You are on page 1of 11

Applied Surface Science 531 (2020) 147303

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Anomalous vibrational behavior of two dimensional tellurium: Layer T


thickness and temperature dependent Raman spectroscopic study

Ram Ashish Yadava, N. Padmab,c, , Shashwati Senb,c, K.R.S. Chandrakumarc,d, H. Donthulae,

Rekha Raoc,f,
a
Government Girls’ College Vidisha, 464001 Madhya Pradesh, India
b
Technical Physics Division, Bhabha Atomic Research Centre, Mumbai 400085, India
c
Homi Bhabha National Institute, Anushaktinagar, Mumbai 400094, India
d
Theoretical Chemistry Section, Bhabha Atomic Research Centre, Mumbai 400085, India
e
Material Science Division, Bhabha Atomic Research Centre, Mumbai 400085, India
f
Solid State Physics Division, Bhabha Atomic Research Centre, Mumbai 400085, India

A R T I C LE I N FO A B S T R A C T

Keywords: A few layer/multilayer two-dimensional (2D) Tellurium (Te) nanoflakes were mechanically exfoliated and their
Tellurene structural and vibrational properties were probed. Layer thickness reduction was confirmed using optical mi-
Exfoliation croscopy, atomic force microscopy and kelvin probe force microscopy. Raman spectroscopy measurements re-
Raman spectroscopy vealed anomalous shift in the frequencies of their intrachain vibrational modes, where, after an initial increasing
Inter-chain and intra-chain interaction
trend (blue shift) with thickness reduction from bulk, they changed to decreasing trend (red shift) on further
layer thinning. Such unusual vibrational behaviour of 2D Te, exhibiting reversal from blue shift to red shift, is
attributed mainly to the deformations occurring in the Te chains on layer thinning, followed by modifications in
the local structures and bondlengths of the corresponding Te atoms. The deformation under relaxed conditions
are also confirmed from first principle calculations. Red shift can also be correlated to lesser hole concentration
for thinner flakes, as revealed by work function changes measured on the nanoflakes surfaces. Interestingly, in
spite of such anomalous behavior, the difference in frequencies of both modes showed initial increase with
thickness and saturation beyond a certain thickness, as generally observed in other 2D materials. Temperature
dependent Raman spectroscopic measurements further confirmed anomalous behaviour for thinner flakes.

1. Introduction same for another class of 2D materials, namely Xenes or mono-ele-


mental 2D materials, are scarce [5–9]. While graphene and group VI A
The discovery of graphene in 2004 and its exotic material properties materials like silicene, germanene etc. are reported to be semi-metals
that ensured its success in various application fields, has directed the and borophene to be metallic, only phosphorene, arsenene and anti-
interest of the researchers towards other two dimensional (2D) mate- monene were suggested to be semiconducting, with their band gap
rials [1]. Most of the 2D materials are generally layered materials opening when thinned down to a monolayer [6,10]. Among these, only
having covalent bonding in the plane of the layers with the individual phosphorene has been widely explored as a p-type semiconductor in
layers held together by Van der Waals bonding. These materials have various applications due to its layer dependent direct band gap from
electronic properties different from their bulk counterparts because of 0.3 eV (bulk) to 1.5 eV (monolayer) [7,11–14]. But the major dis-
quantum confinements of carriers in one dimension, varying from advantage of phosphorene is that it reacts strongly with moisture and
metallic to semiconducting to insulating [2–4]. Therefore, for applica- oxygen from the ambient, thereby forming phosphoric acid and other
tions of these 2D materials in various technologies, it is imperative to related species, leading to degradation of the devices within hours of
explore their intrinsic properties. Though vast literature is available on fabrication [8,13]. Focus was shifted to antimonene whose isolation
the material properties of mono/multilayers of transition metal di- into thin layers is still at an infant stage and the attempts to fabricate
chalcogenides (TMDs), nitrides and carbides etc. studies exploring the gated devices using antimonene has not yet been successful [6,8,12].


Corresponding authors at: Technical Physics Division, Bhabha Atomic Reserach Centre, Mumbai 400094 India (N. Padma). Solid State Physics Division, Bhabha
Atomic Research Centre, Mumbai 400094, India (R. Rao).
E-mail addresses: padman@barc.gov.in (N. Padma), rekhar@barc.gov.in (R. Rao).

https://doi.org/10.1016/j.apsusc.2020.147303
Received 20 April 2020; Received in revised form 26 June 2020; Accepted 16 July 2020
Available online 21 July 2020
0169-4332/ © 2020 Elsevier B.V. All rights reserved.
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Other Xenes like silicene, germanene and borophene are reported to be previous reports have shown continuous blue shift of both A1 and E(2)
substrate dependent 2D materials that are unstable in free standing peaks with reduction in thickness of tellurene flakes, our study shows
form [15]. Therefore, the hunt for other mono-elemental 2D materials an anomalous behavior of the same, which is discussed here.
is still being pursued and, in this context, 2D form of Tellurium (Te) has
caught the attention of researchers recently. 2. Experimental methods
Tellurium belongs to the family of chalcogens and has trigonal
crystal lattice, where the bulk Te is an indirect band gap p type semi- 2.1. Exfoliation and characterization of 2D Te flakes
conductor with a band gap of 0.35 eV. Due to the small spin orbit
coupling, bulk Te has high hole mobility and small effective mass The SiO2 (270 nm)/Si substrates were cleaned by ultra-sonication in
[16,17]. It is basically a one-dimensional van der Waals solid con- trichloroethylene, acetone and methanol for 10 min each. A single
taining helical chains of Te atoms bonded by strong covalent bond crystal whisker of tellurium, synthesized as reported earlier [35], was
along the chain and each of the neighbouring chains bonded by weak initially slid on cleaned SiO2 (270 nm)/Si substrate to exfoliate flakes
van der Waals interaction [15,18]. The large structural anisotropy of Te [18]. Fresh SiO2/Si substrates were subsequently slid on the first sub-
favours the formation of one dimensional nanostructures which has strate with exfoliated flakes. This process was repeated two to three
been explored in a number of studies where such nanostructures are times in order to reduce the thickness of flakes each time. As different
synthesised using different methods [19–24]. Bulk Te was not con- colored flakes have different thicknesses, the exfoliated flakes were first
sidered to be a layered material, until recently, when some groups have observed using Olympus optical microscope with different objectives to
predicted this one- dimensional van der Waals solid to be also stable in identify the optical colour contrast of these flakes. Atomic force mi-
2D form, called as tellurene and have convincingly included the same in croscopy (AFM) images and kelvin probe force microscopy (KPFM)
the family of 2D materials [25–27]. These groups theoretically pre- measurements were recorded by Asylum Research Oxford Instrument
dicted tellurene to be existing in three forms [25]. The α-Te phase, (MFP3D) in non-contact mode. The WSxM software was used for the
which is energetically most stable, is suggested to be having 1 T MoS2 roughness analysis of AFM images and contact potential difference
like structure with three Te atoms per unit cell and two Te atoms having (CPD) analysis of KPFM images of tellurium flakes. For recording TEM
different coordination. The metastable β phase of tellurene was pre- images the nanolayers were prepared by initially exfoliating Te flakes
dicted to be composed of planar four members and chair like six on sapphire substrates and subsequently sonicating these substrates in
member rings arranged alternately. The metastable γ tellurene has 2H isopropyl alcohol for 10 min. A few μL of the sonicated dispersion was
MoS2 like structure. The α and β phases are suggested to be semi- drop cast on carbon coated copper grid. Field emission scanning elec-
conducting and γ phase to be metallic. Since then, the promising tron microscopy (FESEM) was further used to verify the morphology
properties of Te such as thermoelectricity, piezoelectricity, photo- and the size of the nanoflakes. TEM investigation was carried out under
conductivity, etc. have motivated the research community to begin Tecnai T20 microscope having LaB6 filament and images were recorded
exploring the isolation/synthesis of 2D tellurene and ultrathin Te films digitally using 2Kx2K camera. For the analysis of the images DM soft-
[15,18,19,25,28–32]. Huang et al and Chen et al have reported the ware freely provided by Gatan was used. The Raman spectra of colored
growth of epitaxial β-Te films on the surface of graphene and highly tellurium flakes were recorded by Jobin Yvon HR-800 evolution Raman
oriented pyrolytic graphite (HOPG) respectively, by molecular beam spectrometer calibrated with first order silicon peak (520.9 cm−1).
epitaxy (MBE) [28,31]. Wang et al have reported successful synthesis of Raman spectroscopic measurements were carried out at room tem-
tellurene by substrate free solution process [15]. They demonstrated perature and pressure using 632.8 nm laser and a grating of 1800 l/mm
fabrication of tellurene based field effect transistors (FETs) exhibiting to get a dispersion of 0.35 cm−1/pixel. The laser power was kept at 0.3
ON/OFF ratio of the order of 106 charge carrier mobility of the order of mW in order to avoid any sample heating. To identify the orientation of
700 cm2 V−1s−1 and improved stability. Apte et al have grown ultra- the c-axis in a tellurium flake, angle resolved polarized Raman spec-
thin Te films on MgO and SiO2 substrates using pulsed laser deposition troscopic measurements were carried out by rotating the incident po-
(PLD) and physical vapour deposition (PVD) respectively [32]. larization using a half-wave plate. Low temperature Raman spectro-
The feasibility of isolation of these layered materials into mono/ scopic measurements were carried out using Linkam THMS 600
multi layers allow them to exhibit layer thickness dependent band gap/ temperature stage with a temperature stability ± 0.5 K on the sample.
electronic properties, which are well exploited in the applications like
optoelectronics. Tellurium has also been reported to exhibit layer 2.2. Theoretical methods and models
thickness dependent band gap ranging from that corresponding to mid
infra-red region for the bulk to near infra-red region for monolayer The total energy calculations were performed at the first-principles
(~1.13 eV) [15,18,28]. As the thickness of the layers plays an im- density functional theory (DFT) level by employing the Perdew-Burke-
portant role in many applications like FETs, solar cells etc., it is ne- Erzerhoff (PBE) exchange–correlation functional as implemented in the
cessary to identify the thickness of the layers for each of the 2D ma- ORCA suits of electronic structure theory based program package. Full
terials. The layer thickness dependent vibrational properties have geometry optimizations of the Te layers have been performed without
enabled the use of Raman spectroscopy as a successful tool to determine any symmetry constraints with the Los Alamos National Laboratory
the thickness of the layers in these materials, with the difference in the based valence double-zeta (LANL2DZ) basis set and in this calculation,
peak positions of in-plane and out-of-plane vibration modes being used for each tellurium atom, the Hay-Wadt based effective core potential
as the markers [33]. Though layer thickness dependent studies and with the associated basis set is used for the core electrons. All the atoms
markers are almost identified for most of the materials, such studies for were allowed to relax during the geometry optimization until a self-
tellurene are sparse [15,34]. Churchill et al have isolated Te sheets by consistent-field convergence criterion of 10-8 hartree on total energy is
mechanical exfoliation and have shown hardening of the Raman modes achieved. The grid based DFT has been used as present in ORCA, that
of an isolated flake as compared to that of the bulk Te [18]. Apte et al employs a typical grid quadrature to compute the integrals. During the
have also reported a systematic hardening of the A1 and E(2) Raman SCF procedure, the grid consists of 96 radial shells with 36 and 72
modes with reduction in thickness of PLD deposited Te films [32]. angular points. This level of theory has been employed for the de-
Wang et al [15] and Peng et al [34] have reported systematic thickness termination of energy, geometries and properties of variety of systems
dependent Raman shifts for solution processed tellurene, in this study, and the calculated properties have been found to be reliable within
we have attempted to isolate the layers from the bulk by mechanical 2–5% error. Since these layered structured are known to be stabilized
exfoliation, as reported by Churchill et al [18] and have investigated the by the weak van der Waals forces, the atom-pairwise dispersion cor-
thickness dependent vibrational behavior of the same. While the rections to the DFT energy with Becke-Johnson damping has been

2
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 1. (a) Helical chains of Te forming a layered structure, (b) Projection of Te chains forming a hexagonal structure, (c-e) optical images of Te flakes with different
thicknesses and optical contrast, (f-h) Atomic force microscopy images of Te flakes with different thicknesses and (i-k) corresponding height profiles along the lines
across the flakes.

employed. metastable tetragonal and 2H-MoS2 type structures with semi-


The bulk Te geometries have been initially generated using the conducting and metallic behavior respectively. While most of the stu-
single crystal X-ray data and the corresponding layered structures were dies have focussed on synthesis and applications of 1D nanostructures
formed by aligning c-lattice of the bulk structures. Herein, the first of Te [36,37], only a few studies have reported preparation of 2D Te
monolayer is generated by combining three single chains orienting to- films and flakes.
wards the a and b lattice axis. Subsequently, the second and other layers Wang et al [19] have demonstrated first the formation of 2D hex-
up to 5 were initially generated in the similar way. For the multi- agonal Te nanoplates when deposited on mica substrate by van der
layered cases, each layer is arranged with respect to the bulk unit cell Waals epitaxial technique, with c axis perpendicular to the substrate.
dimension along the a or b axis. Though Zhu et al [25] have suggested α Te to be more stable, they have
confirmed the epitaxially grown Te nanolayers on HOPG by MBE
method in their study to be in β phase, with c-axis lying parallel to the
3. Results and discussion
substrate. Many other groups that have also prepared/synthesized Te
films/flakes through different methods like MBE, physical vapour de-
As mentioned before, Te is a one-dimensional van der Waals solid
position, mechanical exfoliation, substrate free solution process, liquid
containing helical chain of Te atoms bonded by strong covalent bond
phase exfoliation (LPE) etc. have also reported c-axis to be parallel to
along the chain, with each of the neighbouring chains bonded by weak
the substrate [15,18,19,25,28–31]. In the present study, the mechani-
van der Waals interaction (Fig. 1a). The projection of each of the chain
cally exfoliated Te flakes, similar to that adapted by Churchill et al [18],
forms a hexagonal lattice in the ab plane, as shown in Fig. 1b. The
are also expected to have c-axis parallel to the SiO2 substrate, as
monolayer form of Te has been predicted to be stable in three phases (α,
mentioned in the above studies. Fig. 1c-e shows the optical images of
β and γ) with α Te to be most stable having 1 T-MoS2 like structure and
some of those exfoliated Te flakes with different colours. It has been
semiconducting in nature [25]. The β and γ Te are suggested to be

3
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 2. FESEM images of (a) Nanoflake 1 and (b) nanoflake 2, EDX intensity mapping of nanoflake 1 at (c) and (d) Te and O elements respectively, EDX intensity
mapping of nanoflake 2 at (e) Te element.

well established that the flakes with different thickness exhibit different are found to be of good crystallinity. It can be seen that the elemental
colours and varying optical contrast with the SiO2 substrate [33,38]. intensity mapping shows uniform distribution of Te atoms in both
Therefore, the dark coloured flakes can be expected to represent thicker thinner and thicker flakes, and the absence of TeO, TeO2 etc.
flakes while those with colours close to that of the substrate could be Kelvin Probe Force Microscopy (KPFM) can be used for quantitative
due to thinner flakes. Hence, the flake in Fig. 1e is suggested to be mapping of contact potential difference (CPD) and work function dis-
thinner than those shown in Fig. 1c and d. tribution of 2D materials surfaces by mapping the local surface poten-
The layer thickness of Te flakes was estimated using AFM images, tials. It has been shown in earlier reports that the thickness variation in
where the step height produced by the different layers of these mate- 2D materials reflects as change in the work function, as measured using
rials on the substrate was measured. Fig. 1f-h show the images of Te KPFM [39]. The present study uses KPFM as another tool to identify the
flake with different thicknesses and Fig. 1i-k show their corresponding change in thickness of the Te flakes by estimating the change in work
height profiles along the lines drawn across the flakes in each of these function of the same. For this purpose, the topography and surface
figures. The height profiles show the thickness of the flake in Fig. 1f to potentials of tellurium flakes were simultaneously recorded. Fig. 3a
be around 80 nm while those shown in Fig. 1g and h are found to be shows the KPFM image of Te flakes that were seen with different col-
about 23 and 7 nm respectively. As the thickness of the flakes decreases, ours, using optical microscopy as shown in Fig. 1. In this image, the
the lateral sizes of most of the flakes are also observed to generally dark points correspond to the Te flakes having larger negative contact
decrease, ranging from about 6–8 µm for thicker flakes (also seen from potential difference. The work function of Te flakes were determined
optical microscopy images) to less than 1 µm for thinner flakes. using the formula
The morphology of the exfoliated Te flakes of different lateral di-
mensions was verified using FESEM imaging, as shown in Fig. 2a (na- ∅sample = ∅tip − exCPD (1)
noflake 1) and b (nanoflake 2). The EDX elemental mapping displayed
in Fig. 2 c and d represents the distribution profile of Te and O for where∅sample and∅tip represent the work functions of sample and tip re-
nanoflake 1 (Fig. 2a) and that in Fig. 2e shows the distribution profile of spectively, e is the electronic charge. The work function of the flakes is
Te for nanoflake 2 (Fig. 2b). It can be seen that while nanoflake 1 has determined with respect to work function of Titanium Iridium tip
length around 2 µm, the other has dimensions of around 500 nm. As (4.9 eV). Fig. 3b shows the variations in work function of Te flakes with
seen in optical and AFM images, the images of the flakes further con- layer thickness that is observed to be increasing with thickness of the
firm the growth direction or the c-axis to be parallel to the substrate and flakes by about 0.1 eV. Increase in work function by about 0.15 eV was
also made in an earlier study on MoS2 which attributes it to the

4
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 3. (a) KPFM image of the tellurium flake on SiO2/Si substrate and (inset) CPD variation of this flake along the line drawn across the flake, (b) Variation of work
function of tellurium flakes with thickness. Continuous line is a guide to eye. (c) TEM image of one of the flakes and (d) corresponding HRTEM image, inset- top left-
SAED pattern, top right- FFT of the selected area shown in white box, bottom right- inverse FFT of selected area.

increasing hole doping with layer thickness [39]. In the present study representation Γ = A1 + A2 + 2E [41]. A1 is a Raman active mode, A2
also, increase in hole doping could be causing the increase in work is an infrared active mode and doubly degenerate E mode is both
function. Since Te is reported to be a p type semiconductor, such in- Raman and infrared active mode. Generally, in 2D layered materials
crease in work function clearly indicates the increase in layer thickness. like TMDs, Ag mode is the out of plane vibrational and Eg modes are the
Transmission electron microscopy (TEM) has been successfully used in plane vibrational modes. In the case of tellurium, A1 mode is an intra-
for lattice parameter and crystal system determination of 2D materials. chain vibrational mode (Fig. 4a) with eigen vectors along a axis in the
Fig. 3c shows a TEM image of one of the typical Te flakes. The high ab basal plane [15,36]. E modes are the two doubly degenerate modes
resolution TEM (HRTEM) image of the flake is shown in Fig. 3d and with E(1) consisting of rotation around a and b -axis and E(2) is an intra-
corresponding SAED pattern of the flake, with planes indexed, is shown chain asymmetric stretching vibrational mode within a single layer of
as the top left inset of Fig. 3d. The HRTEM image shows presence of tellurium (tellurene) with eigenvectors mainly along c- axis (Fig. 4b).
fringe patterns at certain places; where one of those areas is selected as Fig. 4c shows the unpolarized Raman spectra of bulk Te crystal at
shown in white box in Fig. 3d. The corresponding Fast Fourier trans- ambient temperature and pressure, which reveals one Raman active
form (FFT) and the inverse FFT of the selected area are shown as top non-degenerate A1 mode at 121.5 cm−1 and two degenerate E modes at
right and bottom right insets respectively. Both SAED pattern and FFT 92 cm−1 (E(1)) and 141.3 cm−1 (E(2)) as reported earlier [15,36]. The
show bright diffraction spots, indicating that the exfoliated Te flake is A1 mode was found to have higher intensity in comparison to E(2)
single crystalline. The spacing between the fringes as seen from inverse mode. Fig. 5a shows the Raman spectra of tellurium flakes having
FFT was estimated to be about 0.39 nm which is close to the interplanar different thicknesses. With reduction in the thickness of Te flakes,
spacing of (1 0 0) planes of Te. The inverse FFT image also shows mild Raman spectra were found to clearly exhibit changes in peak fre-
deformations in Te chains and hence local structure variations. quencies, intensity and the full width at half maximum (FWHM) as
Layer thickness dependent vibrational behavior of these mechani- compared to that of the bulk Te. The Raman spectra clearly show the
cally isolated Te flakes were investigated using Raman spectroscopy. thickness dependent shift in the frequencies for both A1 and E(2)
Tellurium crystallizes in trigonal structure (space group D34 ) containing modes.
three atoms per unit cell arranged helically about the c-axis [40]. Group In the Raman spectra of Te flakes with lower thicknesses only A1
theory predicts 9 normal vibrational modes of which three are acous- mode and one of the E modes, around 141 cm−1, is visible with de-
tical modes and six are zone centred optic modes having irreducible creasing intensities. As described earlier, optical microscopy, FESEM

5
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 4. (a, b) Vibration of Te atoms in A1 and E(2) modes in the basal plane of tellurium respectively, (c) Raman spectra of bulk Te crystal.

and AFM images show the flakes to be elongated in the direction par- revert to decreasing trend with the shift of about 5 cm−1 for 7 nm thick
allel to the substrate. The orientation of c-axis was further verified using flakes, from the maximum value. In the case of A1 mode also, such
angle resolved polarized Raman spectroscopy measurement of a typical anomalous behavior is observed with layer thinning, where the max-
flake (Fig. 5b). The polarization of the excitation wavelength was ro- imum blue shift is observed to occur at around 23 nm thickness beyond
tated and the total integrated intensity of the Raman modes over the which the peaks show a red shift. In the case of TMDs like MoS2, Ag
entire range shown in Fig. 5a is plotted in Fig. 5b. The Raman mode mode is reported to decrease in frequency and Eg mode to harden as
intensity showed two maxima at 0° and 180° within one full rotation. they are thinned from bulk to monolayer [33,42,43]. While decrease of
This implies that the flakes in the present study have c-axis parallel to Ag mode is attributed to the reduced van der Waals force and hence
the substrate as also reported by Churchill et al [18]. The measured easier vibration of the Mo atoms out of plane, stiffening of Eg mode is
polarization dependence of Raman intensity is fitted with expected suggested to be due to the effective long range coulombic interactions
polarization dependence of Raman intensity from Raman tensor cal- and to changes in the intralayer structure causing increase in the
culation as provided in the Supporting Information. strength of the covalent bonding between the atoms [38,42]. In the case
Fig. 6a shows the shifts in the frequencies of A1 and E(2) modes as a of mono-elemental 2D materials like antimonene, both Ag and Eg modes
function of thickness of tellurium flakes (with a maximum error of ± were reported to show a large shift to high frequency for reduction in
0.3 cm−1 for A1 and ± 0.4 cm−1 for E(2) modes considering all the thickness from bulk to monolayer, as also observed for initial reduction
thicknesses) revealing an anomalous behavior. It can be seen from this in the thickness in the present study [6]. In all these 2D materials, the
figure that for initial reduction in layer thickness from the bulk, both A1 modes either shift to higher or lower frequencies throughout, on re-
and E(2) modes of nano-layered flakes exhibit hardening. E(2) mode at ducing from bulk to monolayer, unlike the anomalous reverting of the
141 cm−1 showed a maximum hardening of about 3.1 cm−1 for a direction of shift observed in the present study. To the best of our
thickness of around 35 nm. Beyond this thickness, E(2) mode is found to knowledge, this anomalous behavior of A1 and E(2) modes are not

Fig. 5. (a) Raman spectra of different thicknesses of Te flakes, dotted lines are guide to eye, (b) Polar plot of total integrated Raman intensity for flake of thickness
80 nm of the range shown in (a), lines with symbol represent experimental data and the solid line represents the fitting.

6
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 6. Variation with thickness of Te flakes- (a) peak positions of A1 and E(2) modes, with error of a maximum of ± 0.3 cm−1 for A1 and ± 0.4 cm−1 for E(2)
modes, (b) difference in peak positions of E(2) and A1 modes (E(2)-A1), (c and d) FWHM of A1 and E(2) modes respectively. Continuous lines are guide to eye.

reported in any 2D materials. of bulk Te, which show hardening of A1 and E(2) modes to higher
As mentioned before, there are only three studies showing the effect frequency, as also seen in the present study around that thickness range.
of layer thinning of Te on Raman shift, with two of them discussing It has to be pointed out that Churchill et al have not shown Raman
systematically about the layer thickness dependent Raman shift spectrum of flakes of lower thicknesses, whereas anomaly in the vi-
[15,34]. Wang et al and Peng et al have carried out systematic studies brational behaviour, showing softening of both the modes, is observed
on the effect of thickness on Raman peaks using Te flakes synthesized for thicknesses below ~ 23 nm in the present study.
by solution route [15,34]. Wang et al [15] have reported an increase of In the present study also, the initial high frequency shift of both the
A1 and E(2) mode frequencies by about 14 and 6 cm−1 respectively and modes when reduced from bulk, could be due to reduced dielectric
Peng et al [34] have also shown similar results. Churchill et al have screening of long-range coulomb interaction and structural changes in
isolated thin layers of Te by mechanical exfoliation as also exfoliated in Te-Te bonds. As the A1 and E(2) modes in Te are due to intrachain
the present study and have also shown a hardening of both the modes vibrations of same Te atoms in different directions respectively, (Fig. 4a
for a 30 nm flake as compared to that of the bulk [18]. Apte et al have and b), hardening of A1 could also be attributed to increased intra chain
also reported blue shift in the Raman modes of PLD deposited ultrathin interaction, as also reported in the above mentioned studies. The shift
Te films, with reduction in thickness [32]. Though their study suggests to lower frequencies of both the modes, after a maximum hardening as
the c-axis to be orthogonal to the substrate surface, it should be noted shown in Fig. 6a, cannot be connected to increased intra chain inter-
that the flakes synthesized in most of the studies are reported to have c- actions. Though the substrate–layer interaction may be expected to
axis parallel to the substrate, as in the present study. It can be pointed become prominent below a certain layer thickness, it is also mostly
out that all these studies have shown stiffening of both E(2) and A1 suggested to stiffen the A1 and E(2) modes, unlike that observed in the
modes with thinning down of bulk Te. Churchill et al attributed the present study [18]. It can be suggested at this point that the attenuated
stiffening of E(2) modes of 30 nm thick Te flakes to mainly substrate van der Waals interaction between the Te chains belonging to different
flake interaction [18]. Wang et al and Apte et al have attributed hard- layers, brought about by layer thinning, can be expected to create de-
ening of E(2) mode to enhancement of long-range interlayer coulombic formation in the chains of Te. This deformation or the local structure
interactions on thinning down, similar to the cases for other 2D mate- variations could be dominating over the increased intra chain interac-
rials like TMDs, phosphorene etc. [15,32]. Peng et al have attributed tion suggested for initial thickness reduction from bulk (blue shift re-
high frequency shift of E(2) modes to laminating induced structural gions), altering the bond length randomly. Irregularity in the fringes
changes [34]. The blue shift/stiffening of A1 mode in thin layers of Te, observed in the HRTEM image (Fig. 3d) could be taken as corroborative
which is unlike that observed for other 2D materials like MoS2, have evidence of the deformation of the chains. It may be noted that KPFM
also been attributed to increased intra chain forces, in these studies. It studies (Fig. 3b) indicated a drastic variation of charge carrier con-
should be mentioned here that while the method of exfoliation in the centration below about 30 nm thickness. It is well known that inter-
present study is similar to that of Churchill et al [18], such an anom- action of charge carriers with phonons results in modification of
alous behavior is not reported by them. In the reports of Churchill et al, phonon frequency as well as FWHM [44]. The change in carrier con-
Raman spectrum of nanoflakes of around 30 nm are compared with that centration in thinner flakes, as noted from the lesser work function seen

7
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

in KPFM analysis (Fig. 3b), could be resulting in variation in phonon in Raman frequency of A1 and E(2) modes to be lower than that of the
spectra and could be further contributing to red shift of Raman modes. bulk. Though the lowest thickness of the exfoliated flake for which
Hence, the resultant reduction in Raman mode frequencies could be a Raman modes could be measured in the present study is around 5 nm
combined effect of changes in the structure of the chains as well as due (about10 layered systems), the obtained E(2) mode peak frequency in
to changes in charge carrier concentration. Fig. 6a, falling below that of the bulk, is found to correlate well with the
As the difference in peak positions of A1 and E(2) modes are gen- above mentioned theoretically observed increase in bond length.
erally used as a facile tool to estimate the layer thickness of 2D mate- For bulk Te, similar to intra-chain covalent bonds, interchain dis-
rials, a plot of the difference in A1 and E(2) peak positions of Te flakes, tances are also observed to be with two different sets of values such as
against the layer thickness is shown in Fig. 6b. While this frequency 3.3 Å and 4.3 Å. The first set with lower value (3.3 Å), which is the next
difference remains constant upto about 30 nm as reduced from the bulk, nearest distance d1, corresponds to that between two Te atoms be-
it shows a clear reducing trend below this thickness. A similar variation longing to two neighbouring chains that come closer to each other in
is also observed for other 2D materials like MoS2, antimonene etc. while their spiralling process [47]. The other set with higher value (4.3 Å) is
opposing trend is also reported for phosphorene [6,33,45]. Therefore, the distance d2 between two similar Te atoms, for example, between Te1
this study shows that even though the peak shifts of E(2) and A1 show atoms (Fig. 4a) belonging to the neighbouring chains. The variations in
an anomalous behavior, the frequency difference follows a common the average values of these distances d1 and d2 as a function of number
trend and hence can be used to determine the thickness of Te layers. of layers, are shown in Fig. 7d and e respectively. As mentioned before,
The FWHM of Raman modes were also systematically studied with since covalent interactions and the bond lengths between the same Te
layer thickness, as shown in Fig. 6c and d for A1 and E(2) modes re- atoms are responsible for both A1 and E(2) Raman modes, changes
spectively. The FWHM of Raman mode is reported to be affected by observed in A1 and E(2) modes should also be in similar direction, with
many factors including defects and disorder. The electron phonon additional contributions from d1 and d2. Since d1 and d2 are also found
coupling has been suggested to play a role in the broadening of Raman to increase with decrease in thickness upto 2 or 3 layers respectively,
modes in 2D materials [46]. The thickness of the flakes also is reported they would further support reduction of A1 as well as E(2) peak fre-
to affect the FWHM of two dimensional materials [34]. Lee et al have quencies with reduction in layer thickness, as observed in this study in
reported increase in the line-width of both Ag and Eg modes with de- the lower thickness regions. The initial increase of A1 and E(2) modes
creasing layer thickness of MoS2, as also observed in the present study. frequencies and reverting below a certain thickness, results in the ex-
Reduction in layer thickness could result in varying structural changes istence of a maximum in both cases. To our knowledge, this behavior
for the outer and inner layers of Te flakes, leading to varying force has not been reported in other 2D materials and can be attributed to the
constants. This could be giving rise to increase in FWHM. Deformation chain like structure of Te and deformation in the same.
in chains or random variations in local structure can also be causing To additionally verify the effect of layer thinning on temperature
larger FWHM. dependent behaviour of Raman modes, measurements were carried out
To gain further insight into these changes taking place in the chains, on bulk and two exfoliated Te nanoflakes with thicknesses around 80
which are suggested to be leading to anomalous layer thickness de- and 23 nm. The A1 mode intensity mapping of these two nanoflakes,
pendent shift in Raman mode positions, first principles calculations along with their respective optical images in the inset, are shown in
were carried out for systems with monolayer to five layers thickness, Fig. 8a and b respectively. Fig. 8c shows the Raman spectra of one of
with each layer containing three adjacent chains. As is well known, the exfoliated thin tellurium flakes (thickness ~ 23 nm) at four different
bulk Te has helical chain like structure with nearest neighbouring temperatures during cooling cycle. Similarly, Raman spectra of the
atoms covalently bonded to each other. The three Te atoms in the he- thicker nanoflake (~80 nm) and the bulk, at different temperatures are
lical cycle, marked as Te1, Te2 and Te3 in Fig. 4a are separated with the shown in Fig. 8d and e respectively. The corresponding plots for the
bond lengths l1, l2 and l3, which are found to be equal (2.91 Ả) [25]. A temperature dependence of the Raman mode frequencies for all the
typical three layers structure used in our calculations in unrelaxed three are given in Supporting Information (Fig. S3). From Fig. 8c, it can
condition is shown in Fig. 7a. The calculations show that on thinning be seen that the Raman intensity of both A1 and E(2) modes of the
down to 5 layers or below and the system allowed to relax, these bond nanoflake with thickness ~ 23 nm, shows enormous increase with de-
lengths become different with l1 and l2 increasing above 3.0 Ả and l3 creasing temperatures. It may be noted that this enormous increase in
decreasing below that in the bulk. These changes in bond lengths or the integrated intensity with decreasing temperature is not observed for
random variations in the local structures, also reported by Zhu et al bulk or thicker flakes. In all the three cases, Raman modes hardened
[25], could be attributed to the deformation occurring in the chains with lowering of temperature regardless of thickness as seen in Fig. S3.
(Fig. 7b). The remaining one, two, four and five layered structures, and But we have observed thickness dependent variation in the integrated
the cross-sectional view of all the layered structures under relaxed total intensity of the two prominent modes, i.e., A1 and E(2) modes. In
condition, are shown in Fig. S1 and S2 respectively, in the Supporting the case of thin flake, there in an increase in integrated total intensity of
Information. the Raman modes about 6 times with cooling whereas it increases only
While Zhu et al [25] have reported l1 and l2 to be equal, our results marginally (by a factor of about 1.3) in the case of bulk/thick flakes.
show slight variations between the same that are also found to be Apte et al reported increase in Raman intensity at low temperatures
slightly different for different chains. Since the Raman modes A1 and E for PVD grown Te flakes during both heating (100 K-350 K) and cooling
(2) mainly depend on vibrations involving l1 and l2, we have ignored l3 cycles (RT-100 K). They found the A1 mode intensity to be affected
in further discussions. The average of l1 and l2 (l) for all the chains in more than the E mode and attributed the increase in intensity to lattice
each layered structure is shown as a function of layer thickness (number strain relaxation at lower temperatures. They also suggested that this
of layers) in Fig. 7c. It can be seen that l exhibits a tendency to increase relaxation is more for in-plane mode than the out-of-plane mode. In the
slightly from 5 layers to 3 layered structure, and to decrease sharply present study, larger increase in intensity with decreasing temperature
below this thickness, exhibiting a maximum at certain layer thickness. is observed for thinner nanoflake, than that for the thicker nanoflake
The observed trend of increasing bond length would result in lowering and the bulk. This anomalous behaviour could be due to the stronger
of frequency of A1 and E(2) modes in the Raman spectra. Such a trend effect of local structure variation/deformation in the thin flake than in
matches with the decrease in frequency of these Raman modes with the thicker nanoflake and the bulk, as also observed for thickness de-
thinning down, observed for lower thickness flakes (after reversal of pendent Raman measurements (Fig. 5a). Unlike the observation made
direction from blue shift to red shift) as shown in Fig. 6a. It can also be by Apte et al, we observe both A1 and E(2) modes exhibit similar in-
noted from Fig. 7c that the bond length l shows a value larger than that crease in intensity with lowering of temperature. This reaffirms our
of the bulk (which is about 2.91 Å) for upto bilayer structures, resulting earlier observation from polarized Raman and thickness dependent

8
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 7. Three layered structure of Te under–(a) unrelaxed and (b) relaxed conditions, variations of- (c) average bond length l, (d) average interchain distance d1 and
(e) average interchain distance d2, as a function of layer thickness. Black solid lines in c, d and e are guide to eye.

Raman measurements, along with imaging techniques, that c-axis is range coulomb interaction and restructuring of covalently interacting
parallel to the substrate in exfoliated flakes and hence, both the modes Te atoms, the subsequent decrease was attributed to deformation that
are affected by deformation in similar fashion. could be occurring in Te chains with further reduction in the layer
Therefore, it can be mentioned that the low temperature Raman thickness, leading to local structure variations. The trend in shifting of
spectroscopic measurements further show the significant influence of both A1 and E(2) modes in same directions, was suggested to be due to
layer thinning on temperature dependent behaviour of Raman modes the involvement of the same Te atoms vibrating along and perpendi-
and hence add to the suggestion of increased local structure variations/ cular to the chains respectively. The decreasing trend of these modes
deformation in thinner flakes, resulting in the anomalous thickness frequencies at low thickness region was also confirmed by the corre-
dependent vibrational behaviour of Te flakes. These results are similar sponding changes in bond lengths and inter-chain distances observed by
to that reported by Hussain et al wherein they observe stronger effect of first principle calculations. Additionally, it is also suggested that the
lattice strain on the Raman spectra of thinner flakes [48]. change in carrier concentration for thinner flakes could further be
contributing for the red shift in the Raman modes. Our study also shows
that though the vibrational modes individually show anomalous be-
4. Conclusions
havior, the difference in their peak positions shows an increase with
increase in layer thickness at low thickness region, tending to exhibit a
In summary, a few layered flakes of tellurium were successfully
saturation beyond a certain thickness. Such a behavior is also reported
isolated by mechanical exfoliation of single crystals of Te on SiO2/Si
for 2D materials like MoS2, and is suggested to be useful for de-
substrates. The thinning down of Te flakes from bulk was identified
termining 2D Te layer thickness. Temperature dependent Raman mea-
using optical microscopy and further confirmed using AFM and KPFM
surements also showed variations with respect to layer thickness, ex-
techniques, from the changes in the height profile in the former and the
hibiting anomalous behaviour for thinner flakes.
work function in the latter. The peak frequencies and FWHM of Raman
active A1 and E(2) modes were found to be varying with thickness of Te
flakes. Though frequencies of both A1 and E(2) modes were found to be Declaration of Competing Interest
increasing initially on thinning from bulk, they exhibited an anomalous
behavior by reverting to decreasing trend beyond a certain layer The authors declare that they have no known competing financial
thickness, which has not been reported in any 2D materials. While the interests or personal relationships that could have appeared to influ-
initial increase was attributed to reduced dielectric screening of long- ence the work reported in this paper.

9
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Fig. 8. A1 mode intensity mapping of the nanoflakes with thickness –(a) ~ 80 nm and (b) ~ 23 nm, along with their respective optical images in the inset, Raman
spectra at different temperatures - of (c) exfoliated thinner nanoflake (~23 nm), (d) thicker nanoflake (~80 nm) and (e) bulk. Raman spectra have been vertically
shifted for clarity.

Acknowledgements Funding sources

The authors would like to thank Central Facility, CAFM, IIT This research did not receive any specific grant from funding
Bombay, Mumbai, India, for providing AFM and KPFM data for the agencies in the public, commercial, or not-for-profit sectors.
exfoliated Te samples. The authors would also like to thank Dr. R.
Tewari, Materials Science Division, Bhabha Atomic Research Centre,
Mumbai, for helping in the analysis of TEM data. The authors ac- Data availability
knowledge Dr. D. Sen and Dr. J. Bahadur, Solid State Physics Division,
Bhabha Atomic Research Centre, Mumbai, for FESEM and EDX mapping The raw/processed data required to reproduce these findings cannot
data. be shared at this time as the data also forms part of an ongoing study.

10
R.A. Yadav, et al. Applied Surface Science 531 (2020) 147303

Credit author statement 112 (2008) 11314–11318.


[23] B. Mayers, Y. Xia, Formation of tellurium nanotubes through concentration deple-
tion at the surfaces of seeds, Adv. Mater. 14 (2002) 279–282.
All authors equal contribution. [24] Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim, H. Yan, One
dimensional nanostructures: synthesis, characterization and applications, Adv.
Appendix A. Supplementary data Mater. 15 (2003) 353–389.
[25] Z. Zhu, X. Cai, S. Yi, J. Chen, Y. Dai, C. Niu, Z. Guo, M. Xie, F. Liu, J.H. Cho, Y. Jia,
Z. Zhang, Multivalency-driven formation of Te based monolayer materials: a
Supplementary data to this article can be found online at https:// combined first-principles and experimental study, Phys. Rev. Lett. 119 (2017)
doi.org/10.1016/j.apsusc.2020.147303. 106101.
[26] B. Wu, X. Liu, J. Yin, H. Lee, Bulk β-tellurenes: indirect to direct band-gap transi-
tions showing semiconducting property, Mater. Res. Express 4 (2017) 095902.
References [27] Y. Pan, S. Gao, L. Yang, J. Lu, Dependence of excited-state properties of tellurium on
dimensionality: from bulk to two dimensions to one dimension, Phys. Rev. B 98
(2018) 085135.
[1] K.S. Novoselov, A.K. Geim, S. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos,
[28] X. Huang, J. Guan, Z. Lin, B. Liu, S. Xing, W. Wang, J. Guo, J. Epitaxial growth and
I.V. Grigorieva, A. Firsov, Electric field effect in atomically thin film, Science 306
band structure of Te on graphene, Nano. Lett. 17 (2017) 4619−4623.
(2004) 666–669.
[29] M. Amani, C. Tan, G. Zhang, C. Zhao, J. Bullock, X. Song, H. Kim, V.R. Shrestha,
[2] S. Balendhran, S. Walia, H. Nili, J.Z. Ou, S. Zhuiykov, R.B. Kaner, S. Sriram,
Y. Gao, K.B. Crozier, M. Scott, A. Javey, Solution-synthesized high-mobility tell-
M. Bhaskaran, K. Kalantar-zadeh, Two-dimensional molybdenum trioxides and di-
urium nanoflakes for short-wave infrared photodetectors, ACS Nano 12 (2018)
chalcogenides, Adv. Funct. Mater. 23 (2013) 3952–3970.
7253–7263.
[3] G. Wang, L. Bao, T. Pei, R. Ma, Y.Y. Zhang, L. Sun, G. Zhang, H. Yang, J. Li, C. Gu,
[30] Z. Xie, C. Xing, W. Huang, T. Fan, Z. Li, J. Zhao, Y. Xiang, Z. Guo, J. Li, Z. Yang,
S. Du, S.T. Pantelides, R.D. Schrimpf, H.J. Gao, Introduction of interfacial charges to
B. Dong, J. Qu, D. Fan, H. Zhang, Ultrathin 2D nonlayered tellurium nanosheets:
black phosphorous for a family of planar devices, Nano Lett. 16 (2016) 6870–6878.
facile liquid-phase exfoliation, characterization, and photoresponse with high per-
[4] V. Nicolosi, M. Chhowalla, M.G. Kanatzidis, M.S. Strano, J.N. Coleman, Liquid
formance and enhanced stability, Adv. Funct. Mater. 28 (2018) 1705833.
exfoliation of layered materials, Science 340 (2013) 1226419.
[31] J. Chen, Y. Dai, Y. Ma, X. Dai, W. Hoa, M. Xie, Ultrathin β-tellurium layers grown on
[5] Z. Lin, A. McCreary, N. Briggs, S. Subramanian, K. Zhang, Y. Sun, X. Li, N. J. Borys,
highly oriented pyrolytic graphite by molecular-beam epitaxy, Nanoscale 9 (2017)
H. Yuan, S. K. Fullerton-Shirey, A. Chernikov, H. Zhao, S. McDonnell, A. M.
15945–15948.
Lindenberg, K. Xiao, B. J. LeRoy, M. Drndi, J. C. M. Hwang, J. Park, M. Chhowalla,
[32] A. Apte, E. Bianco, A. Krishnamoorthy, S. Yazdi, R. Rao, N. Glavin, H. Kumazoe,
R. E. Schaak, A. Javey, M. C. Hersam, J. Robinson, M. Terrones, 2D materials ad-
V. Varshney, A. Roy, F. Shimojo, E. Ringe, R.K. Kalia, A. Nakano, C.S. Tiwary,
vances: from large scale synthesis and controlled heterostructures to improved
P. Vashishta, V. Kochat, P.M. Ajayan, Polytypism in ultrathin tellurium, 2D Mater. 6
characterization techniques, defects and applications, 2D Mater. 3 (2016) 042001.
(2019) 015103.
[6] J. Ji, X. Song, J. Liu, Z. Yan, C. Huo, S. Zhang, M. Su, L. Liao, W. Wang, Z. Ni,
[33] C. Lee, H. Yan, L.E. Brus, T.F. Heinz, J. Hone, S. Ryu, Anomalous lattice vibrations
Y. Hao, H. Zeng, Two-dimensional antimonene single crystals grown by van der
of single and few-layer MoS2, ACS Nano 4 (2010) 2695–2700.
Waals epitaxy, Nat. Commun. 7 (2016) 13352.
[34] J. Peng, Y. Pan, Z. Yu, J. Wu, Y. Zhou, Y. Guo, X. Wu, C. Wu, Y. Xie, Two-dimen-
[7] H. Liu, A.T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tomanek, P.D. Ye, Phosphorene: an
sional tellurium nanosheets exhibiting an anomalous switchable photoresponse
unexplored 2D semiconductor with a high hole mobility, ACS Nano 8 (2014)
with thickness dependence, Angew. Chem. Int. Ed. 57 (2018) 13533–13537.
4033–4041.
[35] S. Sen, U.M. Bhatta, V. Kumar, K.P. Muthe, S. Bhattacharya, S.K. Gupta,
[8] P. Ares, J.J. Palacios, G. Abellán, J. Gómez-Herrero, F. Zamora, Recent progress on
J.V. Yakhmi, Synthesis of tellurium nanostructures by physical vapor deposition
antimonene: a new bidimensional material, Adv. Mater. 30 (2018) 1703771.
and their growth mechanism, Crystal Growth & Design 8 (2008) 239–242.
[9] A. Castellanos-Gomez, L. Vicarelli, E. Prada, J. Q. Island, K. L. Narasimha-Acharya,
[36] Y. Du, G. Qiu, Y. Wang, M. Si, X. Xu, W. Wu, P.D. Ye, One-dimensional van der
S. I. Blanter, D. J. Groenendijk, M. Buscema, G. A. Steele, J. V. Alvarez, H. W.
Waals material tellurium: Raman spectroscopy under strain and magneto-transport,
Zandbergen, J. J. Palacios, H. S. J. van der Zant, Isolation and characterization of
Nano Lett. 17 (2017) 3965–3973.
few-layer black phosphorous, 2D Mater. 1 (2014) 025001.
[37] Z. He, Y. Yang, J.W. Liu, S.H. Yu, Emerging tellurium nanostructures: controllable
[10] H.S. Tsai, C.W. Chen, C.H. Hsiao, H. Ouyangb, J.H. Liang, The advent of multilayer
synthesis and their applications, Chem. Soc. Rev. 46 (2017) 2732–2753.
antimonene nanoribbons with room temperature orange light emission, Chem.
[38] H. Li, Q. Zhang, C.C.R. Yap, B.K. Tay, T.H.T. Edwin, A. Olivier, D. Baillargeat, From
Commun. 52 (2016) 8409–8412.
bulk to monolayer MoS2: Evolution of Raman scattering, Adv. Funct. Mater. 22
[11] Y. Yang, J. Gao, Z. Zhang, H.H. Si Xiao, Z.B. Xie, J.H. Sun, C.H. Wang, Y.W. Zhou,
(2012) 1385–1390.
X.Y. Wang, P.K. Guo, X.F.Yu. Chu, Black phosphorous based photocathodes in
[39] H. Kim, J. Lee, J.H. Kim, C.C. Hwang, C. Lee, J.Y. Park, Work function variation of
wideband bifacial dye-sensitized solar cells, Adv. Mater. 28 (2016) 8937–8944.
MoS2 atomic layers grown with chemical vapor deposition: the effects of thickness
[12] M. Fortin-Deschenes, O. Waller, T. O. Mentes, A. Locatelli, S. Mukherjee, F.
and the adsorption of water/oxygen molecules, Appl. Phys. Lett. 106 (2015)
Genuzio, P. L. Levesque, A. Hebert, R. Martel, O. Moutanabbir, Synthesis of anti-
251606.
monene on germanium, Nano Lett. 17 (2017) 4970−4975.
[40] A.S. Pine, G. Dresselhaus, Raman spectra and lattice dynamics of tellurium, Phys.
[13] L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H.Chen, Y. Zhang, Black
Rev. B 4 (1971) 356–371.
phosphorous field-effect transistors, Nat. Nanotechnol. 9 (2014) 372−377.
[41] C. Marini, D. Chermisi, M. Lavagnini, D. Di Castro, C. Petrillo, L. Degiorgi,
[14] G. Abellan, P. Ares, S. Wild, E. Nuin, C. Neiss, D. R. S. Miguel, P. Segovia, C. Gibaja,
S. Scandolo, P. Postorino, High-pressure phases of crystalline tellurium: a combined
F. G. Michel, A. Gcrling, F. Hauke, J. Glmez-Herrero, A. Hirsch, F. Zamora,
Raman and ab initio study, Phys. Rev. B 86 (2012) 064103.
Noncovalent functionalization and charge transfer in antimonene, Angew. Chem.
[42] S. Golovynskyi, I. Irfan, M. Bosi, L. Seravalli, O.I. Datsenko, I. Golovynska, B. Li,
Int. Ed. 56 (2017) 14389−14394.
D. Lin, J. Qu, Exciton and trion in few-layer MoS2: thickness- and temperature -
[15] Y. Wang, G. Qiu, R. Wang, S. Huang, Q. Wang, Y. Liu, Y. Du, W. A. Goddard, M. J.
dependent photoluminescence, Appl. Surface Sci. (2020), https://doi.org/10.1016/
Kim, X. Xu, P.D. Ye, W. Wu, Field-effect transistors made from solution-grown two-
j.apsusc.2020.146033.
dimensional tellurene, Nat. Electron. 1 (2018) 228−236.
[43] G. Pradhan, A.K. Sharma, Temperature controlled 1T/2H phase ratio modulation in
[16] W. Hoerstel, D. Kusnick, M. Spitzer, High-field transport and low-field mobility in
mono- and few layered MoS2 films, Appl. Surface Sci. 479 (2019) 1236–1245.
tellurium single crystals, Phys. Stat. Sol. B 60 (1973) 213–221.
[44] J. Yan, Y. Zhang, P. Kim, A. Pinczuk, Electric field tuning of electron-phonon
[17] T. Doi, K. Nakao, H. Kamimura, The valence band structure of tellurium, J. Phys.
coupling in graphene, Phys. Rev. Lett. 98 (2007) 166802.
Soc. Jpn. 28 (1970) 36–43.
[45] M. Akhtar, G. Anderson, R. Zhao, A. Alruqi, J.E. Mroczkowska, G. Sumanasekera,
[18] H.O. Churchill, G.J. Salamo, S.Q. Yu, T. Hironaka, X. Hu, J. Stacy, I. Shih, Toward
J.B. Jasinski, Recent advances in synthesis, properties and applications of phos-
single atom chains with exfoliated tellurium, Nanoscale Res. Lett. 12 (2017) 488,
phorene, 2D Mater. Appl. 1 (5) (2017), https://doi.org/10.1038/s41699-017-
https://doi.org/10.1186/s11671-017-2255-x.
0007-5.
[19] Q. Wang, M. Safdar, K. Xu, M. Mirza, Z. Wang, J. He, Van der Waals epitaxy and
[46] W. Zhao, Z. Ghorannevis, K.K. Amara, J.R. Pang, M. Toh, X. Zhang, C. Kloc,
photoresponse of hexagonal tellurium nanoplates on flexible mica sheets, ACS Nano
P.H. Tan, G. Eda, Lattice dynamics in mono- and few-layer sheets of WS2 and WSe2,
8 (2014) 7497–7505.
Nanoscale 5 (2013) 9677–9683.
[20] M. Safdar, X. Zhan, M. Niu, M. Mirza, Q. Zhao, Z. Wang, J. Zhang, L. Sun, J. He, Site
[47] S. Yi, Z. Zhu, X. Cai, Y. Jia, J.H. Cho, The nature of bonding in tellurium composed
specific nucleation and controlled growth of a vertical tellurium nanowire array for
of one-dimensional helical chains, Inorg. Chem. 57 (2018) 5083–5088.
high performance field emitters, Nanotechnology 24 (2013) 185705.
[48] N. Hussain, M. Rafique, T. Anwar, M. Murtaza, J. Liu, F. Nosheen, K. Huang,
[21] C.J. Hawley, B.R. Beatty, G. Chen, J.E. Spanier, Shape-controlled vapor-transport
Y. Huang, J. Lang, H. Wu, A high-pressure mechanism for realizing sub-10 nm
growth of tellurium nanowires, Crystal Growth & Design 12 (2012) 2789–2793.
tellurium nanoflakes on arbitrary substrates, 2D Mater. 6 (2019) 045006.
[22] G. Tai, B. Zhou, W. Guo, Structural characterization and thermoelectric transport
properties of uniform single crystalline lead-telluride nanowires, J. Phys. Chem. C

11

You might also like