0% found this document useful (0 votes)
28 views

Temperature dependent photoluminescence from WS2 nanostructures

This study investigates the temperature-dependent photoluminescence and Raman spectra of WS2 nanostructures, highlighting their potential applications in optoelectronics and spintronics due to their tunable bandgap and strong spin-orbit coupling. The research reveals a red-shift in luminescence with increasing temperature and identifies various optical transitions through systematic spectroscopy. The findings suggest that WS2 quantum dots exhibit enhanced properties compared to their bulk counterparts, making them attractive for future electronic devices.

Uploaded by

phitruong.le7901
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
28 views

Temperature dependent photoluminescence from WS2 nanostructures

This study investigates the temperature-dependent photoluminescence and Raman spectra of WS2 nanostructures, highlighting their potential applications in optoelectronics and spintronics due to their tunable bandgap and strong spin-orbit coupling. The research reveals a red-shift in luminescence with increasing temperature and identifies various optical transitions through systematic spectroscopy. The findings suggest that WS2 quantum dots exhibit enhanced properties compared to their bulk counterparts, making them attractive for future electronic devices.

Uploaded by

phitruong.le7901
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 7

Journal of Materials Science: Materials in Electronics

https://doi.org/10.1007/s10854-018-0137-3

Temperature dependent photoluminescence from ­WS2 nanostructures


Shivani Sharma1 · Shubham Bhagat1 · Jasvir Singh1 · Manzoor Ahmad2 · Sandeep Sharma1

Received: 11 April 2018 / Accepted: 30 September 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Owing to their intriguing physical properties, two-dimensional transition metal dichalcogenides such as ­WS2 and ­MoS2 have
gained significant attention in the research community. Their tunable bandgap justify their use in future optical and nano-
electronics devices. Here, we report the room temperature Raman spectra and temperature dependent photoluminescence
of ­WS2 nanostructures prepared in liquid media. The resonance Raman spectra revealed various first order modes together
with higher order modes that were inaccessible with excitation away from resonance absorption. The luminescence from
these nanostructures displayed red-shift and linear temperature dependence in the range 293–363 K. The observed negative
temperature coefficients are very small and may arise from anharmonicity and thermal expansion. Further, optical measure-
ments revealed that ­WS2 quantum dots exhibits strong spin–orbit coupling ≈ 650 meV, larger than observed for monolayer
sheets of W
­ S2 (≈ 400 meV). The stronger spin–orbit coupling together with highly luminescent nature make them attractive
for applications in spintronics and optoelectronics devices.

1 Introduction different physical properties in monolayer or few layered


form that are not seen in their bulk counterpart. This is
In recent few years two-dimensional materials like graphene attributed to quantum confinement and surface effects [6,
and its inorganic analogue, i.e., transition metal di-chalco- 7]. For instance, in bulk form TMDCs have an indirect band
genides (TMDCs) of type ­MX2 (M = Mo or W and X = S, gap, but when reduced to monolayers the band gap becomes
Se or Te) have attracted world-wide attention due to their direct and give rise to enhanced photoluminescence [8, 9].
intriguing physical properties and potential application in Thus, tuning band gap in these 2-D materials suggest
nanoelectronics and photovoltaic devices [1].Whereas gra- their promising use in light sensitive applications. Although
phene consist of single carbon atomic-thick layer, TMDCs band gap tuning can be achieved by various means such as
(e.g. ­MoS2, ­MoSe2, ­WS2, ­WSe2, etc.), on the other hand induced strain in nanostructures [10] or external application
have a “sandwich” type structure where a transition metal of stress [11–13] doping and alloying [14]. There are other
layer (e.g., Mo or W) is sandwiched between two chalcogen external factors which may influence the physical proper-
layers (e.g., S, Se or Te). As a result TMDCs are character- ties of the semiconducting materials. For instance, it is well
ized by weak and noncovalent interlayer bonding similar to known that optical properties in semiconductors strongly
graphene and other van der Waals materials. This allows the depend upon the temperature [15–17]. For instance, Pandey
exfoliation of bulk TMDCs into single or few layer sheets by et al. has recently reported the existence of various optical
various physical or chemical means including adhesive tape transitions originating from excitons, biexciton and trions.
exfoliation [2, 3] chemical exfoliation through lithium inter- Due to broad spectral features of exciton at normal tempera-
calation and/or solvent assisted exfoliation [4, 5]. Another ture the experimental realization of the biexciton remains
interesting feature of TMDCs is that they exhibit entirely a challenging task. A systematic temperature dependent
photoluminescence spectroscopy enabled them to identify
various optical transitions [18]. The understanding of such
* Sandeep Sharma optical transitions is crucial for future optoelectronic appli-
[email protected]
cations. Such types of systematic investigations are rare in
1
Department of Physics, Guru Nanak Dev University, case of W­ S2. In view of these aspects, it is crucial to under-
Amritsar 143005, India stand the response of these materials to temperature. Both
2
Department of Chemistry, Guru Nanak Dev University, micro-Raman as well as photoluminescence spectroscopy
Amritsar 143005, India

13
Vol.:(0123456789)
Journal of Materials Science: Materials in Electronics

offer a non-destructive method to characterize the structural used for Raman spectroscopy measurements. The optical
and lattice vibrations of a crystal. Here, in this article we characterization was performed on liquid based suspension.
present room temperature Raman spectra and temperature For transmission electron microscopy, a few drops of sus-
dependent photoluminescence from W ­ S2 nanostructures pension were dropped on carbon coated copper grid. The
prepared in liquid media. The emission intensity from these grid was further dried under an electric bulb.
nanostructures decreases with rise in temperature and a High resolution Transmission electron microscopic imag-
weak red-shift with increasing temperature is also observed. ing was carried out using JEOL JEM-2100. The vibrational
spectra of the ­WS2 nanostructure were investigated using
Renishaw Invia Reflex micro Raman spectrometer using vis-
2 Experimental ible excitations (488 nm and 514 nm). For investigating the
optical properties of the samples, absorption spectra were
Crystalline ­WS2 powder and polyvenyle pyrrolidone (PVP) recorded by using Shimadzu UV-3600 spectrophotometer.
were obtained from Sigma-Aldrich and from Loba Chemie Photoluminescence study was carried out by using Fluorolog
India, respectively. A mixture of 2 g ­WS2 in 50 ml of de- Horiba spectrometer.
ionized water together with suitable amount of PVP was
processed using porbe sonicator (PCI-Analytics, India) for
1½ h. This process resulted in the exfoliation of the crystal- 3 Results and discussion
line powder and thin sheets of W­ S2 appeared on the liquid
surface. These thin sheets were removed and remaining sus- Morphology and crystal structure of the nanostructures were
pension was centrifuged at 12,000 rpm for 30 min. From the investigated using transmission electron microscopy. Fig-
suspension, a few drops were transferred on glass substrates ure 1a, b show the HR-TEM images of few layer ­WS2 sheets.
and the sample was dried in oven at 80 °C. This sample was Sheets of different dimensions and thickness are clearly

Fig. 1  HR-TEM images for


­WS2 nanostructures a, b shows
few layer sheets of ­WS2. In b
various overlapping regions
between ­WS2 sheets are clearly
visible. In c quantum dots of
variable size and thickness are
clearly visible. d Is the magni-
fied view of quantum dots

13
Journal of Materials Science: Materials in Electronics

visible in the images. The image contrast in Fig. 1b implies interaction [20]. It should be noted that due to ultrathin
the presence of overlapped thin sheets of W ­ S2. The insets in nature of the quantum dots, the relative intensity of the
Fig. 1a, b display the digitally filtered image showing lattice obtained Raman signal or peak intensity in XRD data reduce
fringes with 0.3 nm spacing corresponding to (004) planes. significantly [19, 20]. Broad peak at 353 cm−1 imply that
In addition to the clear planes some distortions are also vis- 2LA (M) mode overlaps with E12g mode. In addition to these
ible inside the inset in Fig. 1a. This indicates that probe modes the spectra also contains a zone-edge mode at
sonication affects the crystalline quality of the starting mate- 175.5 cm−1 identified as a longitudinal acoustic mode at M
rial. Figure 1c shows the W
­ S2 quantum dots of variable sizes point of the Brillouin zone [LA (M)] [21–23]. The multi-
(ranging from 4 to 10 nm). Some of the quantum dots appear phonon combination of these modes gives rise to additional
to be darker as compared to other, implying the variation in weak peaks in the spectra. When the sample is excited at
the thickness also. Most of the quantum dots have size in the 514 nm, the Raman spectra reveals many second-order peaks
range of ≈ 10 nm. The insets in Fig. 1d show the HR-TEM relatively stronger than those observed in spectra obtained
images of the quantum dots. The fringes with interplanar at 488 nm. The energy for excitation at 514 nm (2.41 eV) is
spacing of 0.18 nm and 0.21 nm confirm the presence of close to the energy of the B exciton in few layer W ­ S2
(105) and (006) planes, respectively (JCPDS 08-0237). (≈ 523 nm, 2.37 eV). Therefore, excitation at 514 nm usually
Figure 2a, b display typical Raman spectra of ­WS2 nano- gives a resonant Raman spectrum of W ­ S2 [22]. Thus, we see
structures acquired at room temperature using 488 nm and that under resonant excitation higher order Raman modes
514 nm excitations, respectively. With 488 nm excitation, become intense. The asymmetric resonant mode at
the spectrum is dominated by two basic first order vibra- ≈ 700 cm−1 is assigned as 4LA(M). The mode at ≈ 580 cm−1
tional modes at the Γ point; ­A1g (Γ) (418.1 cm−1) and E12g (Γ) is leveled as ­A1g (M) + LA(M), which when combined with
(354.14 cm−1). The mode A ­ ig arises due to out-of-plane mode ­A1g (M) LA (M) at ≈ 233 cm−1 can be utilized to
motion of the sulfur atoms whereas E12g originates due to roughly estimate the energy of A­ 1g (M) and LA (M) modes.
relative in-plane motion of the sulfur and tungsten atoms. This gives ­A1g (M) and LA (M) modes at 406 cm−1 and
When compared with bulk (­ A1g (Γ) (419.11 cm−1) and E12g 176 cm−1, respectively, close to experimentally observed
(Γ) (353.79 cm−1) data not shown here) [19], the E12g mode value for LA (M) mode. Thus energy of ­A1g (M) mode
is stiffened whereas A­ 1g mode is softened, possibly due to (406 cm −1 ) is relatively farther from A ­ 1g (Ʃ) mode
ultrathin structure of ­WS2 leading to absence of layer–layer (419 cm−1) by 13 cm−1. It is to be noted that previous reports

Fig. 2  Room temperature


Raman spectra of W­ S2 nano- (b) A1g
λexc = 514 nm
structures obtained with an
excitation wavelength a 488 nm,
2LA(M)

1
and b 514 nm E2g

807 A1g+ E2g+E2g


2
233 A1g(M) - LA(M)
Intensity (arb. units)

295 2LA-2E2g

1
322 2LA -E2g
2

A1g(M)+LA(M)
517 3LA(M)
175.5 LA(M)

4LA(M)

(a) A1g
λexc = 488 nm
1
E2g
2LA(M)
2LA-2E22g

2g

4LA(M)
2

3LA(M)
2LA-E
LA(M)

100 200 300 400 500 600 700 800 900


-1
Raman Shift (cm )

13
Journal of Materials Science: Materials in Electronics

have shown that these two modes, i.e., ­A1g (M) and ­A1g (Σ) (2.1 eV). They have also shown that room temperature pho-
have dispersionless character in monolayer TMDCs [23, 24]. toluminescence from such nanostructures is enhanced and
Figure 3 shows the typical linear absorption spectrum lies in the blue-green spectral region [20]. Thus, it is very
of ­WS2 nanostructures acquired at room temperature. The unlikely that in present samples the absorptions at 395 nm
absorption data is deconvoluted using gaussian function and 327 nm are linked with the absorption spectrum of few
and we see that spectra consists of various spectral peaks at layer sheets of W ­ S2. The band edge absorption in the UV
627 nm, 523 nm, 458 nm, 395 nm and 327 nm together with region is quiet far away from direct or indirect band gap in
a band edge absorption in the ultra voilet region. The first ­WS2, it is indeed the absorption feature arising from the
two peaks at 627 nm and 523 nm and are characteristics exci- ­WS2 quantum dots. These two peaks are likely to be the
ton peaks ‘A’ and ‘B’, respectively. These two peaks belong excitonic absorptions A and B from the quantum dots and
to the lowest energy exciton states and are due to transitions are at relatively shorter wavelength than is the case for ­WS2
from the highest energy spin–orbit coupling induced split-off few layered sheets (627 and 523 nm). The energy separa-
valence bands to the lowest energy conduction band states tion between two peaks (327 nm and 395 nm) is 650 meV
at the K point in the Brillouin zone [20, 25–27]. The energy (near IR region, ≈ 1907 nm), much larger than observed
separation between peaks A and B is close to 0.4 eV. This is for few layer sheets (400 meV). These results, suggest that
in good agreement with the theoretically calculated and pre- by rightly controlling the lateral size of the quantum dots,
viously experimentally reported value for ­WS2 [9, 26–28]. spin–orbit splitting in these nanostructures can be tuned. The
It should be noted that excitonic absorptions A and B are next immediate question is that can we observe such a large
not due to quantum dots rather they corresponds to the few enhancement in band gap when material is transformed from
layered sheets of ­WS2. Another absorption feature marked bulk into quantum dots. The Bohr radius determines the rel-
as C and the one at 458 nm is attributed to the transitions evant particle size at which quantum confinement effects
between density of states peaks between valence band and governing such enhancement in band gap are expected. For
conduction bands [27]. The absorption spectra also reveal ­WS2 the Bohr radius is around 3.7 nm. This implies that
absorptions close to 395 nm (3.14 eV) and 327 nm (3.79 eV) nanostructures with size in the range of ≈ 8 nm may give
together with band edge absorption in the ultraviolet region. rise to larger band gap. Under effective mass approximation,
The indirect band gap in bulk ­WS2 is ≈ 1.2 eV [28] whereas the band gap variation with size is expressed through the
the direct band gap in mono layer of ­WS2 is around 2.1 eV. following relationship [30].
Recently, indirect transitions with energy larger than these
h2 1.8e2
values have been predicted in few-layer sheets of ­MoS2 and E∗ = Eg + 2
− (1)
8𝜇r 4𝜋𝜀0 𝜀r
­WS2 [29]. But the energy separation between correspond-
ing valence band maxima and conduction band minima is where, ­Eg is band gap of bulk material, r is radius of the
much lower than the 395 nm and 327 nm absorptions we quantum dot or nanoparticle, ɛ0 is permittivity of free space
have noticed. In a recent study, Lin et al. have shown that and ɛ r being relative permittivity of the material and
­WS2 quantum dots with size in the range 8–15 nm possess 𝜇 = m e+mh being reduced mass of the exciton. In case of
mm

a very large direct transition energies 3.16 eV (393 nm) and e h

3.73 eV (333 nm) compared to the ­WS2 few layer sheets ­ S2 the effective mass of electron, ­me = 0.33 m0 and that of
W
hole ­mh = 0.43 m0 [31]. The quantum dots size in present
sample ranges from 4 to 10 nm. Equation (1) predicts a band
gap of 3.95–1.71 eV for particle with size ranging from 4 to
10 nm. Thus, such a large enhancement in band gap is
expected and attributed to quantum confinement effects
within ­WS2 nanostructures. Further, it should be noted that
Absorption (arb. units)

possibility of nanostructures with size smaller than 4 nm


327 cannot be excluded. Unlike other semiconductors, the
395
C
absorption spectra for W­ S2 does not display gap between
458 various excitonic features and band edge absorption, thus
B elevating the problem of accurate determination of various
523 A peak positions. Further, difference in exact peak position
627
might arise due to life time broadening and overlap with
higher energy excitonic absorptions, excited states of exci-
300 400 500 600 700 800
Wavelength (nm) tons and/or electron–phonon coupling that results in con-
tinuous absorption [32, 33].
Fig. 3  UV–vis spectra of ­WS2 nanostructures

13
Journal of Materials Science: Materials in Electronics

Figure 4 shows the photoluminescence emission spectra λexc = 310 nm


acquired in the temperature range from 20 to 70 °C with Temperature = 200 C
λexc = 310 nm. As we see, the emission intensity reduces
and the spectra also display a slight red-shift with rise in

Intensity (arb. units)


temperature. In Fig. 5 the emission spectra obtained at 20 °C
is shown. The deconvolution of the spectra reveals three
emission peaks at ≈ 340 nm, 390 nm and 432 nm. These
three emission peaks corresponds to the absorption features
noticed in the UV–visible measurements. The energy sepa-
ration between former two peaks is ≈ 480 meV, larger than
the value found in bulk W ­ S2 (≈ 400 meV). Lin et al. has
attributed these features to the spin–orbit coupling induced
325 350 375 400 425 450 475 500 525 550 575
splitting of valence band in confined W ­ S2 quantum dots Wavelength (nm)
[20]. Zhu et al. has recently demonstrated that due to bro-
ken inversion symmetry a giant spin splitting of the order
Fig. 5  Deconvoluted photoluminescence spectra of ­WS2 nanostruc-
of 456 meV is expected in monolayered W ­ S2 [34]. Due to tures. The spectra were obtained at λexc = 310 nm
inversion asymmetry in monolayer W ­ S2, the d-orbital cor-
responding to W atoms have strong spin-orbit coupling. As
a result, largest splitting at K point of the Brillouin zone is increased electron–phonon interaction and/or lattice expan-
observed. The expected enhancement in the band gap due to sion at higher temperature [16, 17].
quantum confinement effect also suggests that corresponding To investigate other spectral features, the PL emission
excitonic features may appear at relatively higher energies with λexc = 400 nm and in the temperature range 20–90 °C
in such nanostructures. Therefore, larger energy separation was also taken (Fig. 6). A closed analysis reveals four peaks
between two peaks at lower wavelength is justified in quan- at 445 nm, 468 nm, 495 nm and 534 nm. The emission
tum confined ­WS2 quantum dots. Further, Lin et al. also feature at 445 nm might be similar to the one at 432 nm
reported a large PL emission which is blue-shifted as com- (obtained with λexc = 310 nm). Thus, we see that emission
pared with the emission from monolayered sheets. These, peaks are slightly shifted w.r.t the absorption spectra in
results suggest the role of quantum confinement effects in Fig. 3. Further, with two different excitations, the PL spectra
determining the optical properties of such 2-D nanostruc- is different. This is a signature of slight polydisperse nature
tures. Further, little red-shift in the PL spectra is similar of the sample [9]. At lower excitation wavelength emission
to that observed in semiconductors. In semiconductors, the spectrum over wider energy range is obtained whereas at
optical band gap reduces with rise in temperature, causing higher excitation wavelength the emission covers an energy
a red shift in absorption/emission spectra. This is widely range at higher wavelength. It is surprising that we did not
accepted phenomena in semiconductors and is attributed to observe the PL emission corresponding to absorption ‘A’.
Figure 7a shows the variation of intensity with tempera-
ture for the emission peak at 390 nm. This peak corresponds
λexc = 310 nm 70
60
50
40
λexc = 400 nm 90
Intensity (arb. units)

30 80
70
20
60
Intensity (arb. units)

Increasing 50
Temperature 40
Increasing 30
Temperature 20

325 350 375 400 425 450 475 500 525 550 575
Wavelength (nm) 400 450 500 550 600 650 700
Wavelength (nm)
Fig. 4  Photoluminescence spectra of ­WS2 nanostructures at different
temperatures. Spectra were acquired with an excitation wavelength of Fig. 6  Photoluminescence spectra of W
­ S2 nanostructures. The spectra
310 nm were acquired at different temperatures with λexc = 400 nm

13
Journal of Materials Science: Materials in Electronics

Fig. 7  Intensity variation with 1.1


temperature. Data is given 1.0
λexc = 310 nm (a) λexc = 400 nm
(b)
for PL spectra acquired at a TC1 = - 6.48x10-3
1.0 TC1 = - 4.3x10-3
λexc = 310 nm for PL emis-
sion corresponding to peak at 0.9
390 nm and b λexc = 400 nm for 0.9
emission peak at 468 nm

IT/IT0

IT/IT0
0.8
0.8

0.7 0.7

0.6 0.6
0 10 20 30 40 50 0 10 20 30 40 50 60 70
T-T0 (K) T-T0 (K)

to λexc = 310 nm. Similar data is shown in Fig. 7b for emis- α is the first order temperature coefficient. The fit to the
sion peak at 468 nm obtained with λexc = 400 nm. Here ­IT0 emission data is shown in Fig. 7a, b. Negative α values,
and and ­IT are the emission intensities at 20 °C and any 6.48 × 10−3 K−1 and 4.3 × 10−3 K−1 have been obtained for
other temperature, respectively. In the former case the lin- the emission intensity corresponding to peaks at 390 nm
ear fit to the data gives the first order temperature coeffi- (λexc = 310 nm) and 468 nm (λexc = 400 nm).
cient ≈ − 6.48 × 10−3 K−1 whereas in latter case its value is
− 4.3 × 10−3 K−1, relatively smaller than the value observed
in previous case. The temperature dependent emission inten-
sity is usually expressed as [8, 14, 35]. 4 Conclusions
I0× krad (T)
I(T) = (2) We have measured systematically the room temperature
krad (T) + knonrad (T)
Raman spectra and the temperature-dependent photolumi-
where, ­I0 is the maximum emission intensity at low tempera- nescence spectra of few layer sheets and quantum dots of
ture, ­krad(T) and ­knonrad(T) are the temperature dependent ­WS2. The resonance Raman spectra with 514 nm excita-
radiative and non-radiative recombination rates, respectively. tion revealed various higher order modes which were inac-
The latter constitutes, the recombination rates due to trap- cessible with other excitation. The nanostructures display
ping at defects sites and electron relaxation within the con- various excitonic features corresponding to a few layer
duction band and valence bands. The intra-band relaxation ­WS2 and a band edge absorption in the ultraviolet region.
to the band minima is usually accompanied with faster rates We observed that the photoluminescence intensities of
and may give rise to emission at lower energy compared to the ­WS2 nanostructures decrease with rise in temperature.
absorption peaks. Further, the increased electron–phonon This is attributed to the increased electron–phonon interac-
interactions at higher temperature give rise to the thermally tions at higher temperature. A slight red-shift in the pho-
activated non-radiative recombination. As a result the emis- toluminescence emission with rise in temperature is also
sion intensity and hence the quantum efficiency decrease noticed. Such red-shift might arise from the anharmonic
dramatically with increase in the temperature. In general effects. Further, the optical studies revealed that ­WS 2
the increased temperature causes a reduction in the band quantum dots exhibits absorptions at much smaller wave-
gap of a semiconductor, which is attributed to the enhanced lengths with larger spin–orbit coupling ≈ 650 meV, larger
electron–phonon interactions as well as lattice expansion at than observed for monolayer sheets of ­WS2 (≈ 400 meV).
higher temperature [16, 17]. A qualitative idea about red- This is in agreement with the quantum confinement effects
shift can also be gained from the shift of other absorption in these nanostructures and previously observed values for
feature. For instance, the absorption features at ≈ 327 nm ­WS2 quantum dots. These results indicates indicate that by
appear at ≈ 340 nm in the emission spectra. Further, a red- rightly controlling the dimensions of the nanostructures
shift, although very small, is also visible from the emission the strength of spin–orbit splitting can be tuned and hence
spectra at different temperatures. The temperature depend- a control over optical properties can be achieved.
ence of emission intensity is fitted with the equation:
Acknowledgements One of the authors Shivani Sharma acknowledges
IT = IT0 [1 + 𝛼(T − T0 )] (3) the UPE-fellowship provided by GNDU Amritsar. This work was sup-
ported by UGC-New Delhi, India under the Grant no. F.30-137/2015
where, ­IT and ­IT0 are the emission intensities at tempera- BSR.
ture T and reference temperature (20 °C), respectively and

13
Journal of Materials Science: Materials in Electronics

References 20. L. Liangxu, X. Yaoxian, Z. Shaowei, R.M. Ian, C.M. Ong Albert,
D.A. Allwood, Fabrication of luminescent monolayered tungsten
dichalcogenides quantum dots with giant spin-valley coupling.
1. Q.H. Wang, K.K. Zadeh, A. Kis, J.N. Coleman, M.S. Strano, Elec-
ACS Nano 7, 8214–8223 (2013)
tronics and optoelectronics of two-dimensional transition-metal
21. A. Berkdemir, H.R. Gutierrez, A.R. Botello-Mendez et al., Iden-
dichalcogenides. Nat. Nanotechnol. 7, 699–712 (2012)
tification of individual and few layers of ­WS2 using Raman spec-
2. K.S. Novoselov, Graphene: materials in the flatland (Nobel lec-
troscopy. Sci. Rep. 3, 1755- (2013)
ture). Angew. Chem. Int. Ed. 83, 837–849 (2011)
22. X. Zhang, Q. Xiao-Fen, S. Wei, W. Jiang-Bin, J. De-Sheng, T.P.
3. K.S. Novoselov, D. Jiang, F. Schedin, T.J. Booth, V.V. Khotkevich,
Heng, Phonon and Raman scattering of two-dimensional transi-
S.V. Morozov, A.K. Geim (2005) Two-dimensional atomic crys-
tion metal dichalcogenides from monolayer, multilayer to bulk
tals. Proc Natl Acad Sci USA 102:3010451–3010453
material. Chem. Soc. Rev. 44, 2757–2785 (2015)
4. B.K. Miremadi, S.R. Morrison, The intercalation and exfoliation
23. N. Wakabayashi, H.G. Smith, R.M. Nicklow, Lattice dynamics of
of tungsten disulfide. J. Appl. Phys. 63, 4970 (1988)
hexagonal ­MoS2 studied by neutron scattering. Phys. Rev. B 12,
5. A. Ghorai, A. Midya, R. Maiti, S.K. Ray, Exfoliation of ­WS2 in
659–663 (1975)
the semiconducting phase using a group of lithium halides: a new
24. A. Molina, L. Wirtz, Phonons in single-layer and few-layer ­MoS2
method of Li intercalation. Dalton Trans. 45, 14979–14987 (2016)
and ­WS2. Phys. Rev. B 84, 155413 (2011)
6. J.P. Wilcoxon, G.A. Samara, Strong quantum-size effects in a
25. H.M. Hill, A.F. Rigosi, C. Roquelet, A. Chernikov, T.C. Berkel-
layered semiconductor: ­MoS2 nanoclusters. Phys. Rev. B 51,
bach, D.R. Reichman, M.S. Hybertsen, L.E. Brus, T.F. Heinz, F.
7299–7302 (1995)
Tony, Observation of excitonic rydberg states in monolayer M ­ oS2
7. D.H. Feng, Z.Z. Xu, T.Q. Jia, X.X. Li, S.Q. Gong, Quantum size
and ­WS2 by photoluminescence excitation spectroscopy. Nano
effects on exciton states in indirect-gap quantum dots. Phys. Rev.
Lett. 15, 992–2997 (2015)
B 68, 035334 (2003)
26. A. Ramasubramaniam, Large excitonic effects in monolayers
8. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.Y. Jonghwan
of molybdenum and tungsten dichalcogenides. Phys. Rev. B 86,
Chim, G. Galli, F. Wang, Emerging photoluminescence in mon-
115409 (2012)
olayer ­MoS2. Nano Lett. 10, 1271–1275 (2010)
27. W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.H. Tan, G.
9. S. Sharma, S. Bhagat, J. Singh, R.C. Singh, S. Sharma, Excitation-
Eda, Evolution of electronic structure in atomically thin sheets of
dependent photoluminescence from W ­ S2 nanostructures synthe-
­WS2 and ­WSe2. ACS Nano 7, 791–797 (2013)
sized via top-down approach. J. Mater. Sci. 52, 11326–11336
28. R.A. Bromley, R.B. Murray, A.D. Yoffe, The band structures of
(2017)
some transition metal dichalcogenides. III. Group VIA: trigonal
10. S. Bhagat, S. Sharma, J. Singh, S. Sharma, Strain-induced tuning
prism materials. J. Phys. C Solid State Phys. 5, 759 (1972)
of optical properties of layered ­MoS2 in communication
29. W. Zhao, R.M. Ribeiro, M. Toh, A. Carvalho, C. Kloc, A.H.C.
11. S. Bhattacharyya, T. Pandey, A.K. Singh, Effect of strain on elec-
Neto, G. Eda, Origin of indirect optical transitions in few-layer
tronic and thermoelectric properties of few layers to bulk M ­ oS2.
­MoS2,WS2, and ­WSe2. Nano Lett. 13, 5627–5634 (2013)
Nanotechnology 25, 465701 (2014)
30. Z.X. Gan, L.Z. Liu, H.Y. Wu, Y.L. Hao, Y. Shan, X.L. Wu, P.K.
12. H. Keliang, P. Charles, M.K. Fai, S. Jie, Experimental demonstra-
Chu, Quantum confinement effects across two-dimensional planes
tion of continuous electronic structure tuning via strain in atomi-
in ­MoS2 quantum dots. Appl. Phys. Lett. 106, 233113 (2015)
cally thin ­MoS2. Nano Lett. 13, 2931–2936 (2013)
31. J. Chang, L.F. Register, S.K. Banerjee, Ballistic performance com-
13. H. Yeung Yu, L. Xiaofei, J. Wenjing, C.N. Chan Yui, H. Jianhua,
parison of monolayer transition metal dichalcogenide ­MX2(M =
H. Yu-Te, L. Lain-Jong, G. Wanlin, L.S. Ping, Exceptional tun-
Mo, W; X = S, Se, Te) metal-oxide-semiconductor field effect
ability of band energy in a compressively strained trilayer MoS2
transistors. J. Appl.Phys. 115, 084506 (2014)
sheet. ACS Nano 7, 7126–7131 (2013)
32. D.Y. Qiu, F.H. da Jornada, S.G. Louie, Optical spectrum of ­MoS2:
14. Y. Chen, W. Wen, Y. Zhu, N. Mao, Q. Feng, M. Zhang, H.P.
Many-body effects and diversity of exciton states. Phys. Rev. Lett.
Hsu, J. Zhang, Y.S. Huang, L. Xie, Temperature-dependent pho-
111, 216805 (2013)
toluminescence emission and Raman scattering from M ­ o1–xWxS2
33. K. He, N. Kumar, L. Zhao, Z. Wang, K.F. Mak, H. Zhao, J. Shan,
monolayers. Nanotechnology 27, 445705 (2016)
Tightly bound excitons in monolayer W ­ Se2. Phys. Rev. Lett. 113,
15. S.M. Sze, N.K. Kwok. (2006) Physics and Properties of Semicon-
026803 (2014)
ductors A Review. Wiley, New York, pp 5–75
34. Z.Y. Zhu, Y.C. Cheng, U. Schwingenschlogl, Giant spin-orbit-
16. Y.P. Varshni, Temperature dependence of the energy gap in semi-
induced spin splitting in two-dimensional transitional-metal
conductors. Physica 34, 12170–12177 (1967)
dichalcogenide semiconductors. Phys. Rev. B 84, 153402 (2011)
17. K.P.O. Donnell, X. Chen, Temperature dependence of semicon-
35. N. Haiyan, W. Zilu, W. Wenhui, L. Zheng, L. Yan, C. Qian, H.
ductor band gaps. Appl. Phys. Lett. 58, 2924–2926 (1991)
Daowei, T. Pingheng, M. Feng, W. Xinran, W. Jinlan, N. Zhenhua,
18. J. Pandey, A. Soni, Unraveling biexciton and excitonic excited
Strong photoluminescence enhancement of ­MoS2 through defect
states from defect bound states in monolayer ­MoS2. Appl. Surf.
engineering and oxygen bonding. ACS Nano 8, 5738–5745 (2014)
Sci. 463, 52–57 (2019)
19. S. Sharma, J. Singh, S. Bhagat, M. Singh, S. Sharma, Size tunable
photoluminescence from WS2 nanostructures. Mater Res Express
5, 045047 (2018)

13

You might also like