Applied Surface Science: Full Length Article

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Applied Surface Science 526 (2020) 146730

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

First-principles study of mechanical, electronic and optical properties of T


Janus structure in transition metal dichalcogenides

Vuong Van Thanha, , Nguyen Duy Vana, Do Van Truongb,c, Riichiro Saitod,

Nguyen Tuan Hungd,e,
a
Department of Design of Machinery and Robot, School of Mechanical Engineering, Hanoi University of Science and Technology, Hanoi, Viet Nam
b
Department of Mechatronics, School of Mechanical Engineering, Hanoi University of Science and Technology, Hanoi, Viet Nam
c
International Research Center for Computational Materials Science, Hanoi University of Science and Technology, Hanoi, Viet Nam
d
Department of Physics, Tohoku University, Sendai 980-8578, Japan
e
Frontier Research Institute for Interdisciplinary Sciences, Tohoku University, Sendai 980-8578, Japan

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

Keywords: Using first-principles calculations, we investigate mechanical, electronic and optical properties of so-called Janus
Janus transition metal dichalcogenides structure for monolayer transition metal dichalcogenides (TMDs), MXY (M = Mo, W; X or Y = S, Se, Te; X ≠ Y),
Ideal strength in which chalcogen atoms at both side of the TMDs are not the same elements. Our calculated results indicate
Density functional theory that WSSe shows the highest stiffness and the most ideal strength among the Janus TMDs due to their strongest
Optical absorption
ionic bond. In the unstrain cases, WSeTe, WSSe and MoSeTe are direct-gap semiconductors, while MoSSe, MoSTe
and WSTe are indirect-gap semiconductors. The energy band gaps of the Janus TMDs decrease with increasing of
the tensile strain due to the coupling between the p and d orbitals of the X/Y and M atoms, respectively.
Furthermore, the tensile strain effectively modulates the optical absorption of the Janus TMDs. For example, the
optical absorption of MoSSe is three times stronger at a photon energy of 2.5 eV. The calculated results of Janus
TMDs provide useful information for applications in nanoelecromechanical, optoelectronic, and photocatalyst
devices.

1. Introduction 10% for WS2, and 11% for WSe2 and WTe2 [17]. Theoretically, the
effect of strain on the optical properties of the 2D TMDs have been
Two-dimensional (2D) transition metal dichalcogenides (TMDs) widely investigated [17–20], in which real part of the dielectric con-
have been promising for nanoelectronic, optoelectronic and photo- stant, reflectance index, and refractive index of the WX2 are enhanced
catalytic devices due to their unique properties [1–6], such as an in- with increasing the strain [17]. These results show that strain en-
direct-to-direct band gap transition, [7–9], valley-selective circular di- gineering is an effective tool to tune the electronic and optical prop-
chroism [10] and multi-excitons [11]. Monolayer TMDs are direct-gap erties of the TMDs [15–20].
semiconductors in a visible range from 400 nm (3 eV) to 700 nm Recently, Janus MoSSe, which is defined by a sandwiched S-Mo-Se
(1.8 eV) [12]. Thus, the monolayer TMDs might be promising materials layer structure, has been successfully synthesized [21,22] using a che-
for optoelectronics applications, in which the energy band gap and the mical vapor deposition. Janus TMD is labeled by MXY (M = Mo, W; X
band structures are essential properties for discussing the performance. or Y = S, Se, Te; X ≠ Y). Compared with the conventional TMDs, the
In oder to control these properties, applying strain on the monolayer electronic properties of the Janus TMD have novel features due to the
TMDs has been widely investigated both theoretically and experimen- structural symmetry-breaking [23,24]. Since the top and bottom atomic
tally [13,14]. Yue et al. [15] shows that MoS2 has a direct-to-indirect layers of the Janus TMDs consist of different atoms, internal electric
transition at a strain ∼ 1% and a semiconductor-to-metal transition at a field induced by the dipole moment separate an electron–hole pair,
strain ∼ 10%. By using density functional theory (DFT), Scalise et al. which is helpful for application of water splitting [25]. Wang et al. [26]
[16] also theoretically showed that MoS2 shows a direct-to-indirect also showed that the 2D Janus TMDs are semiconductors with a energy
transition at a tensile strain up to ∼ 2% . As for WX2 TMDs (X = S, Se, band gap range from 1.36 to 1.97 eV and that they observe strong op-
and Te), the energy band gaps are closed when the tensile strain reaches tical absorption at the visible-light region (1.61 to 3.10 eV). The effect


Corresponding author at: Frontier Research Institute for Interdisciplinary Sciences, Tohoku University, Sendai 980-8578, Japan (N.T. Hung).
E-mail addresses: thanh.vuongvan@hust.edu.vn (V.V. Thanh), nguyen@flex.phys.tohoku.ac.jp (N.T. Hung).

https://doi.org/10.1016/j.apsusc.2020.146730
Received 17 February 2020; Received in revised form 7 May 2020; Accepted 17 May 2020
Available online 26 May 2020
0169-4332/ © 2020 Elsevier B.V. All rights reserved.
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

of strain on the electronic structures and transport properties of the vacuum space of 30 Å in the direction perpendicular to the monolayer
Janus TMDs has also been considered by the DFT calculations and (z direction) is set to avoid interactions between adjacent layers. To
Boltzmann equation [27], respectively, in which the transport proper- obtain optimized atomic configurations of the Janus TMDs, the atomic
ties are improved by applying strain. Therefore, the strain engineering positions and cell vectors are relaxed using the Broyden-Fletcher-
is a helpful tool to improve the electronic and optical properties not Goldfarb-Shanno (BFGS) minimization method [41–44]. The model is
only for the TMDs but also for the Janus TMDs. In the analysis, the considered as a optimized structure when the Hellmann–Feynman
mechanical properties such as the mechanical breakdown and ideal forces and all components of the stress are less than 0.0005 Ry/a.u. and
strength, which can be defined by a maximum value of the stress that 0.05 GPa, respectively.
the materials can achieve by applying the strain, are important para- The Themo-pw code [45] is used to calculate the elastic constants Cij
meters. Although many studies have focused on synthesis, electronic of the Janus TMDs. The calculated stresses for a 2D material are ex-
and optical properties of the 2D Janus materials [21–23,26–33], their pressed as a 2D force per length in the units of N/m by taking the
mechanical properties have not been clarified yet, in particular ideal product of the 3D Cauchy stresses (in the units of N/m2) and the super-
strength and mechanical breakdown. These properties would be im- cell thickness of 30 Å [46]. The Young modulus, Y, and the Poisson
portant for synthesizing the Janus material and its application to the ratio, ν , of the 2D material are expressed by [47–49]
device. Thus, the theoretical analysis of mechanical property of the 2D 2 2
C11 − C12 C12
Janus TMDs is necessary. Further, understanding the electronic and Y= , and ν = .
optical properties of the Janus TMDs under the tensile strain is also C11 C11 (1)
necessary to investigate for nanoelectronic, optoelectronic, and photo- σi ∊i
To investigate the ideal strength and ideal strain (i.e. the strain
catalytic applications. at the maximum stress) of the Janus TMDs under the strains, a tensile
In the present paper, we investigate the mechanical, electronic and strain is applied to the structures by elongating the cell in uniaxial di-
optical properties of the Janus TMDs MXY as a function of the tensile rection ( x , y ) with an increment of 0.02 of the tensile strain until the
strain using first-principles calculations. We show that the Janus TMDs structure is broken. Here, the dimensionless tensile strain is defined by
show anisotropic behavior for uniaxial tensile strain. We also discuss ε = (L − L0)/ L0 , where L and L0 are the length of the unit cell for the
the effect of the strain on the electronic and optical properties of the strained and unstrained structures, respectively.
Janus TMDs. To obtain the optical properties of the Janus TMDs under the tensile
strain, we calculated the optical absorption coefficient α (ω) using the
2. Computational methods following equation [50]
1/2
In this study, we perform the first-principles calculations based on 4πe ⎧ [ε12 + ε22 ]1/2 − ε1 ⎫
α (ω) = ,
the DFT using Quantum ESPRESSO [34]. A cut-off energy for the wave hc ⎨
⎩ 2 ⎬
⎭ (2)
function is taken to be 60 Ry. k -point grid in the Brillouin zone is set by
where ε1 (ω) and ε2 (ω) are real and imaginary parts of the dielectric
a 16 × 16 × 1 grid for all Janus TMDs in the Monkhorst–Pack method
function, respectively, c is the speed of light in the vacuum and ω is the
[35]. Both the local-density approximations (LDA) [36] and the gen-
angular frequency of light.
eralized-gradient approximation (GGA) [37] are performed for dis-
cussing mechanical properties and electronic (or optical) properties,
respectively. The energy band gap is corrected by the Heyd-Scuseria- 3. Results and discussion
Ernzerhof (HSE) hybrid functional [38]. It is because the results of the
GGA generally underestimate the mechanical properties compared with 3.1. Structure and stability of Janus TMDs
those by the LDA. In particular, Cooper et al. [39] discussed this be-
havior for the monolayer MoS2. On the other hand, the GGA + HSE is In Table 1, we list the optimized lattice parameters of the Janus
better than the LDA to evaluate the whole band structure of the 2D TMDs MXY by using the LDA, in which the lattice parameters of MoXY
Janus TMDs [40]. and WXY have quite similar values to each other. The lattice constants
The atomic structures of the Janus TMDs are shown in Fig. 1. The of the Janus TMDs are arranged as follows MSSe < MSTe < MSeTe
periodic boundary condition is applied in all the models, in which a (M = Mo, W), respectively. The lattice constant of MoSSe is 3.23 Å,
which is in a good agreement with the previous theoretical results
[21,30,51,52], indicating that the present calculations are consistent.
The lattice constant of MoSSe lies between that of MoS2 (a = 3.19 Å)
and MoSe2 (a = 3.31 Å) [48]. We note that the buckling height (h, see
Fig. 1(b)) of MoSSe (3.22 Å) is larger than that of MoS2 (3.12 Å [51],
3.16 Å [53]) and smaller than that of MoSe2 (3.35 Å [51],3.34 Å [53]).
Although the atomic structures of the Janus TMDs have been opti-
mized, they do not guarantee the elastic and dynamic stabilities [54].
Therefore, we need to analyze the phonon dispersion and the elastic
stability conditions for the Janus TMDs. In Fig. 2, we show the phonon
dispersions of all Janus TMDs. It can be seen from the insets of Fig. 2
that WSeTe, MoSeTe, MoSSe, MoSTe are dynamically stable without
any imaginary phonon frequency, while WSSe and WSTe show very
small negative (imaginary) frequencies (∼ −1 cm−1) around the
Γ -point, which may indicate a structural instability. Cheng et al. [53]
pointed out that these negative frequencies are related to a bending of
the X/Y plane in the Janus structures, and they showed MoSTe and
Fig. 1. The top view (a) and front view (b) of the Janus TMDs. The x and y MoSeTe are unstable with negative frequencies ∼ −1 cm−1. Since the
directions correspond to the armchair and zigzag directions, respectively. (c) difference in the method between Cheng et al. and us are GGA and LDA,
The first Brillouin Zone of the Janus TMDs. Under tensile strain, the original respectively, a small value of the imaginary phonon frequency is sen-
crystal symmetry is broken, thus M′ and K′ become no longer equivalent to M sitive. Nevertheless, since the Janus TMDs are synthesized on the sub-
and K, respectively. strate [21], the instability for bending can be avoided experimentally.

2
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

Table 1
i
Lattice constants a (Å), buckling heights h, elastic constants Cij (N/m), the Young modulus Y (N/m), the Poisson ratios ν , 2D stress σxx (N/m) (∊ixx ) and σ yy
i
(N/m) (∊iyy )
of the Janus TMDs.
Materials a h C11 C12 C66 Y ν i
σxx (∊ixx ) i
σ yy (∊iyy )

MoSSe 3.23 3.22 126.78 20.61 53.07 123.43 0.163 15.75 (0.27) 9.06 (0.17)
MoSTe 3.32 3.33 119.82 25.32 47.25 114.47 0.211 13.13 (0.26) 7.32 (0.15)
MoSeTe 3.36 3.47 96.57 20.67 37.95 92.15 0.214 13.07 (0.28) 6.57 (0.15)

WSSe 3.22 3.31 141.27 22.41 59.43 137.72 0.159 17.03 (0.27) 10.49 (0.17)
WSTe 3.32 3.32 133.62 26.25 53.70 128.46 0.196 14.73 (0.27) 8.68 (0.17)
WSeTe 3.38 3.45 111.03 17.31 46.86 108.33 0.156 14.22 (0.30) 7.77 (0.17)

Moreover, for MoSSe, the calculated out-of-plane A1 and in-plane E′ armchair and zigzag directions, respectively. Li et al. [56] showed that
modes are 285 cm−1 and 356 cm−1, which agree well with the ob- the anisotropic behavior in MoS2 comes from the fact that the dis-
served Raman peaks at 288 cm−1 and 355 cm−1 [21], respectively. placement of the S atoms makes the zigzag Mo-S chain more flat than
Compared with MoS2 (A1 = 406 cm−1 and E′ = 387 cm−1), the shift of that in the homogeneous structure under zigzag strain, while there is a
the A1 mode is larger than that of the E′ mode. It is because that the A1 small displacement under the armchair strain. When we discuss the
optical mode might be related to the out-of-plane symmetry-breaking, strain effect, we should consider structural phase transition for a given
while the E′ mode is in-plane oscillation of the bond length [21,25]. stress. In order to check the stability, we optimize structures of MoSeTe
Next we discuss the elastic stability of the Janus TMDs. The neces- with a supercell 2a × 2 3 a at ∊xx = 0.28 and ∊yy = 0.15 for armchair and
sary and sufficient conditions of the elastic stability of the 2D materials zigzag directions, respectively. As shown in Fig. S1 (Supplementary
are C11 > |C12 | > 0 and C66 > 0 [55]. As shown in Table 1, we get that data), the calculated results indicate no structural transition for the
C11 > |C12 | > 0 and C66 > 0 for all Janus TMDs. Therefore, the Janus both cases, which is consistent with the previous works for monolayer
TMDs have the stable structures judging from these conditions. By using MoS2 [59,39].
Cij , the Young modulus Y and the Poisson ratio ν of the Janus TMDs are Since the ideal strain provides a large resistance to the tensile load,
calculated by Eq. (1) and are also listed in Table 1. Y of the Janus TMDs the ideal strength of the Janus TMDs along the armchair direction is
decreases with increasing atomic number of metal ZM (ZMo < ZW ), that larger than that along the zigzag direction, as shown in Figs. 3(a) and
is, Y(MSSe)> Y(MSTe)> Y(MSeTe) (M = Mo or W). Therefore, MoSSe (b). In the armchair direction, WSSe is the strongest in the Janus TMDs,
i
and WSSe are the stiffest materials in the Janus MoXY and WXY com- with σxx = 17.03 N/m. Bertolazzi et al. [58] reported the average value
pounds with Y = 123.43 GPa and Y = 137.72 GPa, respectively. Li of the ideal strength about 15 ± 3 N/m of MoS2 by using nano-in-
et al. [31] also reported a high value of Y(MoSSe) about 106 − 113 GPa. dentation in an atomic force microscope. In this present study, the ideal
i i
strength of MoSSe reaches σxx = 15.75 N/m (or σxx = 25.6 GPa with a
3.2. Ideal strength and failure behaviors of Janus TMDs layer thickness d 0 = 6.145Å [48,59]) at the ideal strain ∊ixx = 0.27 ,
which is in a good agreement with the experimental observations (22±
In Figs. 3(a) and (b), we plot the stress-strain curves of the Janus 4 GPa) [58]. The ideal strength and ideal strain are obtained from the
TMDs in the x and y directions, respectively. We can see that the Janus stress–strain relationship in Figs. 3(a) and (b), and are listed in Table 1.
TMDs show an anisoptropic behavior for σ , that is, σxx ≠ σyy when In order to understand why WSSe has a highest ideal strength, we plot
∊xx = ∊yy , where x and y are the armchair and zigzag directions, re- the ideal strength as a function of the ionicity I as shown in Fig. 3(c).
spectively. The ideal strain in the armchair direction becomes up to The I of the Janus TMDs MXY is calculated by the Pauling expression as
0.26 − 0.30 , which is larger than that of the zigzag direction about
0.15 − 0.17 as shown in Table 1. This anisotropic behavior is also ob- 1
I = 1 − exp ⎡− (χM − χXY )2⎤,
served in MoS2 [56,57] with the ideal strains are 0.28 and 0.18 for the ⎣ 4 ⎦ (3)

Fig. 2. Phonon dispersions of the Janus TMDs by the LDA calculation at the equilibrium state. The insets show a magnified view of the acoustic branches around the
Γ - point.

3
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

Fig. 3. Engineering stress of the Janus TMDs by the LDA calculation as a function of tensile strain: σxx and σyy along (a) armchair and (b) zigzag directions,
respectively.

where χM is the electronegativities of the M atoms and χXY is the Se bond lengths, respectively. For the tensile strain, the M-S, M-Se and
average electronegativities of the X and Y atoms (i.e. M-Te bond lengths increase with increasing strain in the armchair di-
χXY = (χX + χY )/2 ) [60,61]. A large value of the I indicates more ionic rection, while they decrease in the zigzag direction. Since the change of
character in the Janus TMD [62]. As shown in Fig. 3(c), the relationship the bond lengths along the armchair direction is larger than that along
between the ideal strength and the I shows as a linear function for each zigzag direction, the ideal strain along the armchair direction is larger
i i
direction, in which are σxx = 0.51I + 10.4 and σyy = 0.47I + 4.4 for than that along the zigzag directions as shown in Figs. 3(a) and (b).
armchair and zigzag directions, respectively. Based on the relationship
I ∝ σ i , the Pauling ionicity I can be used as an important parameter to
evaluate the stiffness and ideal strength of the ionic materials. Since 3.3. Strain effect on electronic properties of Janus TMDs
WSSe has a highest I values about 13, it exhibits the highest ideal
strength. It is noted here that I of MoS2 is 12 [61]. In Fig. 5, we show the band structures of the Janus TMDs for
In order to understand the deformation mechanism of the Janus ∊xx = ∊yy = 0 by using the GGA + HSE calculation. The obtained results
TMDs, the bond lengths between M and X/Y atoms as a function of show that WSeTe, WSSe and MoSeTe are direct-gap semiconductors
tensile strains in both armchair (solid symbol) and zigzag (open with the bottom of conduction band (CBM) and the top of valence band
symbol) directions are plotted in Fig. 4. Without strain (∊xx = ∊yy = 0 ), (VBM) at the K point in the Brillouin zone, while MoSSe, MoSTe and
the M (Mo, W)-Te bond lengths of MXY are larger than the M-S and M- WSTe are indirect-gap semiconductors with an energy band gap from
1.42 eV (MoSTe) to 2.3 eV (WSSe). In Fig. 5, the CBM and VBM are

Fig. 4. The M (Mo, W)- S, M-Se and M-Te bond lengths of the Janus TMDs as a function of strain for the x (armchair) and y (zigzag) directions, respectively.

4
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

Fig. 5. Band structures of the Janus TMDs by the GGA + HSE calculation without strain (∊xx = ∊yy = 0 ). The Fermi level is set to be zero and the arrow connects the
top of valence band and the bottom of conduction band for each Janus TMD.

connected by arrow for each Janus TMD. The energy band gap of significantly modified by the strain, which is in a good agreement with
MoSSe is 2.08 eV, which is in a good agreement with the previous the previous studies [27]. For ∊ = 0 , MoSeTe is a direct-gap semi-
theoretical value of 2.1 eV [25]. Since the energy difference between conductor with the band gap of 1.75 eV, which is consistent with the
the two valleys of the conduction band at the K point and along K-Γ line previous theoretical study (1.71 eV) using the HSE [25] method. When
is very small, e.g. 0.1 eV for MoSSe, the direct- or indirect-gap is sen- the strain increases gradually, the energy band gap at the K-point
sitive to the approximations of exchange–correlation energy or spi- monotonically decreases. At ∊ = 0.04 , MoSeTe becomes an indirect-gap
n–orbit coupling (SOC). For example, by using GGA + SOC, Liu et al. semiconductor with band gap of ∼ 1.45 eV for both armchair and zigzag
[63] showed that MoSSe and WSSe are direct-gap semiconductors, directions. These results suggest that not only band gap but also di-
whereas other Janus TMDs are indirect-gap semiconductors. On the rect–indirect transition of MoSeTe occur by applying strain. The band
other hand, Wang et al. [26] used the HSE functional and SOC to cal- structures of the other Janus TMDs (MoSSe, MoSTe, WSeTe, WSSe and
culate the electronic structures, and they showed that all Janus TMDs WSTe) are also calculated by using the GGA + HSE calculation, as
except MoSSe are indirect-gap semiconductors. shown in Figs. S3 to S7 in Supplementary data.
In Figs. 6 (a) and (b), we show the band structures of MoSeTe at the In Fig. 7, the energy band gaps of the Janus TMDs are plotted as
different tensile strains in the armchair and zigzag directions, respec- functions of the strain for (a) the armchair and (b) zigzag directions.
tively. As we see from Fig. 6, the band structure of MoSeTe is Since the Janus structures are unstable at a critical strain, we only

Fig. 6. Band structures of the monolayer MoSeTe by the GGA + HSE calculation for the tensile strains ∊ = 0 , 0.04, 0.08 and 0.12 for (a) armchair and (b) zigzag
directions, respectively. The Fermi level is set to be zero and the arrow connects the top of valence band and the bottom of conduction band for each strain.

5
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

Janus TMDs might be as a good candidate materials for the photo-


voltaic solar cells. For the out-of-plane, the Janus TMDs has the largest
value of 2.5 × 10 4 cm−1 at 3.1 eV, as shown in Fig. 8(b). In order to
understand the peaks of the absorption spectrum, we calculate the joint
density of states (JDOS) as shown in Fig. 8(c). Here, the JDOS is defined
as [67]
1
JDOS(E ) =
4π 3
∫ d3kδ [Ev (k ) − Ec (k ) − E ], (4)
where k is wave-vector, and Ev (k ) and Ec (k ) are the valence and con-
duction bands, respectively. The JDOS is associated with the transitions
from the valence bands to the conduction bands. Therefore, a large
peak in the JDOS corresponds to the peak in the absorption spectrum as
shown in Figs. 8. From Eq. (4), a peak is found at a photon of energy
Fig. 7. Energy band gaps of the Janus TMDs by the HSE calculation as a
E = Ek, v − Ek, c , which corresponds to the two-dimensional Van Hove
function of tensile strain in (a) the armchair and (b) zigzag directions, respec-
singularities in the Brillouin zone. The Van Hove singularities usually
tively.
occur at high-symmetry points in the k -space [68]. As discussed in
Section 3.3, however, the electron states at the high-symmetry points
consider the strain up to the ideal strain for each Janus TMD (refer such as K and Γ can be shifted by applying strain. Therefore, the strain
Table S1 for the GGA case). We note that the ideal strains obtained by engineering can work as an effective tool to improve the optical prop-
using the GGA and LDA show a difference of about 10%, as shown in erties of the Janus TMDs.
Table S1 and Fig. S3. Chang et al. [40] showed that the electronic states In Figs. 9(a) and (b), we show the optical absorption of MoSSe for
at the top of the valence band and the bottom of the conduction band several strains along the armchair and zigzag directions, respectively.
are given by couplings between p orbitals of X/Y atoms and d z 2, d x 2 − y 2 , For tensile strains along the armchair direction, α// (ω) xx is much en-
and d xy orbitals of M atoms. Therefore, the change of the X-M-Y bond hanced at around 2.5 eV, while α// (ω) yy and α⊥ (ω) zz do not show a
angle under the strain will modify the couplings of orbitals. For a tensile significant change. As shown in Fig. 9(a), at the photon energy of
strain, the X/Y atoms move toward the M atoms and the X-M-Y bond 2.5 eV, α// (ω) xx increases from 7.5 × 10 4 cm−1 to 28 × 10 4 cm−−1. It is
angle is reduced, which lead to weaken the coupling between the p and because that at ∊xx = 0.12, the band gap of MoSSe is reduced by 1.0 eV.
out-of-plane d z 2 orbitals and to strengthen the coupling between the p Thus, the peak position of α// (ω) xx at 3.5 eV is shifted to 2.5 eV. The
and in-plane d x 2 − y 2 and d xy orbitals. As a result the electronic energies at scenario is similar to the case of tensile strains along the zigzag direc-
the Γ and K points increase and decrease by the tensile strain, respec- tion. That is, the peak position of α// (ω) yy at 3.5 eV is shifted to 2.5 eV.
tively. Thus the energy band gap of the Janus TMDs decreases with Therefore, the optical absorption is enhanced around 2.5 eV, in which
increasing the tensile strain. In addition, the direct-to-indirect transi- α// (ω) yy increases from 8 × 10 4 cm−1 to 31 × 10 4 cm−1 as shown in
tions also occur for MoSeTe at ∊ = 0.04 , as shown in Fig. 5. Fig. 9(b). The calculated strain-dependent α (ω) of other Janus TMDs,
including MoSeTe, MoSTe, WSeTe, WSSe and WSTe, are shown in Figs.
3.4. Strain effect on optical properties of Janus TMDs S8 to S12 in Supplementary data.

In Figs. 8(a) and (b), we show the optical absorption α (ω) of the 4. Conclusions
Janus TMDs by using the GGA + HSE calculation at the unstrain
(∊xx = ∊yy = 0 ) in the energy range from ℏω = 0 to 5 eV, in which α (ω) In summary, we investigate the mechanical, electronic and optical
is calculated by Eq. (2). We find that α (ω) exhibits a strong anisotropy properties of the Janus TMDs under the strain using the LDA and
along the polarization directions (α// (ω) xx = α// (ω) yy > α⊥ (ω) zz ). In the GGA + HSE calculations. The obtained results show that WSSe is the
visible-light region, α (ω) of the Janus TMDs along the in-plane strongest material in the Janus MXY compounds. The calculated
(α// (ω) xx = α// (ω) yy ) direction is much larger than that of the out-of- strengths of WSSe are 17.03 N/m and 10.49 N/m for the armchair and
plane (α⊥ (ω) zz ) direction with one order of magnitude, which is con- zigzag directions, respectively. Based on the relationship I ∝ σ i , we find
sistent with the previous DFT calculations using the HSE + SOC method that the Pauling ionicity I can be used as an important parameter to
[26]. For the in-plane direction, the Janus TMDs exhibit strong ab- evaluate the stiffness and the ideal strength of the Janus materials. At
sorption when the photon energy is more than 2 eV, as shown in the unstrain condition, the energy band gaps of the Janus TMDs from
Fig. 8(a). The optical absorption of all the Janus TMDs reaches around 1.42 eV to 2.3 eV, in which WSeTe, WSSe and MoSeTe are direct-gap
15 × 10 4 − 30 × 10 4 cm−1 at the visible-light region, which is in a good semiconductors, while MoSSe, MoSTe and WSTe are indirect-gap
agreement with the earlier DFT results [25], and the values are com- semiconductors. When the tensile strain is applied, the band gap of the
parable to CdTe and CIGS materials [65]. Since the photocurrent J (ω) is Janus TMDs decreases with increasing the strain due to the coupling
proportional to the optical absorption coefficient, J (ω) ∝ α (ω) , [66] the between the p and d orbitals of the X/Y and M atoms, respectively.

Fig. 8. Optical absorption of six Janus MXY monolayers by GGA + HSE calculation at the equilibrium condition (∊xx = ∊yy = 0 ) along the (a) in-plane (α// (ω) ), (b)
out-of-plane directions (α⊥ (ω) ) and (c) JDOS of Janus TMDs. The dash lines denote the range of visible light from 1.61 eV to 3.10 eV.

6
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

Fig. 9. Optical absorption of the monolayer MoSSe as a function of photo energy for several tensile strains (∊ = 0 , 0.04, 0.08 and 0.12) along (a) armchair and (b)
zigzag directions, respectively. The dash lines denote the range of visible light from 1.61 eV to 3.10 eV.

Furthermore, the optical absorption coefficient of the Janus TMDs is [7] K.F. Mak, C. Lee, J. Hone, J. Shan, T.F. Heinz, Atomically thin MoS2: a new direct-
enhanced by applying the tensile strain. The use of various Janus TMDs gap semiconductor, Phys. Rev. Lett. 105 (13) (2010) 136805 .
[8] I.G. Lezama, A. Arora, A. Ubaldini, C. Barreteau, E. Giannini, M. Potemski,
under applying strain would allow to create new flexible optoelectronic A.F. Morpurgo, Indirect-to-direct band gap crossover in few-layer MoTe2, Nano
devices. Lett. 15 (4) (2015) 2336–2342.
[9] Y. Sun, D. Wang, Z. Shuai, Indirect-to-direct band gap crossover in few-layer
transition metal dichalcogenides: a theoretical prediction, J. Phys. Chem. C. 120
Declaration of Competing Interest (38) (2016) 21866–21870.
[10] T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, et al.,
Valley-selective circular dichroism of monolayer molybdenum disulphide, Nat.
The authors declare that they have no known competing financial
Commun. 3 (2012) 887.
interests or personal relationships that could have appeared to influ- [11] K.F. Mak, K. He, C. Lee, G.H. Lee, J. Hone, T.F. Heinz, J. Shan, Tightly bound trions
ence the work reported in this paper. in monolayer MoS2, Nature Mater. 12 (3) (2013) 207.
[12] S. Lebegue, O. Eriksson, Electronic structure of two-dimensional crystals from ab
initio theory, Phys. Rev. B. 79 (11) (2009) 115409 .
Acknowledgements [13] S.-D. Guo, Y. Wang, Small compressive strain-induced semiconductor-metal tran-
sition and tensile strain-enhanced thermoelectric properties in monolayer PtTe2,
V.V.T. and D.V.T acknowledge the Vietnam’s National Foundation Semicond. Sci. Tech. 32 (5) (2017) 055004 .
[14] X. Chen, G. Wang, Tuning the hydrogen evolution activity of MS2 (M= Mo or Nb)
for Science and Technology Development (NAFOSTED) with No. monolayers by strain engineering, Phys. Chem. Chem. Phys. 18 (14) (2016)
107.02–2016.18. R.S. acknowledges JSPS KAKENHI (No. 9388–9395.
JP18H01810). N.T.H. acknowledges JSPS KAKENHI (No. JP20K15178) [15] Q. Yue, J. Kang, Z. Shao, X. Zhang, S. Chang, G. Wang, S. Qin, J. Li, Mechanical and
electronic properties of monolayer MoS2 under elastic strain, Phys. Lett. A. 376
and the financial support from the Frontier Research Institute for (12–13) (2012) 1166–1170.
Interdisciplinary Sciences, Tohoku University. [16] E. Scalise, M. Houssa, G. Pourtois, V. Afanas’ev, A. Stesmans, Strain-induced
semiconductor to metal transition in the two-dimensional honeycomb structure of
MoS2, Nano Res. 5 (1) (2012) 43–48.
Appendix A. Supplementary data [17] J. Liu, H. Liu, J. Wang, H. Sheng, G. Tang, J. Zhang, D. Bai, Optical and electronic
properties of dichalcogenides WX2 (X= S, Se, and Te) monolayers under biaxial
Supplementary data associated with this article can be found, in the strain, Physica B: Condensed Matter. 568 (2019) 18–24.
[18] M. Feierabend, A. Morlet, G. Berghäuser, E. Malic, Impact of strain on the optical
online version, athttps://doi.org/10.1016/j.apsusc.2020.146730. fingerprint of monolayer transition-metal dichalcogenides, Phys. Rev. B. 96 (4)
(2017) 045425 .
References [19] J. Lee, J. Huang, B. G. Sumpter, M. Yoon, Strain-engineered optoelectronic prop-
erties of 2D transition metal dichalcogenide lateral heterostructures, 2D Mater. 4
(2), 2017, 021016.
[1] K. Kam, B. Parkinson, Detailed photocurrent spectroscopy of the semiconducting [20] M. Farkous, M. Bikerouin, D.V. Thuan, Y. Benhouria, M. El-Yadri, E. Feddi,
group vib transition metal dichalcogenides, J. Phys. Chem. 86 (4) (1982) 463–467. H. Erguig, F. Dujardin, C.V. Nguyen, N.V. Hieu, et al., Strain effects on the elec-
[2] R. Tenne, A. Wold, Passivation of recombination centers in n-WSe2 yields high ef- tronic and optical properties of van der Waals Heterostructure MoS2/WS2: A first-
ficiency (> 14%) photoelectrochemical cell, Appl. Phys. Lett. 47 (7) (1985) principles study, Physica E: Low-Dimens. Syst. Nanostruct. 116 (2020) 113799 .
707–709. [21] A.-Y. Lu, H. Zhu, J. Xiao, C.-P. Chuu, Y. Han, M.-H. Chiu, C.-C. Cheng, C.-W. Yang,
[3] E. Gourmelon, O. Lignier, H. Hadouda, G. Couturier, J. Bernede, J. Tedd, J. Pouzet, K.-H. Wei, Y. Yang, et al., Janus monolayers of transition metal dichalcogenides,
J. Salardenne, MS2 (M= W, Mo) photosensitive thin films for solar cells, Sol. Energ. Nat. Nanotechnol 12 (8) (2017) 744.
Mat. Sol. C. 46 (2) (1997) 115–121. [22] J. Zhang, S. Jia, I. Kholmanov, L. Dong, D. Er, W. Chen, H. Guo, Z. Jin, V.B. Shenoy,
[4] Y. Zheng, Y. Huang, H. Shu, X. Zhou, J. Ding, X. Chen, W. Lu, The effect of lithium L. Shi, et al., Janus monolayer transition-metal dichalcogenides, ACS Nano 11 (8)
adsorption on the formation of 1T- MoS2 phase based on first-principles calculation, (2017) 8192–8198.
Phys. Lett. A 380 (20) (2016) 1767–1771. [23] F. Li, W. Wei, P. Zhao, B. Huang, Y. Dai, Electronic and optical properties of pristine
[5] J. Du, P. Song, L. Fang, T. Wang, Z. Wei, J. Li, C. Xia, Elastic, electronic and optical and vertical and lateral heterostructures of Janus MoSSe and WSSe, J. Phys. Chem.
properties of the two-dimensional PtX2 (X= S, Se, and Te) monolayer, Appl. Surf. Lett. 8 (23) (2017) 5959–5965.
Sci. 435 (2018) 476–482. [24] W. Zhou, J. Chen, Z. Yang, J. Liu, F. Ouyang, Geometry and electronic structure of
[6] Z. Cui, K. Ren, Y. Zhao, X. Wang, H. Shu, J. Yu, W. Tang, M. Sun, Electronic and monolayer, bilayer, and multilayer Janus WSSe, Phys. Rev. B 99 (2019) 075160 .
optical properties of van der Waals heterostructures of g-GaN and transition metal [25] C. Xia, W. Xiong, J. Du, T. Wang, Y. Peng, J. Li, Universality of electronic char-
dichalcogenides, Appl. Surf. Sci. 492 (2019) 513–519. acteristics and photocatalyst applications in the two-dimensional Janus transition

7
V.V. Thanh, et al. Applied Surface Science 526 (2020) 146730

metal dichalcogenides, Phys. Rev. B 98 (2018) 165424 . [45] A. Dal Corso, Elastic constants of beryllium: a first-principles investigation, J. Phys:
[26] J. Wang, H. Shu, T. Zhao, P. Liang, N. Wang, D. Cao, X. Chen, Intriguing electronic Condens. Matter 28 (7) (2016) 075401 .
and optical properties of two-dimensional Janus transition metal dichalcogenides, [46] Q. Peng, W. Ji, S. De, Mechanical properties of the hexagonal boron nitride
Phys. Chem. Chem. Phys. 20 (27) (2018) 18571–18578. monolayer: Ab initio study, Comput. Mater. Sci. 56 (2012) 11–17.
[27] S.-D. Guo, J. Dong, Biaxial strain tuned electronic structures and power factor in [47] N.T. Hung, A.R. Nugraha, R. Saito, Two-dimensional MoS2 electromechanical ac-
Janus transition metal dichalchogenide monolayers, Semicond. Sci. Tech. 33 (8) tuators, J. Phys. D: Appl. Phys. 51 (7) (2018) 075306 .
(2018) 085003 . [48] V. Van Thanh, N.T. Hung, et al., Charge-induced electromechanical actuation of
[28] L. Dong, J. Lou, V.B. Shenoy, Large in-plane and vertical piezoelectricity in Janus Mo-and W-dichalcogenide monolayers, RSC adv. 8 (67) (2018) 38667–38672.
transition metal dichalcogenides, ACS Nano 11 (8) (2017) 8242–8248. [49] V. Van Thanh, N.T. Hung, et al., Charge-induced electromechanical actuation of
[29] Y. Ji, M. Yang, H. Lin, T. Hou, L. Wang, Y. Li, S.-T. Lee, Janus structures of transition two-dimensional hexagonal and pentagonal materials, Phys. Chem. Chem. Phys. 21
metal dichalcogenides as the heterojunction photocatalysts for water splitting, J. (40) (2019) 22377–22384.
Phys. Chem. C 122 (5) (2018) 3123–3129. [50] X. Li, J. Zhao, J. Yang, Semihydrogenated bn sheet: a promising visible-light driven
[30] K.D. Pham, N.N. Hieu, H.V. Phuc, B.D. Hoi, V.V. Ilysov, B. Amin, C.V. Nguyen, First photocatalyst for water splitting, Sci. Rep. 3 (2013) 1858.
principles study of the electronic properties and schottky barrier in vertically [51] S.-D. Guo, Phonon transport in janus monolayer mosse: a first-principles study,
stacked graphene on the Janus MoSeS under electric field, Comput. Mater. Sci. 153 Phys. Chem. Chem. Phys. 20 (10) (2018) 7236–7242.
(2018) 438–444. [52] S. Deng, L. Li, O.J. Guy, Y. Zhang, Enhanced thermoelectric performance of
[31] X. Li, X. Wang, W. Hao, C. Mi, H. Zhou, Structural, electronic, and electro- monolayer MoSSe, bilayer MoSSe and graphene/MoSSe heterogeneous nanor-
mechanical properties of MoSSe/blue phosphorene heterobilayer, AIP Adv. 9 (11) ibbons, Phys. Chem. Chem. Phys. 21 (33) (2019) 18161–18169.
(2019) 115302 . [53] Y. Cheng, Z. Zhu, M. Tahir, U. Schwingenschlögl, Spin-orbit-induced spin splittings
[32] X. Zhao, B. Qiu, G. Hu, W. Yue, J. Ren, X. Yuan, Transition-metal doping/adsorption in polar transition metal dichalcogenide monolayers, EPL (Europhysics Letters) 102
induced valley polarization in Janus WSSe: First-principles calculations, Appl. Surf. (5), 2013, 57001.
Sci. 490 (2019) 172–177. [54] R.A. Cowley, Acoustic phonon instabilities and structural phase transitions, Phys.
[33] L. Yu, S. Sun, X. Ye, Electronic and magnetic properties of janus MoSSe/WSSe su- Rev. B 13 (1976) 4877.
perlattice nanoribbon: A first-principles study, Phys. Chem. Chem. Phys. 22 (2020) [55] F. Mouhat, F. Coudert, Necessary and sufficient elastic stability conditions in var-
2498–2508. ious crystal systems, Phys. Rev. B 90 (2014) 224104 .
[34] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, [56] T. Li, Reply to ”Comment on “ideal strength and phonon instability in single-layer
G.L. Chiarotti, M. Cococcioni, I. Dabo, et al., Quantum espresso: a modular and MoS2”’, Phys. Rev. B 90 (2014) 167402 .
open-source software project for quantum simulations of materials, J. Phys. [57] R.C. Cooper, J.W. Kysar, C.A. Marianetti, Comment on ”ideal strength and phonon
Condens. Matter. 21 (39) (2009) 395502 . instability in single-layer MoS2”, Phys. Rev. B 90 (2014) 167401 .
[35] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. [58] S. Bertolazzi, J. Brivio, A. Kis, Stretching and breaking of ultrathin MoS2, ACS nano
B 13 (12) (1976) 5188. 5 (12) (2011) 9703–9709.
[36] C. Hartwigsen, S. Gœdecker, J. Hutter, Relativistic separable dual-space Gaussian [59] T. Li, Ideal strength and phonon instability in single-layer MoS2, Phys. Rev. B. 85
pseudopotentials from H to Rn, Phys. Rev. B 58 (1998) 3641. (23) (2012) 235407 .
[37] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made [60] L. Pauling, The Nature of the Chemical Bond vol. 260, Cornell University Press
simple, Phys. Rev. Lett. 77 (18) (1996) 3865. Ithaca, NY, 1960.
[38] J. Heyd, G.E. Scuseria, M. Ernzerhof, Hybrid functionals based on a screened [61] S.P. Hind, P.M. Lee, Kkr calculations of the energy bands in NbSe2, MoS2 and alpha
Coulomb potential, J. Chem. Phys. 118 (18) (2003) 8207. MoTe2, J. Phys. C: Solid State Phys. 13 (1980) 349.
[39] R.C. Cooper, C. Lee, C.A. Marianetti, X. Wei, J. Hone, J.W. Kysar, Nonlinear elastic [62] N. T. Hung, L. Yin, P. D. Tran, R. Saito, Simultaneous anionic and cationic redox in
behavior of two-dimensional molybdenum disulfide, Phys. Rev. B 87 (3) (2013) Mo3S11 polymer electrode of sodium-ion battery, J. Phys. Chem. C. 123 (51), 2019,
035423 . 30856-30862.
[40] C.-H. Chang, X. Fan, S.-H. Lin, J.-L. Kuo, Orbital analysis of electronic structure and [63] H. Liu, Z. Huang, C. He, Y. Wu, L. Xue, C. Tang, X. Qi, J. Zhong, Strain engineering
phonon dispersion in mos 2, MoSe2, WS2, and WSe2 monolayers under strain, Phys. the structures and electronic properties of Janus monolayer transition-metal di-
Rev. B 88 (2013) 195420 . chalcogenides, J. Appl. Phys. 125 (2019) 082516 .
[41] C.G. Broyden, The convergence of a class of double-rank minimization algorithms: [65] M.R. Kim, D. Ma, Quantum-dot-based solar cells: recent advances, strategies, and
2. the new algorithm, IMA J. Appl. Math. 6 (3) (1970) 222–231. challenges, J. Phys. Chem. Lett. 6 (1) (2014) 85–99.
[42] R. Fletcher, A new approach to variable metric algorithms, Comput. J. 13 (3) (1970) [66] M. Bernardi, M. Palummo, J.C. Grossman, Extraordinary sunlight absorption and
317–322. one nanometer thick photovoltaics using two-dimensional monolayer materials,
[43] D. Goldfarb, A family of variable-metric methods derived by variational means, Nano Lett. 13 (2013) 3664–3670.
Math. Comput. 24 (109) (1970) 23–26. [67] C. Kittel, Introduction to solid state physics, Wiley, New York, 1976.
[44] D.F. Shanno, Conditioning of quasi-newton methods for function minimization, [68] D. Brust, J.C. Phillips, F. Bassani, Critical points and ultraviolet reflectivity of
Math. Comput. 24 (111) (1970) 647–656. semiconductors, Phys. Rev. Lett. 9 (1962) 94.

You might also like