Você está na página 1de 7

Journal of Physics and Chemistry of Solids 135 (2019) 109081

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


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

Mechanical elasticity and piezoelectricity in monolayer transition-metal T


dichalcogenide alloys
Chen Yu, Xiaoxing Chen, Chunyu Wang, Zhiguo Wang∗
School of Electronic Science and Engineering, Center for Public Security Technology, University of Electronic Science and Technology of China, Chengdu, 610054, PR China

ARTICLE INFO ABSTRACT

Keywords: Two-dimensional (2D) materials have attracted considerable attention due to their application in nanoscale
mechanical elasticity devices. The intrinsic physical properties of 2D materials are crucial for their practical applications. In this work,
Piezoelectricity the structural, electronic, mechanical and piezoelectric properties of transition metal dichalcogenide (TMD)
2D materials monolayer alloys were investigated using first-principles calculations. The results show that for MX2(1-x)Y2x
Transition metal dichalcogenide alloys
(where M = Cr, Mo and W; and X, Y = S, Se and Te) monolayer alloys, the electronic, mechanical and piezo-
Density functional theory
electric properties greatly depend on the composition. The band gap decreases with increasing x in MX2(1-x)Y2x
alloys if the Y atom has a larger atomic number than the X atom. The in-plane stiffness and piezoelectric
coefficients of 2D MX2(1-x)Y2x alloys lie in the range between that of MX2 to MY2 monolayers. Thus, these
properties can be finely tuned by modifying the chemical composition. These results provide a useful guideline
for the design of nanoscale devices based on 2D TMD alloys.

1. Introduction study of 2D piezoelectric materials. The piezoelectric properties of


many 2D monolayer materials have been well studied using density
Two-dimensional (2D) materials have attracted much attention in functional theory (DFT), such as graphene [16], TMD monolayers
recent years owing to their extraordinary mechanical, chemical and [17,18], group-II monochalcogenides and oxides [17,19], group III
physical properties. After the successful synthesis of graphene, which monochalcogenides [20], group IV monochalcogenides [21], the
was exfoliated from graphite in 2004 [1], many other atomic-layered buckled monolayers of group-IV elements [22] and group V binary
materials were successfully synthesized, such as hexagonal BN (h-BN) monolayers [23], tellurene [24], platinum dichalcogenides [25], group
[2] and transition metal dichalcogenides (TMDs) [3–5]. Extensive re- III nitrides [26], vanadium dichalcogenides [27], buckled V-structure
search has shown that these mono- and few-layered materials have group III–V monolayers [28] and ternary In-containing sesquichalco-
significant potential uses in various fields such as energy conversion genides [29]. These 2D materials show great piezoelectric properties.
and storage [6], catalysis [7,8], anode materials [9], phototransistors Moreover, the piezoelectric properties of some materials have been
[10] and spintronics [11]. experimentally studied. For instance, Lee et al. [30] measured the
Piezoelectric materials can convert applied mechanical energy into piezoelectric coefficient of large monolayer WSe2 using a lateral piezo-
electrical energy. Conversely, application of an electric field to such response force microscopy and Esfahani et al. [31] developed a lateral
materials can result in a change in mechanical strain. The strain-de- excitation scanning probe microscopy technique to measure a WSe2
pendent charge output of piezoelectric materials can be typically used sample – the measured piezoelectric coefficients were 3.26 and 5.2 pm/
for power harvesting and sensor applications. Research in nanoscience V, respectively, in good agreement with previous calculations [18,32].
has developed novel piezoelectric materials for next generation nano- Moreover, the study of piezoelectricity is not limited to in-plane pie-
generators. The 2D materials theoretically predicted to show electric zoelectric properties. Wu et al. [14] measured the piezoelectric re-
polarization [12,13] have attracted much attention since experimental sponse of MoS2 flakes and found that piezoelectric output decreased
observation of the piezoelectric effect in a MoS2 monolayer [14,15]. with increasing numbers of atomic layers and disappeared for an even
The MoS2 monolayer (d = 3.73 pm/V) [14,15] exhibits a piezoelectric number of layers. Based on theoretical and experimental studies, 2D
coefficient comparable to that found for 3D wurtzite structure mate- piezoelectric devices have already been demonstrated for many prac-
rials, such as ZnO (d = 9.93 pm/V), which has triggered growth in the tical applications, including energy conversion and electronics [14],


Corresponding author.
E-mail address: zgwang@uestc.edu.cn (Z. Wang).

https://doi.org/10.1016/j.jpcs.2019.109081
Received 11 February 2019; Received in revised form 26 June 2019; Accepted 27 June 2019
Available online 27 June 2019
0022-3697/ © 2019 Elsevier Ltd. All rights reserved.
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

nanoelectromechanical systems [15], strain sensors [33], piezo-photo-


tronics [34,35] and nanogenerators [30].
Piezoelectric materials with intrinsic non-centrosymmetric crystal
structures [36] allow for conversion between electrical and mechanical
energy. Bulk and 2D MoS2 with an even number of layers exhibit
centrosymmetric structures, which do not show piezoelectric proper-
ties. The symmetry is broken in 2D MoS2 with an odd number of layers,
leading to in-plane intrinsic piezoelectricity [15]. Other approaches
such as surface adsorption can artificially break inversion symmetry
and induce piezoelectricity in 2D materials [16,37]. In addition to in-
plane piezoelectric properties, the out-of-plane piezoelectric coeffi-
cients of Janus monolayers such as MoSSe and Ga2SSe have been cal-
culated [38,39].
Monolayer TMDs (MX2) consist of one layer of transition metal
atoms sandwiched between two layers of chalcogen atoms. Due to the
different electronegativities for transition metal and chalcogen atoms, it
was reported that the electronic properties of monolayer MX2 can be
tuned by substitution of M with Mʹ atoms or X with Xʹ atoms to form
ternary monolayer MxMʹ(1-x)X2 or MX2xXʹ2(1-x) alloys [40–46]. Ternary
TMD alloys have also been successfully synthesized in experiments
[47–50]. The Mo1-xWxSe2 [49,51–53] and WS2(1-x)Se2x [54–56]
monolayers show tunable band gaps and different electrical properties
depending on the composition ratio x. The symmetry of the MX2
monolayer is broken by forming ternary TMD alloys. However, there is
no report on the piezoelectricity of monolayer TMD alloys. First-prin-
ciples calculations based on DFT have been widely used to predict the Fig. 1. (a) Top and (b) side views of MX2(1-x)Y2x alloy where M = Cr, Mo and W
structure, electronic properties and mechanical properties of novel and X, Y = S, Se and Te. P denotes the direction of the polarization. The
functional materials [57,58] In this work, we systematically studied the hexagonal and orthorhombic unit cell used in the calculations is labeled.
piezoelectric properties of monolayer MX2(1-x)Y2x (M = Cr, Mo and W;
and X, Y = S, Se and Te) alloys using DFT. It was shown that the
piezoelectricity of monolayer ternary TMD alloys can be continuously
tuned by varying the value of x in MX2(1-x)Y2x. The relationship be-
tween the piezoelectric coefficient and the component ratio of chal-
cogen atoms was analyzed. These results provide an effective method to
tune the piezoelectric properties of monolayer ternary TMD alloys.

2. Computational details

All DFT calculations were performed using the SIESTA code [59],
with the electron exchange–correlation interactions described by the
generalized gradient approximation in the scheme of the Perdew–Bur-
ke–Ernzerhof functional [60]. Norm-conserving pseudopotentials were
used to describe the electron–ion interactions, and double-ζ basis sets
with the polarization function were adopted to expand the valence
electron wave-functions. The structure optimizations used the con-
jugate gradient method with a Brillouin zone sampled by a 11 × 11 × 1
k-point mesh within the Monkhorst–Pack scheme [61]. All atomic co-
ordinates and lattice parameters were fully relaxed. The convergence
criteria were 10−5 eV and 0.01 eV/Å for the total energy and maximum
Fig. 2. (a)–(f) Possible atomic configurations of monolayer MX2(1-x)Y2x alloy
Hellmann–Feynman force on each atom, respectively. A 30 Å vacuum
with x = 1/2. (g) Energy difference per unit cell of monolayer MoSSe (blue),
was used in the z-direction to avoid spurious periodic interaction be-
MoSeTe (red) and MoSTe (green) alloys for three configurations. (For inter-
tween adjacent images. pretation of the references to colour in this figure legend, the reader is referred
to the Web version of this article.)
3. Results and discussion

x)Y2x alloys, there are one, three, six, three and one possible config-
The MX2 monolayer has a crystal structure consisting of X–M–X
urations for x = 0, 1/4, 1/2, 3/4 and 1, respectively. The six possible
sandwich layers with covalent bonding, with the M and X atoms ar-
configurations for monolayer MX2(1-x)Y2x alloy with x = 1/2 are shown
ranged in a graphite-like configuration. Monolayer MX2(1-x)Y2x alloy is
in Fig. 2a–f, and their relative energy in Fig. 2g. The configuration with
formed by substituting part of the X atoms with Y atoms. In this work,
the same chalcogen superimposed over each other has higher energy
the monolayer MX2(1-x)Y2x is modeled with a 2 × 2 supercell (Fig. 1),
than that with the X atom superimposed over the Y atom (Fig. 2), in-
which includes eight chalcogen atoms and four transition metal atoms.
dicating that chalcogen atoms are preferably distributed evenly. This
As zero, two, four and six X atoms are replaced by Y atoms, monolayer
configuration pattern is consistent with previous theoretical and ex-
MX2(1-x)Y2x (x = 0, 1/4, 1/2, 3/4 and 1.0) alloys are formed. All pos-
perimental research [47,62].
sible substitution configurations are investigated to find the most en-
The calculated lattice constants as a function of x in monolayer
ergetically stable structure for monolayer MX2(1-x)Y2x (x = 0, 1/4, 1/2,
MX2(1-x)Y2x alloys are shown in Fig. 3a. The lattice constants increase
¾ and 1.0) alloys. Because a 2 × 2 supercell is used to model the MX2(1-

2
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

Fig. 3. (a) Lattice constants of MX2(1-x)Y2x alloys as a function of the relative concentration of their constituents. (b) Band gaps of MX2(1-x)Y2x alloys.

with increasing x if Y has a larger atomic number than the X atoms in


the periodic table. As x increases from 0 to 1, the lattice constant in-
creases from 3.20 Å for MoS2 to 3.32 Å for MoSe2 in MoS2(1-x) Se2x alloy,
and from 3.32 Å for MoSe2 to 3.55 Å for MoTe2 in MoSe2(1-x)Te2x alloy.
The increased lattice constants result from the larger size of Y compared
to the X atom. The calculated lattice constants for MoS2, MoSe2 and
MoTe2 are in good agreement with experimental [48,63,64] and other
simulation values [32,65,66]. The lattice constant shows a linear var-
iation in the Y content that obeys the well-known Vegard's law [67],
which has been theoretically and experimentally confirmed [62,68].
The phonon dispersion spectra calculated using the frozen-phonon
method along the high-symmetry direction for monolayer MX2(1-x)Y2x
alloys are shown in Fig. 4. All the monolayer MX2(1-x)Y2x alloys are
dynamically stable with very small, insignificant imaginary frequency.
The largest imaginary frequencies are −7.1, −9.7 and −6.2 cm−1 for
CrS1/4Te3/4, MoSSe and WSTe, respectively. Although the most stable
configuration of MoTe2 is a distorted 1T-phase (1Tʹ) [5,69], the 1Tʹ-
MoTe2 monolayer is a metallic material [70] and its piezoelectric re-
sponse cannot be calculated. The 2H–MoTe2 monolayer may also be
Fig. 5. Band structures of MoX2(1-x)Y2x alloys. The vacuum energy was set to
stable due to the small imaginary frequencies (−1.9 cm−1); we con- zero; the Fermi energy level is denoted by horizontal dotted lines.
sidered the 2H-phase for all structures in the present work.
The MoX2 monolayers are direct band gap semiconductors with
both the conduction band minimum (CBM) and the valence band atomic number than X atoms in the periodic table. The band gaps
maximum (VBM) located at the K point (Fig. 5). The calculated band calculated for all MX2(1-x)Y2x alloys considered are shown in Fig. 3b. As
gaps of MoS2 and WS2 monolayers are 1.71 and 1.92 eV, respectively, x increases from 0 to 1, the band gap decreases from 1.71 eV for MoS2
which agree well with corresponding previous theoretical calculated to 1.48 eV for MoSe2 in MoS2(1-x) Se2x alloy, and from 1.48 eV for MoSe2
values of 1.67 and 1.81 eV [32]. The band gap is gradually tuned by the to 1.17 eV for MoTe2 in MoSe2(1-x)Te2x alloy. Thus, the band gap of
formation of MX2(1-x)Y2x alloys. Fig. 5 shows the evolution of the band monolayer MoX2 can be tuned by the formation of MX2(1-x)Y2x alloy
structures of MoX2(1-x)Y2x alloys with content. The direct band gap [71], which is promising for broad application prospects in the pho-
character remains, with CBM and VBM located at the K point. The band toelectric field.
gap decreases with increasing x in MX2(1-x)Y2x alloys if Y has a larger To reveal the mechanism underlying evolution of the band structure

Fig. 4. Phonon dispersion for CrX2(1-x)Y2x, MoX2(1-x)Y2x and WX2(1-x)Y2x alloys.

3
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

Table 1
Calculated relaxed-ion and clamped-ion elastic stiffness coefficients (N/m) and
Poisson's ratio v.
Material Relaxed-ion Clamped-ion

x C11 C12 v C11 C12 v

CrS2(1-x)Se2x 0 122.6 35.6 0.29 147.1 44.7 0.30


1/4 120.3 34.7 0.29 141.3 45.0 0.32
1/2 116.6 35.3 0.30 139.9 44.7 0.32
3/4 106.2 34.0 0.32 132.4 44.7 0.34
CrSe2(1-x)Te2x 0 107.6 36.3 0.34 126.6 44.1 0.35
1/4 98.9 35.0 0.35 121.1 44.1 0.36
1/2 87.0 35.6 0.41 118.1 45.5 0.39
3/4 79.8 34.5 0.43 107.1 42.8 0.40
CrS2(1-x)Te2x 1/4 118.3 34.8 0.29 134.5 45.2 0.34
1/2 108.1 36.2 0.33 127.3 46.1 0.36
3/4 93.6 33.1 0.35 114.8 43.7 0.38
1 81.7 34.7 0.42 101.6 41.8 0.41
MoS2(1-x)Se2x 0 131.7 35.7 0.27 157.4 51.1 0.32
1/4 123.5 31.1 0.25 149.8 47.6 0.32
1/2 120.9 31.4 0.26 148.4 46.2 0.31
3/4 113.2 29.2 0.26 142.8 45.3 0.32
MoSe2(1-x)Te2x 0 112.6 28.3 0.25 141.3 45.1 0.32
1/4 106.8 27.7 0.25 133.4 42.5 0.32
1/2 98.9 27.2 0.27 124.6 41.2 0.33
Fig. 6. PDOS of MoX2(1-x)Y2x monolayers. The vacuum energy was set to zero; 3/4 93.8 27.5 0.29 119.0 40.5 0.34
the Fermi energy level is indicated by dotted lines. MoS2(1-x)Te2x 1/4 115.5 28.3 0.25 145.5 45.5 0.31
1/2 102.3 26.3 0.26 141.2 42.9 0.30
3/4 96.3 27.4 0.28 124.5 41.7 0.33
of the MX2(1-x)Y2x alloys with composition x, the projected density of
1 88.7 24.1 0.27 112.9 37.0 0.33
states (PDOS) of the MoX2(1-x)Y2x alloys are shown in Fig. 6. The VBM WS2(1-x)Se2x 0 151.4 33.8 0.22 182.5 52.0 0.29
and CBM of the MoX2(1-x)Y2x alloys are composed of the d-states of Mo 1/4 145.4 32.2 0.22 174.6 50.3 0.29
and p-states of X and Y atoms. The upshift of the VBM to high energy 1/2 142.3 28.1 0.20 171.3 48.2 0.28
3/4 136.9 28.0 0.20 166.5 46.2 0.28
induces a decrease in the band gap for MoX2(1-x)Y2x alloy as x increases
WSe2(1-x)Te2x 0 131.0 28.6 0.22 161.9 43.8 0.27
from 0 to 1. 1/4 124.8 24.9 0.20 155.6 43.1 0.27
Piezoelectricity and elastic stiffness of the MX2(1-x)Y2x alloys were 1/2 123.3 25.5 0.21 148.0 41.2 0.28
investigated. To align the direction of polarization and elastic strain 3/4 117.7 24.1 0.21 140.5 39.4 0.28
along the desired direction, a rectangular supercell (Fig. 1a) is used to WS2(1-x)Te2x 1/4 138.0 27.2 0.20 172.8 47.7 0.27
1/2 121.2 21.3 0.18 156.2 44.4 0.28
calculate the piezoelectricity and elastic stiffness. The elastic properties
3/4 116.5 23.5 0.20 151.1 42.1 0.28
of MX2(1-x)Y2x alloys can be described by three independent constants: 1 109.8 22.5 0.20 135.5 36.6 0.26
the in-plane stiffness (C11 and C12) and Poisson's ratio (v). The in-plane
stiffness is defined as C11 = (1/A0) × (∂2ΔE/ 11 2
) and C12 = (1/
2
A0) × (∂ ΔE/∂ε11∂ε12), where A0 is the equilibrium area of the 2D The in-plane stiffness decreases as X changes from S to Te in the
monolayer, ε is the uniaxial strain applied to the in-plane direction and periodic table in monolayer MX2 (Table 1). For example, in-plane
ΔE = ET(ε) − ET(0), where ET(0) is equilibrium total energy and ET(ε) is stiffness is 121.0, 105.5 and 82.2 N/m for MoS2, MoSe2 and MoTe2
total energy of the monolayer with strain ε. The v, defined as monolayers, respectively. The in-plane stiffness increases when transi-
v = −εtrans/εaxial, governs the transverse variation in response to axial tion metal elements change from Cr to W in the periodic table. These
strain. The elastic stiffness was calculated using the scheme developed results indicate that a monolayer composed of a high atomic-number
by Topsakal et al. [72]. The strains along the armchair and zigzag di- transition metal and low atomic-number chalcogen atom will be strong.
rections, ε11 and ε22, respectively, were varied from −0.02 to 0.02 with The in-plane stiffness of 2D MX2(1-x)Y2x alloy is in the range between
steps of 0.01. The strained monolayer is realized by elongation or that of MX2 to MY2 monolayers. For example, in-plane stiffness of
shrinkage of the lattice constant, i.e. the deformed lattice constant is monolayer MoS2(1-x)Se2x decreases as x increases from 0 to 1, thus in-
a = a0(1 + ε), where a0 is the equilibrium lattice constant. The total plane stiffness of monolayer MoX2 can be tuned by formation of MX2(1-
energy of the monolayer under varying strain is calculated, resulting in x)Y2x alloys.
a grid containing 25 data points. By fitting these energy data to the The piezoelectric effect can be described by third-rank piezoelectric
quadratic polynomial, ΔE = a1ε112 + a2ε222 + a3ε11ε22, the three tensors including piezoelectric stress coefficients (eijk) and piezoelectric
coefficients can be obtained. Due to isotropy in the structure symmetry, strain coefficients (dijk) as follows [32]:
a1 = a2. The in-plane elastic stiffness can be calculated: C11 = 2a1/A0
Pi Pi
and C12 = a3/A0. Poisson's ratio is calculated with v = a3/2a1 [72]. eijk = , dijk =
The calculated relaxed- and clamped-ion elastic constants, in-plane jk jk (1)
stiffness and Poisson's ratio are listed in Table 1 – these results are in where Pi is the piezoelectric polarization along the direction i induced
good agreement with previous DFT calculations [17,18]. For example, by applied strains along the direction j and k, and εjk and σjk represent
the relaxed-ion in-plane stiffness C11 of MoS2 (131.7 N/m) is compar- strains and stresses, respectively. The indices i, j and k correspond to the
able with reported values of 130 N/m [32] and 132.7 N/m [18]; and the x (armchair), y (zigzag) and z (vertical) directions, respectively. Due to
relaxed-ion C11 of CrS2 (122.6 N/m) is in agreement with reported the symmetry of these 2D hexagonal crystals, the number of in-
values of 122.5 N/m [17] and 120.6 N/m [18]. The elastic constants are dependent components of the piezoelectric tensor is reduced. The e11
all positive for all MX2(1-x)Y2x alloys, and they satisfy the mechanical and d11 (Voigt notation is employed to limit the number of indices)
stability criteria for 2D hexagonal crystals: C11 > 0, C11–C12 > 0 and values were calculated and used to characterize the in-plane piezo-
C11 > |C12| [73], indicating that these 2D MX2(1-x)Y2x alloys are me- electric properties of 2D hexagonal crystals. These piezoelectric coef-
chanically stable. ficients are correlated with the in-plane elastic constants:

4
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

number transition metal and a high atomic-number chalcogen atom has


large relaxed-ion piezoelectric coefficients. For a given transition metal
element M, the piezoelectric coefficients d11 and e11 increase when the
chalcogen element changes from S to Te. In addition, for a given
chalcogen atom, the piezoelectric coefficients d11 and e11 decrease as
the transition metal atom changes from Cr to W. Thus, monolayer WS2
possesses the smallest relaxed-ion e11 (2.51 × 10−10 C/m) and d11
(2.13 pm/V) piezoelectric coefficients, and correspondingly monolayer
CrTe2 possesses the largest among the studied MX2 monolayers with
6.39 × 10−10 C/m and 13.60 pm/V.
Evolution of the piezoelectric coefficients d11 and e11 for the 2D
MX2(1-x)Y2x alloys with content is shown in Fig. 8a and b, respectively.
The values of the piezoelectric coefficients lie in the range between that
of MX2 and MY2 and increase with increasing Y content if the atomic
number of Y is larger than X. For example, the piezoelectric coefficients
of the WS2(1-x)Se2x alloys are 2.26, 2.33 and 2.53 pm/V for x = 1/4, 1/2
and 3/4, respectively, which are larger than that of monolayer WS2
(2.13 pm/V) and smaller than that of monolayer WSe2 (2.70 pm/V).
The result shows that alloying is an effective way to engineer the pie-
zoelectricity of TMD alloys, and this feature can be useful to find and
synthesize appropriate materials for a variety of applications. Con-
sidering many monolayer alloys have been successfully synthesized in
experiments, such as MoSe2(1-x)Te2x, MoS2(1-x)Se2x, WS2(1-x)Se2x and
WSe2(1-x)Te2x [48–50,55,77], other considered structures are expected
to be successfully synthesized with suitable approaches.
The piezoelectric coefficients are much smaller for the clamped-ion
than the relaxed-ion (Fig. 8). Thus, the polarization significantly in-
creases after relaxation of the ion positions under strain. Detailed
analysis of the relaxed atomic structure shows that with increasing
atomic number for chalcogen elements in the same group, the internal
Fig. 7. (a) Polarization change under an applied uniaxial strain (ε11) along the relaxation contributions to the piezoelectric coefficients become gen-
x-axis for monolayer CrX2(1-x)Y2x, MoX2(1-x)Y2x and WX2(1-x)Y2x. Piezoelectric erally more important. This trend can be interpreted as due to the large
coefficient e11 is determined from the slope of the curves. The change in M–X
motion of the ion in response to an applied strain. The changes in M–X
bond length as a function of uniaxial strain for the (b) relaxed- and (c) clamped-
bond length as a function of uniaxial strain are shown in Fig. 7b and c
ion conditions.
for the relaxed- and clamped-ion conditions, respectively. The change
in bond length with uniaxial strain can be fitted linearly. Differences in
e11 = d11 (C11 C12) (2) slopes between the clamped- and relaxed-ion conditions show an in-
crease as the chalcogen elements change from S to Te (Table 2), in-
In this study, e11 was obtained by fitting the differential change in
dicating that internal relaxation mainly contributes to the increased
the piezoelectric polarization per area along the armchair direction as a
piezoelectric coefficients.
function of the uniaxial strain ε11 ranging from −0.01 to 0.01 with a
The physical properties of alloy MX2(1−x)Y2x as a function of con-
step of 0.005, i.e. e11 was calculated using the Berry-phase method [74]
centration often obey a quadratic rule. The relationship between pie-
as implemented within the SIESTA code:
zoelectricity and concentration x for MX2(1−x)Y2x alloy was fitted using
P ( 11) P (0) = e11 11 (3) the bowing equation:

d11(MX2(1-x)Y2x) = (1 − x)d11(MX2) + xd11(MY2) − bx(1 − x) (4)


Fig. 7a shows the evolution of the polarizations in the monolayers
MX2(1-x)Y2x (M = Cr, Mo and W) along the x direction under applied where d11(MX2(1-x)Y2x), d11(MX2) and d11(MY2) are piezoelectric coef-
uniaxial strain ε11; the piezoelectric coefficient e11 can be derived from ficients of MX2(1-x)Y2x, MX2 and MY2 monolayers, respectively; and b is
the slope of the curves. The piezoelectric coefficients d11 can be esti- the bowing parameter, which accounts for nonlinear dependence on
mated using equation (2). The calculated piezoelectric coefficients d11 composition x. The calculated values of the bowing parameter are listed
and e11 of the 2D MX2(1-x)Y2x alloys are shown in Fig. 8a and b, re- in Table 3. The bowing parameters follow the sequence
spectively. The relaxed-ion piezoelectric coefficients d11 for these |b(S,Te)| > |b(Se,Te)| > |b(S,Se)|, indicating that polarizability increases
monolayer MX2(1-x)Y2x alloys lie in the range of 2.13–13.60 pm/V, from S–Se to Se–Te to S–Te. The Cr-based alloys have the largest value
which agrees with other theoretical values [17,18]. The predicted re- for the bowing parameter, indicating that cation polarization can
laxed-ion piezoelectric coefficient d11 for the WSe2 monolayer (2.70 counteract the change in the piezoelectricity.
pm/V) is in good agreement with the experimentally measured result of
3.26 pm/V [30].
The calculated piezoelectric coefficients, such as CrS2 (5.42 pm/V), 4. Conclusion
CrTe2 (13.60 pm/V), MoS2 (3.89 pm/V), MoTe2 (7.16 pm/V), WS2
(2.13 pm/V) and WTe2 (4.64 pm/V), are also consistent with the In this work, the structural, electronic, mechanical and piezoelectric
available reference data of 6.15, 13.45, 3.73, 7.39, 2.12 and 4.60 pm/V, properties of TMD monolayer alloys were investigated using DFT cal-
respectively [17,32]. These monolayer MX2(1-x)Y2x alloys have piezo- culations. The results show that these ternary alloys have comparable
electric coefficients that are the same order of magnitude as that of or even better piezoelectric properties than extensively used bulk pie-
many important piezoelectric bulk materials, including α-quartz (2.3 zoelectric materials. The electronic, mechanical and piezoelectric
pm/V) [75], wurtzite GaN (3.1 pm/V) [76] and wurtzite AlN (5.1 pm/ properties can be modulated by changing the MX2(1-x)Y2x alloy com-
V) [76]. The monolayer MX2(1-x)Y2x alloy composed of a low atomic- position. The band gap decreases with increasing x in MX2(1-x)Y2x alloys

5
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

Fig. 8. Calculated (a) piezoelectric strain (d11) and (b) piezoelectric stress (e11) coefficients. Dashed lines denote clamped-ion coefficients, whereas the solid line
denotes relaxed-ion coefficients.

Table 2 [8] Y. Chen, H. Sun, W. Peng, 2D transition metal dichalcogenides and graphene-based
Differences in fitting slope for the change in the M–X bond length (Å) between ternary composites for photocatalytic hydrogen evolution and pollutants degrada-
tion, Nanomaterials 7 (2017) 62.
clamped- and relaxed-ion conditions.
[9] H. Shu, F. Li, C. Hu, P. Liang, D. Cao, X. Chen, The capacity fading mechanism and
Material MS2 MSe2 MTe2 improvement of cycling stability in MoS2-based anode materials for lithium-ion
batteries, Nanoscale 8 (2016) 2918–2926.
Cr 0.27 0.39 0.43 [10] Y. Wang, E. Liu, A. Gao, T. Cao, M. Long, C. Pan, L. Zhang, J. Zeng, C. Wang, W. Hu,
S.J. Liang, F. Miao, Negative photoconductance in van der Waals heterostructure-
Mo 0.42 0.45 0.48
based floating gate phototransistor, ACS Nano 12 (2018) 9513–9520.
W 0.39 0.43 0.50
[11] X. Chen, T. Yan, B. Zhu, S. Yang, X. Cui, Optical control of spin polarization in
monolayer transition metal dichalcogenides, ACS Nano 11 (2017) 1581–1587.
[12] E.J. Mele, P. Král, Electric polarization of heteropolar nanotubes as a geometric
Table 3 phase, Phys. Rev. Lett. 88 (2002) 056803.
[13] N. Sai, E.J. Mele, Microscopic theory for nanotube piezoelectricity, Phys. Rev. B 68
Bowing parameters for piezoelectric coefficients d11 of MX2(1−x)Y2x alloys.
(2003) 241405.
MS2(1-x)Se2x MSe2(1-x)Te2x MS2(1-x)Te2x [14] W. Wu, L. Wang, Y. Li, F. Zhang, L. Lin, S. Niu, D. Chenet, X. Zhang, Y. Hao,
T.F. Heinz, J. Hone, Z.L. Wang, Piezoelectricity of single-atomic-layer MoS2 for
Cr 1.18 −2.82 8.67 energy conversion and piezotronics, Nature 514 (2014) 470.
Mo −0.27 1.37 2.51 [15] H. Zhu, Y. Wang, J. Xiao, M. Liu, S. Xiong, Z.J. Wong, Z. Ye, Y. Ye, X. Yin, X. Zhang,
Observation of piezoelectricity in free-standing monolayer MoS2, Nat. Nanotechnol.
W 0.23 1.47 1.80
10 (2014) 151.
[16] M.T. Ong, E.J. Reed, Engineered piezoelectricity in graphene, ACS Nano 6 (2012)
1387–1394.
if Y has a larger atomic number than the X atoms in the periodic table. [17] M.N. Blonsky, H.L. Zhuang, A.K. Singh, R.G. Hennig, Ab Initio prediction of pie-
zoelectricity in two-dimensional materials, ACS Nano 9 (2015) 9885–9891.
The in-plane stiffness and piezoelectric coefficients of the 2D MX2(1-
[18] M.M. Alyörük, Y. Aierken, D. Çakır, F.M. Peeters, C. Sevik, Promising piezoelectric
x)Y2x alloys lie in the range between that of MX2 to MY2 monolayers. performance of single layer transition-metal michalcogenides and dioxides, J. Phys.
Especially, CrSe2(1-x)Te2x and CrS2(1-x)Te2x alloys have large piezo- Chem. C 119 (2015) 23231–23237.
electric coefficients d11, and should be synthesized experimentally for [19] C. Sevik, D. Çakır, O. Gülseren, F.M. Peeters, Peculiar piezoelectric properties of soft
two-dimensional materials, J. Phys. Chem. C 120 (2016) 13948–13953.
designing novel nanoscale piezoelectric or piezo-phototronic devices. [20] L. Huang, N. Huo, Y. Li, H. Chen, J. Yang, Z. Wei, J. Li, S.-S. Li, Electric-field tunable
These results provide valuable guidance for experiments and promise band offsets in black phosphorus and MoS2 van der Waals p-n heterostructure, J.
for novel nanoscale piezoelectric applications. Phys. Chem. Lett. 6 (2015) 2483–2488.
[21] R. Fei, W. Li, J. Li, L. Yang, Giant piezoelectricity of monolayer group IV mono-
chalcogenides: SnSe, SnS, GeSe, and GeS, Appl. Phys. Lett. 107 (2015) 173104.
References [22] J. Shi, Y. Gao, X.L. Wang, S.N. Yun, Electronic, elastic and piezoelectric properties
of two-dimensional group-IV buckled monolayers, Chin. Phys. Lett. 34 (2017)
087701.
[1] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos,
[23] H. Yin, J. Gao, G.-P. Zheng, Y. Wang, Y. Ma, Giant piezoelectric effects in monolayer
I.V. Grigorieva, A.A. Firsov, Electric field effect in atomically thin carbon films,
Group-V binary compounds with honeycomb phases: a first-principles prediction, J.
Science 306 (2004) 666.
Phys. Chem. C 121 (2017) 25576–25584.
[2] C. Jin, F. Lin, K. Suenaga, S. Iijima, Fabrication of a freestanding boron nitride
[24] Y. Chen, J. Liu, J. Yu, Y. Guo, Q. Sun, Symmetry-breaking induced large piezo-
single layer and its defect assignments, Phys. Rev. Lett. 102 (2009) 195505.
electricity in Janus tellurene materials, Phys. Chem. Chem. Phys. 21 (2019)
[3] Q.H. Wang, K. Kalantar-Zadeh, A. Kis, J.N. Coleman, M.S. Strano, Electronics and
1207–1216.
optoelectronics of two-dimensional transition metal dichalcogenides, Nat.
[25] Z. Kahraman, A. Kandemir, M. Yagmurcukardes, H. Sahin, Single-layer Janus-type
Nanotechnol. 7 (2012) 699.
platinum dichalcogenides and their heterostructures, J. Phys. Chem. C 123 (2019)
[4] Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey, H. Zhang, Single-layer
4549–4557.
semiconducting nanosheets: high-yield preparation and device fabrication, Angew.
[26] Y. Guo, H. Zhu, Q. Wang, Piezoelectric effects in surface-engineered two-dimen-
Chem. Int. Ed. 50 (2011) 11093–11097.
sional Group III nitrides, ACS Appl. Mater. Interfaces 11 (2019) 1033–1039.
[5] C.H. Naylor, W.M. Parkin, J. Ping, Z. Gao, Y.R. Zhou, Y. Kim, F. Streller,
[27] J. Yang, A. Wang, S. Zhang, J. Liu, Z. Zhong, L. Chen, Coexistence of piezoelectricity
R.W. Carpick, A.M. Rappe, M. Drndić, J.M. Kikkawa, A.T.C. Johnson, Monolayer
and magnetism in two-dimensional vanadium dichalcogenides, Phys. Chem. Chem.
single-crystal 1T′-MoTe2 grown by chemical vapor deposition exhibits weak antil-
Phys. 21 (2019) 132–136.
ocalization effect, Nano Lett. 16 (2016) 4297–4304.
[28] S. Norouzi, N. Shahtahmassebi, M. Behdani, M. Rezaee Roknabadi, High piezo-
[6] F. Bonaccorso, L. Colombo, G. Yu, M. Stoller, V. Tozzini, A.C. Ferrari, R.S. Ruoff,
electricity in the buckled V-structure monolayers of group III-V: an Ab initio cal-
V. Pellegrini, Graphene, related two-dimensional crystals, and hybrid systems for
culation, Phys. E Low-dimens. Syst. Nanostruct. 102 (2018) 88–94.
energy conversion and storage, Science 347 (2015) 1246501.
[29] H. Yin, G.-P. Zheng, Y. Wang, B. Yao, New monolayer ternary In-containing ses-
[7] D. Voiry, J. Yang, M. Chhowalla, Recent strategies for improving the catalytic ac-
quichalcogenides BiInSe3, SbInSe3, BiInTe3, and SbInTe3 with high stability and
tivity of 2D TMD nanosheets toward the hydrogen evolution reaction, Adv. Mater.
extraordinary piezoelectric properties, Phys. Chem. Chem. Phys. 20 (2018)
28 (2016) 6197–6206.

6
C. Yu, et al. Journal of Physics and Chemistry of Solids 135 (2019) 109081

19177–19187. monolayer Mo1−xWxSe2, Appl. Phys. Lett. 104 (2014) 012101.


[30] J.H. Lee, J.Y. Park, E.B. Cho, T.Y. Kim, S.A. Han, T.H. Kim, Y. Liu, S.K. Kim, [52] F. Chen, S. Ding, W. Su, The synthesis and tunable optical properties of two-di-
C.J. Roh, H.J. Yoon, H. Ryu, W. Seung, J.S. Lee, J. Lee, S.W. Kim, Reliable piezo- mensional alloyed Mo1-xWxS2 monolayer with in-plane composition modulations
electricity in bilayer WSe2 for piezoelectric nanogenerators, Adv. Mater. 29 (2017) (0≤x≤1), J. Alloy. Comp. 784 (2019) 213–219.
1606667. [53] J. Mann, Q. Ma, P.M. Odenthal, M. Isarraraz, D. Le, E. Preciado, D. Barroso,
[31] E. Nasr Esfahani, T. Li, B. Huang, X. Xu, J. Li, Piezoelectricity of atomically thin K. Yamaguchi, G. von Son Palacio, A. Nguyen, T. Tran, M. Wurch, A. Nguyen,
WSe2 via laterally excited scanning probe microscopy, Nano Energy 52 (2018) V. Klee, S. Bobek, D. Sun, T.F. Heinz, T.S. Rahman, R. Kawakami, L. Bartels, 2-
117–122. dimensional transition metal dichalcogenides with tunabled direct band gaps:
[32] K.A.N. Duerloo, M.T. Ong, E.J. Reed, Intrinsic piezoelectricity in two-dimensional MoS2(1–x)Se2x monolayers, Adv. Mater. 26 (2014) 1399–1404.
materials, J. Phys. Chem. Lett. 3 (2012) 2871–2876. [54] X. Duan, C. Wang, F. Zheng, G. Hao, L. Kou, U. Halim, H. Li, X. Wu, Y. Wang,
[33] L. Chen, F. Xue, X. Li, X. Huang, L. Wang, J. Kou, Z.L. Wang, Strain-gated field effect J. Jiang, Synthesis of WS2xSe2–2x alloy nanosheets with composition-tunable elec-
transistor of a MoS2–ZnO 2D–1D hybrid structure, ACS Nano 10 (2016) 1546–1551. tronic properties, Nano Lett. 16 (2016) 264.
[34] F. Xue, L. Chen, J. Chen, J. Liu, L. Wang, M. Chen, Y. Pang, X. Yang, G. Gao, J. Zhai, [55] Q. Fu, L. Yang, W. Wang, A. Han, J. Huang, P. Du, Z. Fan, J. Zhang, B. Xiang,
Z.L. Wang, p-type MoS2 and n-type ZnO diode and its performance enhancement by Synthesis and enhanced electrochemical catalytic performance of monolayer
the piezophototronic effect, Adv. Mater. 28 (2016) 3391–3398. WS2(1–x)Se2x with a tunable band gap, Adv. Mater. 27 (2015) 4732–4738.
[35] P. Lin, L. Zhu, D. Li, L. Xu, C. Pan, Z. Wang, Piezo-phototronic effect for enhanced [56] Y. Meng, T. Wang, Z. Li, Y. Qin, Z. Lian, Y. Chen, M.C. Lucking, K. Beach,
flexible MoS2/WSe2 van der Waals photodiodes, Adv. Funct. Mater. 28 (2018) T. Taniguchi, K. Watanabe, S. Tongay, F. Song, H. Terrones, S.F. Shi, Excitonic
1802849. complexes and emerging interlayer electron–phonon coupling in BN encapsulated
[36] E. Sun, W. Cao, Relaxor-based ferroelectric single crystals: growth, domain en- monolayer semiconductor alloy: WS0.6Se1.4, Nano Lett. 19 (2019) 299–307.
gineering, characterization and applications, Prog. Mater. Sci. 65 (2014) 124–210. [57] Q. Fan, R. Niu, W. Zhang, W. Zhang, Y. Ding, S. Yun, t-Si, A novel silicon allotrope,
[37] M.T. Ong, K.-A.N. Duerloo, E.J. Reed, The effect of hydrogen and fluorine coad- ChemPhysChem 20 (2019) 128–133.
sorption on the piezoelectric properties of graphene, J. Phys. Chem. C 117 (2013) [58] Q. Fan, W. Zhang, S. Yun, J. Xu, Y. Song, III-nitride polymorphs: XN (X=Al, Ga, In)
3615–3620. in the Pnma phase, Chem. Eur J. 24 (2018) 17280–17287.
[38] L. Dong, J. Lou, V.B. Shenoy, Large in-plane and vertical piezoelectricity in Janus [59] M.S. José, A. Emilio, D.G. Julian, G. Alberto, J. Javier, O. Pablo, S.P. Daniel, The
transition metal dichalchogenides, ACS Nano 11 (2017) 8242–8248. SIESTA method for ab initio order-N materials simulation, J. Phys. Condens. Matter
[39] Y. Guo, S. Zhou, Y. Bai, J. Zhao, Enhanced piezoelectric effect in Janus group-III 14 (2002) 2745.
chalcogenide monolayers, Appl. Phys. Lett. 110 (2017) 163102. [60] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
[40] M. Zhang, J. Wu, Y. Zhu, D.O. Dumcenco, J. Hong, N. Mao, S. Deng, Y. Chen, simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
Y. Yang, C. Jin, Two-dimensional molybdenum tungsten diselenide alloys: photo- [61] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Phys. Rev.
luminescence, Raman scattering, and electrical transport, ACS Nano 8 (2014) B 13 (1976) 5188–5192.
7130–7137. [62] H.P. Komsa, A.V. Krasheninnikov, Two-dimensional transition metal dichalco-
[41] C. Xia, J. An, S. Wei, Y. Jia, Q. Zhang, Electronic structures and optical properties of genide alloys: stability and electronic properties, J. Phys. Chem. Lett. 3 (2012)
SnSe2(1−x)O2x alloys, Comput. Mater. Sci. 95 (2014) 712–717. 3652–3656.
[42] S.H. Su, C.H. Chen, M.H. Chiu, W.H. Chang, J.H. He, L.J. Li, Controllable synthesis [63] P.A. Young, Lattice parameter measurements on molybdenum disulphide, J. Phys. D
of band-gap-tunable and monolayer transition-metal dichalcogenide alloys, Front. Appl. Phys. 1 (1968) 936–938.
Energy Res. 2 (2014) 1–8. [64] F. Tumino, C.S. Casari, M. Passoni, V. Russo, A. Li Bassi, Pulsed laser deposition of
[43] H.P. Komsa, A.V. Krasheninnikov, Two-dimensional transition metal dichalco- single-layer MoS2 on Au(111): from nanosized crystals to large-area films,
genide alloys: stability and electronic properties, J. Phys. Chem. Lett. 3 (2012) Nanoscale Adv. 1 (2019) 643–655.
3652. [65] F.A. Rasmussen, K.S. Thygesen, Computational 2D materials database: electronic
[44] Y. Wang, L. Huang, B. Li, J. Shang, C. Xia, C. Fan, H.X. Deng, Z. Wei, J. Li, structure of transition-metal dichalcogenides and oxides, J. Phys. Chem. C 119
Composition-tunable 2D SnSe2(1−x)S2x alloys towards efficient bandgap en- (2015) 13169–13183.
gineering and high performance (opto)electronics, J. Mater. Chem. C 5 (2016) [66] C. Ataca, M. Topsakal, E. Aktürk, S. Ciraci, A comparative study of lattice dynamics
84–90. of three- and two-dimensional MoS2, J. Phys. Chem. C 115 (2011) 16354–16361.
[45] Y. Huang, X. Chen, D. Zhou, H. Liu, C. Wang, J. Du, L. Ning, S. Wang, Stabilities, [67] A.R. Denton, N.W. Ashcroft, Vegard's law, Phys. Rev. A 43 (1991) 3161–3164.
electronic and optical properties of SnSe2(1–x)S2x alloys: a first-principles study, J. [68] L.F. Schneemeyer, M.J. Sienko, Crystal data for mixed-anion molybdenum dichal-
Phys. Chem. C 120 (2016) 5839–5847. cogenides, Inorg. Chem. 19 (1980) 789–791.
[46] T. Pei, L. Bao, G. Wang, R. Ma, H. Yang, J. Li, C. Gu, S. Pantelides, S. Du, H. Gao, [69] Y. Zhou, E.J. Reed, Structural phase stability control of monolayer MoTe2 with
Few-layer SnSe2 transistors with high on/off ratios, Appl. Phys. Lett. 108 (2016) adsorbed atoms and molecules, J. Phys. Chem. C 119 (2015) 21674–21680.
10451. [70] W. Shi, Z. Wang, Mechanical and electronic properties of Janus monolayer transi-
[47] Y. Gong, Z. Liu, A.R. Lupini, G. Shi, J. Lin, S. Najmaei, Z. Lin, A.L. Elías, tion metal dichalcogenides, J. Phys. Condens. Matter 30 (2018) 215301.
A. Berkdemir, G. You, H. Terrones, M. Terrones, R. Vajtai, S.T. Pantelides, [71] J. Kang, S. Tongay, J. Li, J. Wu, Monolayer semiconducting transition metal di-
S.J. Pennycook, J. Lou, W. Zhou, P.M. Ajayan, Band gap engineering and layer-by- chalcogenide alloys: stability and band bowing, J. Appl. Phys. 113 (2013) 143703.
layer mapping of selenium-doped molybdenum disulfide, Nano Lett. 14 (2014) [72] M. Topsakal, S. Cahangirov, S. Ciraci, The response of mechanical and electronic
442–449. properties of graphane to the elastic strain, Appl. Phys. Lett. 96 (2010) 091912.
[48] A. Apte, A. Krishnamoorthy, J.A. Hachtel, S. Susarla, J.C. Idrobo, A. Nakano, [73] F. Mouhat, F.X. Coudert, Necessary and sufficient elastic stability conditions in
R.K. Kalia, P. Vashishta, C.S. Tiwary, P.M. Ajayan, Telluride-based atomically thin various crystal systems, Phys. Rev. B 90 (2014) 224104.
layers of ternary two-dimensional transition metal dichalcogenide alloys, Chem. [74] D. Vanderbilt, Berry-phase theory of proper piezoelectric response, J. Phys. Chem.
Mater. 30 (2018) 7262–7268. Solids 61 (2000) 147–151.
[49] X. Duan, C. Wang, Z. Fan, G. Hao, L. Kou, U. Halim, H. Li, X. Wu, Y. Wang, J. Jiang, [75] R. Bechmann, Elastic and piezoelectric constants of alpha-quartz, Phys. Rev. 110
A. Pan, Y. Huang, R. Yu, X. Duan, Synthesis of WS2xSe2–2x alloy nanosheets with (1958) 1060–1061.
composition-tunable electronic properties, Nano Lett. 16 (2016) 264–269. [76] C.M. Lueng, H.L.W. Chan, C. Surya, C.L. Choy, Piezoelectric coefficient of alu-
[50] P. Yu, J. Lin, L. Sun, Q.L. Le, X. Yu, G. Gao, C.H. Hsu, D. Wu, T.R. Chang, Q. Zeng, minum nitride and gallium nitride, J. Appl. Phys. 88 (2000) 5360–5363.
F. Liu, Q.J. Wang, H.T. Jeng, H. Lin, A. Trampert, Z. Shen, K. Suenaga, Z. Liu, [77] Q. Gong, L. Cheng, C. Liu, M. Zhang, Q. Feng, H. Ye, M. Zeng, L. Xie, Z. Liu, Y. Li,
Metal–semiconductor phase-transition in WSe2(1-x)Te2x monolayer, Adv. Mater. 29 Ultrathin MoS2(1–x)Se2x alloy nanoflakes for electrocatalytic hydrogen evolution
(2017) 1603991. reaction, ACS Catal. 5 (2015) 2213–2219.
[51] S. Tongay, D.S. Narang, J. Kang, W. Fan, Two-dimensional semiconductor alloys:

Você também pode gostar