plane vibration shows at least crystal growth, quartz tubes containing Br2 and the reacting elements (Re: 99.95%
a factor of 2 weaker dependence on pressure. pure, S: 99.999%) were evacuated and sealed. The quartz tube was placed in a three-
zone furnace and the charge pre-reacted for 24 h at 800 °C with the growth zone at
1,000 °C, thereby preventing transport of the product. The furnace was then
equilibrated to give a constant temperature across the reaction tube and was
significantly reduces the interlayer coupling energy, akin to programmed over 24 h to produce the temperature gradient at which single-crystal
turbostratic graphite where the interlayer interaction is growth takes place. The best results were obtained with a temperature gradient of
significantly weaker than that of Bernal-stacked graphite. 1,060 to 1,100 °C.
Another factor contributing to the vanishing interlayer coupling
comes from a much weaker polarization inside the ReS2 layers. High-resolution TEM. HRTEM and diffractometry was performed on JEOL2100F
Our charge density calculations show that, compared with MoS2, microscope equipped with EDS and EELS spectrometers. The exfoliated ReS2 was
there is a smaller charge difference between the neighbouring Re transferred to electron transparent supports (C-flat carbon support and patterned
SiN membranes from EMS). A number of 20–50-nm-thick ReS2 crystals were
and S planes, and hence a smaller dipole moment residing in each studied in various orientations including the in-plane and out-of-plane
layer. As a result of the weak intra-layer polarization, the van der configurations.
Waals force and the binding energy between adjacent ReS2 layers
are very weak (Supplementary Fig. 4). DFT calculations. Structural, electronic and vibrational properties have been
Therefore, ReS2 bulk crystals would be an ideal platform to calculated using GGA-PBE (Perdew-Burke-Ernzerhof) functional, as implemented
probe 2D excitonic and lattice physics, circumventing the in the VASP code18,19 within the DFT, which is demonstrated to yield rather
accurate results for sTMDs. Calculations were performed using the Generalized 18. Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys.
Gradient Approximation20 and projector augmented-wave potentials21. The kinetic Rev. B 47, 558–561 (1993).
energy cutoff for the plane-wave basis set was 500 eV. In the self-consistent 19. Kresse, G. & Furthmuller, J. Efficient iterative schemes for ab initio total-
potential and total energy calculations of ReS2 layers, a set of (31 31 1) k-point energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186
samplings was used for Brillouin zone integration. The convergence criterion of (1996).
self-consistent calculations for ionic relaxations was 10 5 eV between two 20. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation
consecutive steps. By using the conjugate gradient method, all atomic positions and made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
unit cells were optimized until the atomic forces were o0.01 eV/Å. Pressures on 21. Grimme, S. Semiempirical GGA-type density functional constructed with a
the lattice unit cell were decreased to values o0.5 kbar. Phonon dispersion curves long-range dispersion correction. J. Comput. Chem. 27, 1787–1799 (2006).
were obtained using the small displacement method22.
22. Alfe, D. PHON A program to calculate phonons using the small displacement
method. Comp. Phys. Commun. 180, 2622–2633 (2009).
References
1. Wang, Q. H., Kalantar-Zadeh, K., Kis, A., Coleman, J. N. & Strano, M. S.
Electronics and optoelectronics of two-dimensional transition metal
Acknowledgements
This work was supported by the United States Department of Energy Early Career Award
dichalcogenides. Nat. Nanotech. 7, 699–712 (2012).
DE-FG02-11ER46796. The high pressure part of this work was supported by COMPRES,
2. Radisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V. & Kis, A. Single-layer
the Consortium for Materials Properties Research in Earth Sciences, under National
MoS2 transistors. Nat. Nanotech. 6, 147–150 (2012).
Science Foundation Cooperative Agreement EAR 11-577758. The electron microscopy
3. Chhowalla, M. et al. The chemistry of two-dimensional layered transition metal
and nano-Auger measurements were supported by the user programme at the Molecular
dichalcogenide nanosheets. Nat. Chem. 5, 263–275 (2013).
Foundry, which was supported by the Office of Science, Office of Basic Energy Sciences,
4. Mak, K., Lee, C., Hone, J., Shan, J. & Heinz, T. F. Atomically thin MoS2: a new
of the United States Department of Energy under contract no. DE-AC02-05CH11231.
direct-gap semiconductor. Phys. Rev. Lett. 105, 136805 (2010).
S.A. gratefully acknowledges Dr Virginia Altoe of the Molecular Foundry for help with
5. Splendiani, A. et al. Emerging photoluminescence in monolayer MoS2. Nano
the TEM data acquisition and analysis. J.L. acknowledges support from the Natural
Lett. 10, 1271–1275 (2010).
Science Foundation of China for Distinguished Young Scholar (grant nos. 60925016 and
6. Novoselov, K. S. et al. Two-dimensional atomic crystals. Proc. Natl Acad. Sci.
91233120). Y.-S.H. and C.-H.H. acknowledge support from the National Science Council
USA 102, 10451–10453 (2005).
of Taiwan under project nos. NSC 100-2112-M-011-001-MY3 and NSC 101-2221-E-011-
7. Tongay, S. et al. Thermally driven crossover from indirect toward direct
052-MY3. H.S. was supported by the FWO Pegasus Marie Curie Long Fellowship pro-
bandgap in 2D semiconductors: MoSe2 versus MoS2. Nano Lett. 12, 5576–5580
gramme. The DFT work was supported by the Flemish Science Foundation (FWO-Vl)
(2012).
and the Methusalem programme of the Flemish government. Computational resources
8. Zhao, W. et al. Evolution of electronic structure in atomically thin sheets of
were partially provided by TUBITAK ULAKBIM, High Performance and Grid Com-
WS2 and WSe2. ACS Nano 7, 791–797 (2013).
puting Centre.
9. Mao, H. K., Xu, J. & Bell, P. M. Calibration of the ruby pressure gauge to
800 kbar under quasi-hydrostatic conditions. J. Geophys. Res.: Solid Earth 91,
4673–4676 (1986). Author contributions
10. Webb, A. W., Feldman, J. L., Skelton, E. F. & Towle, L. C. High pressure S.T. and J.W. conceived and designed the measurements. S.T. performed all optical
investigations of MoS2. J. Phys. Chem. Solids 37, 329–335 (1976). experiments. C.K. and D.F.O. contributed to nano-AES experiments. C.K., J.Y. and A.L.
11. Connell, G. A. N., Wilson, J. A. & Yoffe, A. D. Effects of pressure and performed DAC measurements. W.F., J.J., J.Z., Y.-S.H., C.-H.H. and K.L. prepared
temperature on exciton absorption and band structure of layer crystals: crystals and flakes for AFM, scanning electron microscopy, TEM and optical measure-
Molybdenum disulphide. J. Phys. Chem. Solids 30, 287–296 (1969). ments. S.A. performed HRTEM experiments and provided interpretation. H.S., S.L., J.L.
12. Lee, C. et al. Anomalous lattice vibrations of single- and few-layer MoS2. ACS and F.M.P. carried out theoretical analysis and DFT calculations. All authors read and
Nano 4, 2695–2700 (2010). edited the manuscript.
13. Sahin, H. et al. Anomalous Raman spectrum and dimensionality effects in
WSe2. Phys. Rev. B 87, 165409 (2013).
14. Berkdemir, A. et al. Identification of individual and few layers of WS2 using Additional information
Raman Spectroscopy. Sci. Rep. 3, 1755 (2013). Supplementary Information accompanies this paper at http://www.nature.com/nature
15. Sugai, S. & Ueda, T. High-pressure Raman spectroscopy in the layered communications.
materials 2H-MoS2, 2H-MoSe2, and 2H-MoTe2. Phys. Rev. B 26, 6554–6558
(1982). Competing financial interests: The authors declare no competing financial interests.
16. Ho, C. H., Huang, Y. S., Tiong, K. K. & Liao, P. C. Absorption-edge anisotropy Reprints and permission information is available online at http://npg.nature.com/
in ReS2 and ReSe2 layered semiconductors. Phys. Rev. B 58, 16130–16135 reprintsandpermissions/
(1998).
17. Friemelt, K. et al. Scanning tunneling microscopy with atomic resolution on How to cite this article: Tongay, S. et al. Monolayer behaviour in bulk ReS2 due to
ReS2 single crystals grown by vapor phase transport. Ann. Phys. 1, 248–253 electronic and vibrational decoupling. Nat. Commun. 5:3252 doi: 10.1038/ncomms4252
(1992). (2014).