Novel Two-Dimensional Layered Mosi Z (Z P, As) : New Promising Optoelectronic Materials

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

nanomaterials

Article
Novel Two-Dimensional Layered MoSi2Z4 (Z = P, As): New
Promising Optoelectronic Materials
Hui Yao 1,2 , Chao Zhang 3 , Qiang Wang 1 , Jianwei Li 1 , Yunjin Yu 1, * , Fuming Xu 1 , Bin Wang 1, *
and Yadong Wei 1

1 College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China;
[email protected] (H.Y.); [email protected] (Q.W.); [email protected] (J.L.);
[email protected] (F.X.); [email protected] (Y.W.)
2 Laboratory of Optoelectronic Devices and Systems of Ministry of Education and Guangdong Province,
College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
3 Beijing Computational Science Research Center, Beijing 100193, China; [email protected]
* Correspondence: [email protected] (Y.Y.); [email protected] (B.W.)

Abstract: Very recently, two new two-dimensional (2D) layered semi-conducting materials MoSi2 N4
and WSi2 N4 were successfully synthesized in experiments, and a large family of these two 2D
materials, namely MA2 Z4 , was also predicted theoretically (Science, 369, 670 (2020)). Motivated
by this exciting family, in this work, we systematically investigate the mechanical, electronic and
optical properties of monolayer and bilayer MoSi2 P4 and MoSi2 As4 by using the first-principles
calculation method. Numerical results indicate that both monolayer and bilayer MoSi2 Z4 (Z = P,

 As) present good structural stability, isotropic mechanical parameters, moderate bandgap, favorable
carrier mobilities, remarkable optical absorption, superior photon responsivity and external quantum
Citation: Yao, H.; Zhang, C.; Wang,
efficiency. Especially, due to the wave-functions of band edges dominated by d orbital of the middle-
Q.; Li, J.; Yu, Y.; Xu, F.; Wang, B.; Wei,
layer Mo atoms are screened effectively, the bandgap and optical absorption hardly depend on the
Y. Novel Two-Dimensional Layered
MoSi2 Z4 (Z = P, As): New Promising
number of layers, providing an added convenience in the experimental fabrication of few-layer
Optoelectronic Materials. MoSi2 Z4 -based electronic and optoelectronic devices. We also build a monolayer MoSi2 Z4 -based
Nanomaterials 2021, 11, 559. https:// 2D optoelectronic device, and quantitatively evaluate the photocurrent as a function of energy and
doi.org/10.3390/nano11030559 polarization angle of the incident light. Our investigation verifies the excellent performance of a
few-layer MoSi2 Z4 and expands their potential application in nanoscale electronic and optoelectronic
Academic Editor: Jung-Ho Yun, devices.
Jeonghun Kim and Francisco Javier
García Ruiz Keywords: DFT; remarkable optical absorption; superior external quantum efficiency; optoelectronic
devices
Received: 3 February 2021
Accepted: 20 February 2021
Published: 24 February 2021

1. Introduction
Publisher’s Note: MDPI stays neutral
Two-dimensional (2D) materials have attracted extensive attention due to their distinc-
with regard to jurisdictional claims in
tive physical and material properties and the potential application on account of monolayer
published maps and institutional affil-
limit [1–9]. As a typical representative, graphene has been widely expected to be a proper
iations.
material for the preparation of a new generation of nanoelectronic devices due to remark-
able high carrier mobility, but its zero bandgap reminds us that it may not be an effective
solution [1,10,11]. Transition metal dichalcogenides (TMDCs) possess tunable bandgap,
but its comparatively low carrier mobilities is a not neglectable obstacle for practical appli-
Copyright: © 2020 by the authors.
cations [3,12–14]. For example, the carrier mobility of MoS2 is roughly 72 cm2 V−1 s−1 for
Licensee MDPI, Basel, Switzerland.
electron and 200 cm2 V−1 s−1 for hole, which are roughly four to six orders of magnitude
This article is an open access article
smaller than graphene and even much lower than those of low-doped Si (1350 cm2 V−1 s−1
distributed under the terms and
for electron and 480 cm2 V−1 s−1 for hole). Beyond TMDCs, a large 2D family of transi-
conditions of the Creative Commons
Attribution (CC BY) license (https://
tion metal carbides and carbonitrides, called MXenes, has been discovered in recent
creativecommons.org/licenses/by/
years [15–17]. MXenes are produced by the etching out of the A layers from MAX phases
4.0/).
of Mn+1 AXn , where M is a transition metal, A is mainly a group IIIA or IVA element, and

Nanomaterials 2021, 11, 559. https://doi.org/10.3390/nano11030559 https://www.mdpi.com/journal/nanomaterials


Nanomaterials 2021, 11, 559 2 of 14

X is C or N with n = 1, 2, 3. So far, more than 60 different pure MXenes have been explored.
Their electronic properties, such as band-gap and magnetism, can be tuned by changing
the MXene elemental composition and the surface terminations.
Very recently, a new kind of hexagonal 2D MXene, MoSi2 N4 and WSi2 N4 , was
successfully synthesized by chemical vapor deposition method with large size up to 15
mm × 15 mm [18]. They show good environmental stability, and even have no structural
deformation within six months. Monolayer MoSi2 N4 is septuple-atomic-layer structure of
N-Si-N-Mo-N-Si-N, which can be view as a MoN2 layer sandwiched by two SiN layers.
It exhibits indirect bandgap semiconducting behavior with band-gap roughly equal to
1.94 eV. The elastic modulus is four times that of monolayer MoS2 , and electron/hole
mobility is also roughly four-to-six times larger than that of monolayer MoS2 . In addition,
a large family of MA2 Z4 is predicted by first-principles calculation, where M represents the
elements of IVB, VB, or VIB groups, A represents Si or Ge, and Z represents the elements
of VA group. The nanosheets in this family are expected to have wide tunable bandgap
and magnetic properties, meaning potential application in electronics, optoelectronics and
spintronics.
Motivated by the exciting properties of MoSi2 N4 and WSi2 N4 , some theoretical works
have been carried out to further explore the mechanical and physical properties of their
family by using the first-principles calculation method [19,20]. The lattice thermal conduc-
tivity, piezoelectric and flexoelectric response, and photocatalytic and electronic feature of
monolayer MA2 Z4 (M = Cr, Mo, W; A = Si, Ge; Z = N, P) were systematically calculated.
They show diverse electronic properties from antiferromagnetic metal to half metal and
semiconductor with band gaps ranging from 0.31 to 2.57 eV. Monolayer MoSi2 N4 and
WSi2 N4 were predicted to show outstandingly high lattice thermal conductivity of 440
and 500 W/mK, respectively [19]. The piezoelectricity property was calculated for six
different configurations of MSi2 N4 (M = Mo, W) which are built through translation, mirror
and rotation operations. The maximum piezoelectric strain and stress coefficients is 3.53
pm/V and 13.95 × 10−10 C/m for MoSi2 N4 , and 2.91 pm/V and 12.17 × 10−10 C/m for
WSi2 N4 , respectively, which are much larger than those of 2D TMD, metal oxides, III-V
semiconductor and Janus TMD [20]. By tuning biaxial in-plane strain to monolayer VSi2 P4 ,
a continuous phase transition can be occurred from a ferromagnetic metal to a spin-gapless
semiconductor to a ferromagnetic semiconductor to spin-gapless semiconductor to a fer-
romagnetic half-metal. At the ferromagnetic semiconductor phase, ferromagnetism and
piezoelectricity can exist together due to broken inversion symmetry [21]. The van der
Waals hetero-structures composed of MoSi2 N4 contacted by graphene and NbS2 monolay-
ers were predicted to exhibit ultra-low Schottky barrier height, which can be modulated
via the interlayer distance or external electric field [22]. Due to the intrinsic inversion sym-
metry breaking and strong spin–orbital coupling, remarkable spin-valley coupling in the
0
inequivalent valleys at K and K points can be found for MoSi2 X4 (X = N, P, As). It induces
spin-valley coupled optical selection properties, which can be tuned by in-plane strain [23].
Beyond traditional two-level valleys, monolayer MoSi2 N4 shows multiple folded valleys,
implying an additional intrinsic degree of freedom. The valley-contrasting properties in
monolayer MoSi2 N4 were discussed by using a three-band low-power Hamiltonian, where
each valley and energy band can be selectively controlled [24].
In this paper, we systematically investigate the structural, electronic, optoelectronic
and quantum transport properties of monolayer and bilayer MoSi2 Z4 (Z = P, As). All these
2D materials possess stable configuration, moderate direct band-gap, high and anisotropic
carrier mobilities, large optical absorption coefficient, superior photon responsivity and
external quantum efficiency in the visible light region. An optoelectronic device based on
monolayer MoSi2 Z4 is also built to model the adjustable photocurrent. Our investigation
further expands the application prospect of few-layer MoSi2 Z4 in nanoelectronics and
optoelectronics.
The rest of this paper is organized as follows. In Section 2, the computational methods
are briefly introduced. In Section 3, the numerical results of the structural, electronic
Nanomaterials 2021, 11, 559 3 of 14

and optoelectronic properties are presented. In addition, the photocurrent of monolayer


MoSi2 Z4 -based nanodevice is also calculated. In Section 4, a brief summary is presented.

2. Numerical Methods
A first-principles calculation is performed by using the Vienna ab initio simulation
package [25,26] based on the density functional theory (DFT). Both the generalized gradient
approximation with a PBE form [27] and the Heyd-Scuseria-Ernzerhof (HSE06) [28] hybrid
functional is adopted to calculate the band structures and optical-electronic properties. The
energy cutoff and reciprocal k-points are chosen as 500 eV and 16 × 16 × 1 in structure
relaxation and electronic calculation. A vacuum space of 20 Å perpendicular to the 2D
plane is applied to separate the periodic images. The weak vdW interaction between
adjacent layers is described by the DFT-D2 functional with Grimme correction [29]. The
convergence criteria of force and energy are set to 0.01 eV/Å and 10−5 eV. To examine the
stability of all the structures, PHONOPY code is used to calculate the phonon dispersion
curves [30], and ab initio molecular dynamics (AIMD) simulation [31] is carried out to
examine the total energy evolution at high temperature. To calculate the photocurrent of
2D layered MoSi2 Z4 based nanodevice, Nanodcal software is evaluated which is developed
based on the combination of DFT and non-equilibrium Green’s function (NEGF-DFT) [32].
In the calculation, norm-conserving pseudopotential, double-zeta polarization basis set
and exchange-correlation functional at PBE level are employed.

3. Results and Discussion


3.1. Structural and Mechanical Properties of Few-Layer MoSi2 Z4 (Z = P, As)
Figure 1 shows the optimized schematic structures of monolayer (a) and bilayer
(b–d) MoSi2 Z4 from top view and side view, where Z = P, As. Monolayer MoSi2 Z4 is
constructed from septuple atomic layers of Z–Si–Z–Mo–Z–Si–Z, which can be viewed as
a MoZ2 layer sandwiched by two SiZ layers. It presents A–B stacked hexagonal lattice
from the top view, and its primitive cell includes one Mo atom, two Si atoms and four
Z atoms as labeled by the parallelogram in Figure 1a. The lattice parameters a = b =
3.470 Å and 3.620 Å for Z = P and As, respectively, which are well coincident with those
predicted in previous work [18]. Figure 1b–d present three most likely stacking patterns
of bilayer MoSi2 Z4 , namely AA, AB and AC, where the Si atoms in the lower layer are
aligned with the Si, Z, and Mo atoms in the upper layer, respectively. The relaxed lattice
parameters a and interlayer distances d are listed in Table 1 for each stacking pattern and
two kinds of Z atoms. We find that the interlayer distance of AB stacking is the smallest
compared to the other two stacking patterns for both MoSi2 P4 and MoSi2 As4 .

Figure 1. (a) Schematic structure of monolayer MoSi2 Z4 (Z = P, As) from top view and side view. The
parallelogram indicates its primitive cell. (b) AA, (c) AB, (d) AC stacking patterns of bilayer MoSi2 Z4 .
Nanomaterials 2021, 11, 559 4 of 14

Table 1. Lattice constants a, interlayer distance d, cohesive energy Ec , binding energy Eb and band gap of few-layer
MoSi2 Z4 (Z = P, As).

System Monolayer Bilayer MoSi2 P4 Monolayer Bilayer MoSi2 As4


Patterns MoSi2 P4 AA AB AC MoSi2 As4 AA AB AC
a( Å) 3.470 3.449 3.450 3.450 3.620 3.581 3.583 3.583
d( Å) — 3.850 3.075 3.081 — 3.825 3.108 3.112
Ec (eV ) −6.089 — — — −5.475 — — —
Eb (eV ) — −3.536 −3.614 −3.613 — −4.272 −4.385 −4.384
Bandgap(eV) 1.015 0.994 1.019 1.021 0.891 0.888 0.894 0.894

Firstly, we check the stability of monolayer and bilayer MoSi2 Z4 before further study-
ing their physical properties. For monolayer MoSi2 Z4 , the cohesive energy is calculated
by
Ec = ( E Mo + 2ESi + 4EZ − E MoSi2 Z4 )/7, (1)
where E Mo , ESi , EZ and E MoSi2 Z4 are total energies of isolated Mo atom, Si atom, Z atom
and a primitive cell of MoSi2 Z4 . The calculated cohesive energies are 6.089 eV/atom
for MoSi2 P4 and 5.475 eV/atom for MoSi2 As4 . They are smaller than that of graphene
(7.46 eV/atom), while larger than those of MoS2 (4.98 eV/atom) and phosphorene (3.30
eV/atom) [33–35] indicting proper stability. For bilayer MoSi2 Z4 , the stability is generally
measured by the binding energy defined as

Eb = EBL − 2E ML , (2)

where EBL and E ML stand for total energies of bilayer and monolayer MoSi2 Z4 , respectively.
As listed in Table 1, the binding energies are negative for all the bilayer MoSi2 Z4 , and the
AB stacking has the smallest value indicting the most stable stacking patten. Thus, we only
focus on the AB stacking pattern for the bilayer MoSi2 Z4 in the rest of this paper.
Next, the phonon dispersion spectrums of monolayer MoSi2 Z4 are calculated to
examine their dynamic stability. Figure 2a presents the phonon dispersion spectrum of
monolayer MoSi2 P4 . The low-frequency band near Γ point is roughly linear and there is no
imaginary modes in the Brillouin zone, which indicates monolayer MoSi2 P4 is dynamically
stable. An AIMD simulation is performed at 300 K to further examine the thermal stability
of the structure by employing a 4 × 4 supercell. As shown in Figure 2b, the total energy of
monolayer MoSi2 P4 oscillates slightly in the vicinity of −720 eV for a long time without
decay. Neither bond-breaking nor geometry reconstruction appears in the structure at
10 fs indicating thermal stability of monolayer MoSi2 P4 at room temperature. Similar
phonon dispersion spectrums and total energy evaluations are also obtained for all the
other monolayer and bilayer structures, as shown in Figure 3.
Finally, we examine the mechanical properties of all the structures under external force
2 >0
by calculating elastic constants Cij . As listed in Table 2, the Born criteria C11 C22 − C12
and C66 > 0 are both satisfied for the monolayer and bilayer MoSi2 Z4 meaning their
mechanical stability [36]. Based on Ci,j , Young’s modulus Y (θ ) and the Poisson’s ratio ν(θ )
along the in-plane angle θ and the layer modulus γ are also calculated. Y (θ ) indicates the
reciprocal of the response of strain to stress along a specific direction along θ in the 2D
plane. ν(θ ) is the ratio of the absolute value of transverse normal strain to axial normal
strain. γ represents the resistance of the 2D surface to stretching, and thus is independent
of θ. These physical quantities can be calculated by the following formulas [37]
2
C11 C22 − C12
Y (θ ) = ,
C11 sin4 θ + A sin2 θ cos2 θ + C22 cos4 θ

C12 sin4 θ − B sin2 θ cos2 θ + C12 cos4 θ


ν(θ ) = ,
C11 sin4 θ + A sin2 θ cos2 θ + C22 cos4 θ
Nanomaterials 2021, 11, 559 5 of 14

1
γ= (C + C22 + 2C12 ),
4 11
in which A = (C11 C22 − C12 2 ) /C − 2C 2
66 12 and B = C11 + C22 − (C11 C22 − C12 ) /C66 .
Figure 2c,d show the Y (θ ) and ν(θ ) of monolayer MoSi2 Z4 . Y (θ ) is isotropic and ν(θ ) is
roughly isotropic for both monolayers. Y (θ ) of MoSi2 P4 is larger than that of MoSi2 As4 ,
while ν(θ ) of the former is smaller to that of the latter. This means monolayer MoSi2 As4 is
easier to deform under in plane external force than monolayer MoSi2 P4 . It is reasonable
because the As–Mo and As–Si bonds are longer and deformable than the P–Mo and P–Si
bonds. Similarly, γ of MoSi2 P4 is larger than that of MoSi2 As4 . Y and γ of monolayer
MoSi2 Z4 are slightly smaller than that of monolayer graphene (340 N/m and 215.9 N/m)
and BN (318 N/m and 177.0 N/m) [38], while comparable to those of SiC (179.7 N/m
and 116.5 N/m) [38] and monolayer PC3 (180.4 N/m and 102.1 N/m) [39] implying their
similar mechanical response. In terms of bilayer MoSi2 Z4 , both Y (θ ) and γ are nearly two
times as those of monolayer MoSi2 Z4 (see Table 2 and Figure 3). Such behavior is physically
reasonable and in good accordance with that of multilayer graphene [38] and PC3 [39]. The
calculated moduli indicate that few-layered MoSi2 Z4 are stretchable and flexible as most
of the other common 2D materials, indicting potential application in flexible electronic
devices.

Figure 2. (a) Phonon dispersion curves and (b) total energy variation at 300 K of monolayer MoSi2 P4 .
Inset in (b) shows the top view and side view of a snapshot at 10ps. (c) Young’s modulus Y (θ ) and (d)
Poisson’s ratio ν(θ ) of monolayer MoSi2 P4 (purple curve) and MoSi2 As4 (blue curve) along arbitrary
in-plane directions.
Nanomaterials 2021, 11, 559 6 of 14

Figure 3. (a) Phonon band dispersion curves of the monolayer MoSi2 As4 . Variations of total energy at
300 K as functions of time for (b) monolayer MoSi2 As4 , (c) bilayer MoSi2 P4 , (d) bilayer MoSi2 As4 . (e)
Young’s modulus Y (θ ) and (f) Poisson’s ratio ν(θ ) of bilayer MoSi2 P4 (purple curve) and MoSi2 As4
(blue curve) along arbitrary in-plane directions.

Table 2. The calculated elastic constants Cij , Young’s modulus Y and Poisson’s ratio ν along the x (θ = 0) and y (θ = π/2)
directions, layer modulus γ for monolayer (ML-) and bilayer (BL-) MoSi2 Z4 (Z = P, As).

Type C11 ( N/m) C22 ( N/m) C12 ( N/m) C66 ( N/m) Yx ( N/m) Yy ( N/m) νx νy γ( N/m)
ML-
217.70 222.65 56.35 80.67 203.43 208.06 0.253 0.259 138.26
MoSi2 P4
BL-
476.59 479.76 130.42 173.09 441.14 444.07 0.272 0.274 304.30
MoSi2 P4
ML-
182.38 188.67 52.01 65.18 168.04 173.84 0.276 0.285 118.77
MoSi2 As4
BL-
415.86 423.43 124.00 145.93 379.54 386.45 0.293 0.298 271.82
MoSi2 As4
Nanomaterials 2021, 11, 559 7 of 14

3.2. Electronic Properties of Few-Layer MoSi2 Z4


Figure 4a,b show the band structure and projected density of states (PDOS) of mono-
layer MoSi2 P4 and MoSi2 As4 based on PBE and HSE06 exchange-correlation functionals.
For each configuration, the band structure based on the PBE exchange-correlation func-
tional is similar to that based on the HSE06 functional except the smaller bandgap. Both
structures show a direct bandgap, and both conduction band minimum (CBM) and valence
band maximum (VBM) locate at K point. This is different from monolayer MoSi2 N4 , whose
CBM sits K point while VBM locates at Γ point, presenting indirect band-gap semicon-
ducting behavior (Ref. [18], also see Figure 5a). To get more insight into this difference,
PDOS and charge distribution at VBM and CBM are plotted in Figure 5c,d. For monolayer
MoSi2 P4 and MoSi2 As, both CBMs and VBMs are mainly originated from the d orbitals of
Mo atoms which locate in the middle layer of the structures. While, for monolayer MoSi2 N4 ,
VBM is dominated by both d orbital of Mo atoms and p orbital of Z atom (see Figure 5b).
Due to the orbital hybridization, an obvious extension of VBM from the middle Mo atoms
to beside the Z atoms occurs, which is much different from the charge distribution of VBMs
in MoSi2 P4 and MoSi2 As4 . This is reasonable because the N–Mo bonds are shorter than the
P–Mo and As–Mo bonds, and thus the orbital hybridization is more likely to happen in
MoSi2 N4 .

Figure 4. The electronic band structure and projected density of states (PDOS) of monolayer (a) MoSi2 P4
and (b) MoSi2 As4 . (c–d): Corresponding charge distribution at valence band maximum (VBM) and
conduction band minimum (CBM) dominated by the d orbital of Mo atoms in the middle layer.

In terms of bilayer MoSi2 Z4 , similar direct bandgap semiconducting behavior


to monolayer MoSi2 Z4 are obtained, where both CBM and VBM locate at K points
(see Figure 5c,d). The bandgap of bilayer MoSi2 Z4 changes very little in comparison to
that of monolayer MoSi2 Z4 (see Table 1). The independence of bandgap with the number
of layers can be attributed to the orbital shield. Because the states at CBM and VBM are
dominated by the d orbital of Mo atoms, they are effectively screened inside the monolayer
MoSi2 Z4 because the Mo atoms located in the middle layer of seven atomic layers. For
bilayer MoSi2 Z4 , the rather weak interlayer vdW interaction makes the Mo atoms at the up
layer and those at the down layer have nothing to do with each other, and thus the band
gap is very close to that of the monolayer. Similar layer number independent bandgap
behavior has also been found in layered 2D KAgSe [40]. The layer number independent
electronic properties provide enormous convenience and less difficulty in experimental
fabrication of finite layer MoSi2 Z4 -based electronic devices.
Nanomaterials 2021, 11, 559 8 of 14

Figure 5. (a) The electronic band structure and PDOS of monolayer MoSi2 N4 , (b) Corresponding
charge distribution at VBM and CBM of monolayer MoSi2 N4 . The electronic band structure and
PDOS of bilayer (c) MoSi2 P4 and (d) MoSi2 As4 .

3.3. Carrier Mobilities of Few-Layer MoSi2 Z4


Carrier mobility is an important factor to describe the transport ability of electronic
and optoelectronic materials, which can be evaluated by using the deformation potential
method as follows [41,42],
eh̄C
µ= 2
, (3)
k B Tm∗ md EDP

where T is the temperature and equal to 300 K in this calculation; m∗ = ±h̄2 (d2 Ek /dk2 ) is
the effective mass of electrons and holes depending on the change of energy with wave
vector k along
q different transport directions; md is the averaged effective mass defined
as md = m∗x m∗y ; C = (∂2 E/∂2 ε)/S0 is the elastic modulus related to the change of
total energy with strain along different directions; EDP = dEedge /dε is the deformation
potential constant given by the change rate of band edges with strain. The calculated carrier
mobilities and corresponding parameters of layered MoSi2 Z4 are summarized in Table 3.
Three pieces of information can be obtained from Table 3. Firstly, the carrier mobility
of holes is roughly three to four times larger than that of electrons for both monolayer
and bilayer MoSi2 Z4 along with both x and y directions, which mainly attributes to the
smaller deformation potential constant EDP of holes. This difference of carrier mobilities
can effectively facilitate the spatial separation of electrons and holes, which reduces the
recombination probability of photo-excited carriers and suggests satisfactory performances
for nanoscale electronic and optoelectronic devices. Secondly, the carrier mobilities of
bilayer MoSi2 Z4 are largely improved in contrast to those of monolayer MoSi2 Z4 due to
the roughly doubled elastic modulus C. Similar properties were also found for MXs [43].
Thirdly, the carrier mobilities of MoSi2 P4 are slightly higher than that of MoSi2 As4 for
both monolayer and bilayer structures, which are also independent of carrier types and
directions. Especially, these carrier mobilities are relatively high, which are much larger
than those of MoS2 (200–500 cm2 V−1 s−1 ) [44] and even comparable to those of black
phosphorene [42] indicating potential application in 2D electronic devices.
Nanomaterials 2021, 11, 559 9 of 14

Table 3. The effective mass m∗ , elastic modulus C2D , deformation potential constant EDP , and carrier mobility µ along x
and y directions for monolayer and bilayer MoSi2 Z4 at 300 K.

System Carrier Type m∗ /m0 C2D ( Nm−1 ) E DP (eV ) µ(cm2 V −1 s−1 )


e (x) 0.325 214.88 6.82 828.76
ML- e (y) 0.415 218.74 6.28 778.90
MoSi2 P4 h (x) 0.339 214.88 3.43 3171.83
h (y) 0.430 218.74 3.65 2131.78
e (x) 0.313 481.13 6.94 1919.84
BL- e (y) 0.403 484.88 6.40 1759.76
MoSi2 P4 h (x) 0.344 481.13 2.99 8652.25
h (y) 0.435 481.88 3.55 4860.34
e (x) 0.499 178.40 4.05 823.19
ML- e (y) 0.640 178.37 3.76 743.11
MoSi2 As4 h (x) 0.419 178.40 3.04 2093.38
h (y) 0.524 178.37 3.16 1552.98
e (x) 0.496 432.14 4.19 1855.52
BL- e (y) 0.659 432.12 3.88 1629.69
MoSi2 As4 h (x) 0.425 432.14 2.79 5905.10
h (y) 0.528 432.12 2.77 4819.97

3.4. Optical Absorption Spectrums of Layered MoSi2 Z4


Monolayer and bilayer MoSi2 Z4 with direct band gaps about 0.85–1.0 eV exhibit
potential application for visible–light solar harvesting/utilizing techniques or making
narrow-gap semiconductor devices. Recent studies revealed that such narrow band
gap materials are good candidates of infrared photodetectors, such as phosphorus
carbides and black arsenic phosphorus [6,45,46]. Thus, we further investigate the opto-
electronic performance of few-layer MoSi2 Z4 by calculating the absorption coefficient
as follows [47,48]

rq
α(ω ) = 2ω ε21 (ω ) + ε22 (ω ) − ε 1 (ω ), (4)

where c, ω, ε 1 (ω ) and ε 2 (ω ) stands for the light velocity, frequency of incident light, real
part and imaginary part of the frequency-dependent dielectric function, respectively. ε 1 (ω )
and ε 2 (ω ) can be calculated by using the Kramers–Kronig relation and summing all the
empty states in the Brillouin zone.
Figure 6 shows the optical absorption coefficients of monolayer and bilayer MoSi2 Z4
based on PBE and HSE06 calculations, where the polarization direction of incident light is
parallel to the 2D plane. All the few-layer MoSi2 Z4 display very similar and remarkably
high absorption coefficients (∼105 cm−1 ) in the visible-ultraviolet light region, which
agrees well with their similar band gaps as shown in Figure 4. The large absorption is even
comparable to that of graphene, phosphorene and MoS2 [6]. The strong optical absorption
and broad absorption ranges make layered MoSi2 Z4 promising materials for photovoltaic
solar cells and optoelectronic devices. Especially, the layer number independence to the
bandgap and optical absorption makes the experimental fabrication more convenient of
few-layer MoSi2 Z4 -based 2D optoelectronic devices.

3.5. Photocurrent in Monolayer MoSi2 Z4 Nanodevice


On account of the similar and excellent optical absorption performance of layered
MoSi2 Z4 , we build a monolayer MoSi2 Z4 -based two-probe 2D optoelectronic device as
shown in Figure 7 and evaluate its photoinduced current. To solve the quantum transport
problem in this identical system, the device can be separated into three parts theoretically
including a central scattering region and two semi-infinite electrodes. When the incident
light energy in the scattering region is larger than the bandgap, the electrons at the valence
band can be excited to the conduction band by absorbing photons. When a tiny external
bias is applied between the source and the drain, the excited electrons can be driven to
Nanomaterials 2021, 11, 559 10 of 14

produce photocurrent in the system. Note that the potential difference between the left
and the right leads should be much smaller than the bandgap of the system to ensure
that the detected current in the electrode is completely generated by the light but not bias.
The photocurrent flowing into the left probe can be expressed in terms of the NEGF as
follows [49–51],

ie
Z h i
ph
JL = Tr Γ L { G <( ph) + f L ( E)( G >( ph) − G <( ph) )} dE, (5)
h

where f L , Γ L and G </>( ph) denotes the Fermi distribution function, line-width function
and greater/lesser Green’s function of the two-probe system including electron-photon
interaction, respectively.

Figure 6. Optical absorption coefficients versus the energy of incident light for 2D (a) monolayer
MoSi2 P4 , (b) monolayer MoSi2 As4 , (c) bilayer MoSi2 P4 and (d) bilayer MoSi2 As4 based on PBE and
HSE06 functionals. For each panel, the polarization vector of incident light is set parallel to the plane,
and the two vertical dashed lines indicate the region of visible light.

Figure 7. Schematic structure of monolayer MoSi2 Z4 -based 2D optoelectronic device. The yellow
zone in the center scattering region stands for the lighting area. The left blue region and right red
region represent the drain and source, respectively.

In this calculation, the incident light is perpendicular to the 2D plane, and the angle
between polarization direction and transport direction is labeled as θ. Figure 8a,c show
the photocurrent versus energy of the linearly polarized light with power density equal to
103 µW/mm2 and θ equal to 0◦ and 90◦ . When energy is smaller than 0.5 eV, photocurrent
is equal to zero for both MoSi2 P4 and MoSi2 As4 because the energy is smaller than their
Nanomaterials 2021, 11, 559 11 of 14

band gaps. With further increase of energy, photocurrent appears and oscillates with the
energy for both θ = 0◦ and 90◦ depending on the detailed behavior of band structures.
Photocurrent reaches local maximum in the visible region at θ = 0◦ , and in the ultraviolet
region at θ = 90◦ . To further explore the influence of incident polarization angle θ,
photocurrent as a function of θ under different photon energy in the visible light region
are evaluated as shown in Figure 6b,d. For both MoSi2 P4 and MoSi2 As4 , the photocurrent
is roughly symmetrical with respect to θ = 90◦ , and reaches maximums at θ = 0◦ and
θ = 180◦ . Similar symmetrical distribution of photocurrent with polarization angle was
also reported for monolayer KAgSe-based 2D optoelectronic device [40].
The responsivity R ph and external quantum efficiency τeqe are generally used to
measure the photovoltaic performances, which are defined as
ph
JL
R ph = (6)
eFph

and
hc
τeqe = R ph , (7)

in which the photon flux Fph stands for the number of incident photons in unit area and
unit time. R ph of monolayer MoSi2 P4 and MoSi2 As4 in the visible light region are 0.060
AW −1 and 0.046 AW −1 , respectively, which are the same order as those of MoS2 (0.016
AW −1 ) and monolayer chalcogenides (0.035 AW −1 for GeS and 0.075 AW −1 for SnS), while
two orders higher than that of graphene (5 × 10−4 AW −1 ) [52]. τeqe of monolayer MoSi2 P4
and MoSi2 As4 in the visible light region can reach 18.60% and 13.33%, respectively, which
are comparable to those of KAgSe (17.92%) [40] and monolayer chalcogenides (10.27% for
GeS and 22.01% for SnS) [53]. In addition, R ph and τeqe of monolayer MoSi2 Z4 are greatly
increased within the whole light region, ie, 0.143 AW −1 and 64.26% for MoSi2 P4 , 0.098
AW −1 and 41.16% for MoSi2 As4 . Here, it is worth mentioning that the above values of
photon responsivity R ph and external quantum efficiency τeqe are all calculated theoretically
based on the computational models. It is hoping that there will be more experimental
results to support in the future. Once again, these ideal performances of MoSi2 Z4 suggest
their powerful potential application in optoelectronic and photovoltaic devices.

Figure 8. Photocurrent as a function of photon energy with polarization angle θ of the incident
light equal to 0◦ and 90◦ for monolayer (a) MoSi2 P4 and (c) MoSi2 As4 nanodevice. (c,d) show the
photocurrent versus photon energy and θ for monolayer (b) MoSi2 P4 and (d) MoSi2 As4 nanodevice.
Nanomaterials 2021, 11, 559 12 of 14

4. Conclusions
Recently synthesized 2D semiconductors MoSi2 N4 and WSi2 N4 exhibit prominent
material and physical properties, including remarkable stability, high strength and large
carrier mobility, which also inspires increasing theoretical researches to further explore the
physical properties of their family MA2 Z4 . First principle calculations indicate that MA2 Z4
materials possess wide tunable band gaps, magnetic properties and valley-contrasting
properties, indicating potential applications in electronics, optoelectronics, spintronics and
valleytronics. In this case, we investigated the electronic and photoelectrical properties
of monolayer and bilayer 2D MoSi2 Z4 (Z = P, As) by using the first-principles calculation
method. Firstly, the structural, dynamic, thermal and mechanical stabilities of the few-layer
MoSi2 Z4 were numerically verified. Secondly, both monolayer and bilayer MoSi2 Z4 show
direct bandgap semiconducting behavior, which is different from MoSi2 N4 with indirect
bandgap. Moreover, the band gaps of layered MoSi2 Z4 are roughly independent of the
number of layers due to effective screening to the atomic orbital of Mo atoms. Thirdly,
monolayer and bilayer MoSi2 Z4 show high carrier mobilities and remarkable optical
absorption coefficients. Monolayer MoSi2 Z4 -based optoelectronic device displays large
photon responsivity and external quantum efficiency. All these appealing properties make
MoSi2 Z4 promising candidates for application in electronic and optoelectronic devices.

Author Contributions: conceptualization, H.Y. and B.W.; methodology, C.Z. and Q.W.; software,
H.Y., J.L., F.X. and Y.W.; formal analysis, H.Y., B.W. and Y.Y.; investigation, H.Y. and C.Z.; writing-
original draft preparation, H.Y.; writing-review and editing, B.W.; All authors have read and agreed
to the published version of the manuscript.
Funding: This work was financially supported by grants from the National Natural Science Foun-
dation of China (NSFC) (Grant No. 11774238), Shenzhen Natural Science Foundations (Grant No.
JCYJ20190808150409413, JCYJ20190808115415679 and JCYJ20190808152801642) and the Natural Sci-
ence Foundation of Guangdong Province (GDNSF) (Grant No. 2020A1515011418).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: There are no conflicts of interest to declare.

References
1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in
atomically thin carbon films. Science 2004, 306, 666–669.
2. Ci, L.; Song, L.; Jin, C.; Jariwala, D.; Wu, D.; Li, Y.; Srivastava, A.; Wang, Z.F.; Storr, K.; Balicas, L.; Liu, F.; Ajayan, P.M. Atomic
Layers of Hybridized Boron Nitride and Graphene Domains. Nat. Mater 2010, 9, 430.
3. Wang, Q.H.; Kalantarzadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition
metal dichalcogenides. Nat. Nanotechnol. 2017, 7, 699–712.
4. Hu, Z.Y.; Ding, Y.C.; Hu, X.M.; Zhou, W.H.; Yu, X.C.; Zhang, S.L. Recent progress in 2D group IV-IV monochalcogenides:
synthesis, properties and applications. Nanotechnology 2019, 30, 252001.
5. Xu, K.; Yin, L.; Huang, Y.; Shifa, T.A.; Chu, J.W; Wang, F.; Cheng, R.Q.; Wang, Z.X.; He, J. Synthesis, properties and applications of
2D layered M I I I XV I (M= Ga, In; X=S, Se, Te) materials. Nanoscale 2016, 8, 16802–16818.
6. Yu, T.; Zhao, Z.Y.; Sun, Y.H.; Bergara, A.; Lin, J.Y.; Zhang, S.T.; Xu, H.Y.; Zhang, L.J.; Yang, G.C.; Liu, Y.C. Two-dimensional PC6
with direct band gap and anisotropic carrier mobility. J. Am. Chem. Soc. 2019, 141, 1599–1605.
7. Niedzielski, B.; Jia, C.; Berakdar, J. Supercurrent Induced by Chiral Coupling in Multiferroic/Superconductor Nanostructures.
Nanomaterials 2021, 11, 184.
8. Chittari, B.L.; Lee, D.; Banerje,e N.; MacDonald, A.H.; Hwang, E.; Jung, J. Carrier- and strain-tunable intrinsic magnetism in
two-dimensional MAX3 transition metal chalcogenides. Phys. Rev. B 2020, 101, 085415.
9. Hou, W.; Mi, H.; Peng, R.; Peng, S.; Zeng, W.; Zhou, Q. First-Principle Insight into Ga-Doped MoS2 for Sensing SO2 , SOF2 and
SO2 F2 . Nanomaterials 2021, 11, 314.
10. Chen, J.; Xi, J.; Wang, D.; Shuai, Z. Carrier Mobility in Graphyne Should Be Even Larger than That in Graphene: A Theoretical
Prediction. J. Phys. Chem. Lett. 2013, 4, 1443.
Nanomaterials 2021, 11, 559 13 of 14

11. Lherbier, A.; Botello-Mendez, A.R.; Charlier, J.C. Electronic and Transport Properties of Unbalanced Sublattice N-Doping in
Graphene. Nano Lett. 2013, 13, 1446.
12. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2 : A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010,
105, 136805.
13. Ahmadi, M.; Zabihi, O.; Jeon, S.; Yoonessi, M.; Dasari, A.; Ramakrishna, S.; Naebe, M. 2D transition metal dichalcogenide
nanomaterials: advances, opportunities, and challenges in multi-functional polymer nanocomposites. J. Mater. Chem. A 2020,
8, 845–883.
14. Cai, Y.Q.; Zhang, G.; Zhang, Y.W. Polarity-Reversed Robust Carrier Mobility in Monolayer MoS2 Nanoribbons. J. Am. Chem. Soc.
2014, 136, 6269–6275.
15. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D metal carbides and nitrides (MXenes) for energy storage, Nat. Rev. Mater. 2017,
2, 16098.
16. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th anniversary article: MXenes: A new family of two-dimensional
materials. Adv. Mater. 2014, 26, 992–1005.
17. Huang, K.; Li, Z.; Lin, J.; Han, G.; Huang, P. Two-dimensional transition metal carbides and nitrides (MXenes) for biomedical
applications. Chem. Soc. Rev. 2018, 47, 5109–5124.
18. Hong, Y.L.; Liu, Z.; Wang, L.; Zhou, T.; Ma, W.; Xu, C.; Feng, S.; Chen, L.; Chen, M.L.; Sun, D.M.; et al. Chemical vapor deposition
of layered two-dimensional MoSi2 N4 materials. Science 2020, 369, 670–674.
19. Mortazavi, B.; Javvaji, B.; Shojaei, F.; Rabczuk, T Shapeev, A.V.; Zhuang, X.Y. Exceptional piezoelectricity, high thermal conductiv-
ity and stiffness and promising photocatalysis in two-dimensional MoSi2 N4 family confirmed by first-principles. Nano Energy
2021, 82, 105716.
20. Guo, S.D.; Zhu, Y.T.; Mu, W.Q.; Wang, L.; Chen, X.Q. Structure effect on intrinsic piezoelectricity in septuple-atomic-layer
MoSi2 N4 (M = Mo and W). Comput. Mater. Sci. 2021, 188, 110223.
21. Guo, S.D.; Mu, W.Q.; Zhu, Y.T.; Wang, L.; Chen, XQ Shapeev, A.V.; Zhuang, X.Y. Coexistence of intrinsic piezoelectricity and
ferromagnetism induced by small biaxial strain in septuple-atomic-layer VSi2 P4 . Phys. Chem. Chem. Phys. 2020, 22, 28359–28364.
22. Cao, L.M.; Zhou, G.H.; Wang, Q.Q.; Ang, L.K.; Ang, Y.S. Two-dimensional van der Waals electrical contact to monolayer MoSi2 N4 .
Appl. Phys. Lett. 2021, 118, 013106.
23. Ai, H.Q.; Liu, D.; Geng, J.Z.; Wang, S.P.; Lo K.H.; Pan, H. Theoretical evidence of the spin-valley coupling and valley polarization
in two-dimensional MoSi2 X4 (X = N, P, and As). Phys. Chem. Chem. Phys. 2021, 23, 3144–3151.
24. Mortazavi, B.; Javvaji, B.; Shojaei, F.; Rabczuk, T Shapeev, A.V.; Zhuang, X.Y. Valley pseudospin in monolayer MoSi2 N4 and
MoSi2 As4 . Phys. Rev. B 2021, 103, 035308.
25. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953.
26. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys.
Rev. B 1996, 54, 11169.
27. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
28. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003,
118, 8207–8215.
29. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H.A. consistent and accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104.
30. Togo, A.; Tanaka, I. First Principles Phonon Calculations in Materials Science. Scr. Mater 2015, 108, 1–5.
31. Barnett, R.; Landman, U. Born-Oppenheimer Molecular-Dynamics Simulations of Finite Systems: Structure and Dynamics of
(H2 O)2 . Phys. Rev. B Condens. Matter Mater. Phys. 1993, 48, 2081–2097.
32. Taylor, J.; Guo, H.; Wang, J. Ab Initio Modeling of Quantum Transport Properties of Molecular Electronic Devices. Phys. Rev. B
Condens. Matter Mater. Phys. 2001, 63, 245407.
33. Guan, J.; Zhu, Z.; Tománek, D. Phase Coexistence and Metal-Insulator Transition in Few-Layer Phosphorene: A Computational
Study. Phys. Rev. Lett. 2014, 113, 46804.
34. Chen, P.J.; Jeng, H.T. High Applicability of Two-dimensional Phosphorous in Kagome Lattice Predicted from First-principles
Calculations, Sci. Rep. 2016, 6, 23151.
35. Ahmad, S.; Mukherjee, S. A Comparative Study of Electronic Properties of Bulk MoS2 and Its Monolayer using DFT Technique:
Application of Mechanical Strain on MoS2 Monolayer. Graphene 2014, 3, 52–59.
36. Lee, C.; Wei, X.; Kysar, J.W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science,
2008, 321, 385–388.
37. Michel, K.H.; Verberck, B. Theory of Elastic and Piezoelectric Effects in Two-dimensional Hexagonal Boron Nitride, Phys. Rev. B
Condens. Matter Mater. Phys. 2009, 80, 224301.
38. Andrew, R.C.; Mapasha, R.E.; Ukpong, A.M.; Chetty, N. Mechanical properties of graphene and boronitrene. Phys. Rev. B Condens.
Matter Mater. Phys. 2012, 85, 125428
39. Yao, H.; Wang, Q.; Li, J.W.; Cai, W.S.; Wei, Y.D.; Wang, B.; Wang, J. Two-dimensional few-layer PC3 as promising photocatalysts
for overall water splitting. Phys. Chem. Chem. Phys. 2020, 22, 1485–1492.
40. Wang, Q.; Li, J.W.; Liang, Y.; Nie, Y.N.; Wang, B. KAgSe: A New Two-Dimensional Efficient Photovoltaic Material with
Layer-Independent Behaviors. ACS Appl. Mater. Interfaces 2018, 10, 41670–41677.
Nanomaterials 2021, 11, 559 14 of 14

41. Xie, J.; Zhang, Z.Y.; Yang, D.Z.; Xue, D.S.; Si, M.S. Theoretical Prediction of Carrier Mobility in Few-Layer BC2 N. J. Phys. Chem.
Lett. 2014, 5, 4073–4077.
42. Qiao, J.; Kong, X.; Hu, Z.X.; Yang, F.; Ji, W. High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus.
Nat. Commun. 2014, 5, 4475.
43. Cui, Y.; Peng, L.; Sun, L.P.; Qian, Q.; Huang, Y.C. Two-dimensional few-layer group-III metal monochalcogenides as effective
photocatalysts for overall water splitting in the visible range. J. Mater. Chem. A 2018 6, 22768–22777.
44. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147–150.
45. Tan, W.C.; Huang, L.; Ng, R.J.; Wang, L.; Hasan, D.N.; Duffin, TJ Kumar, K.S.; Nijhuis, C.A.; Lee, C.; Ang, K.W. A Black
Phosphorus Carbide Infrared Phototransistor. Adv. Mater. 2018, 30, 1705039.
46. Long, M.; Gao, A.; Wang, P.; Xia, H.; Ott, C.; Pan, C.; Fu, Y Liu, E.; Chen, X.; Lu, W.; Nilges, T.; Xu, J.; Wang, X.; Hu, W.; Miao, F.
Room Temperature High-Detectivity Mid-Infrared Photodetectors Based on Black Arsenic Phosphorus. Sci. Adv. 2017, 3, 700589.
47. Kuzmenko, A.B. Kramers-Kronig. Constrained Variational Analysis of Optical Spectra. Rev. Sci. Instrum. 2005, 76, 083108.
48. Gajdoš, M.; Hummer, K.; Kresse, GFurthmüller, J.; Bechstedt, F. Linear Optical Properties in the Projector-Augmented Wave
Methodology. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 73, 045112.
49. Zhang, L.; Gong, K.; Chen, J.; Liu, L.; Zhu, Y.; Xiao, D.; Guo, H. Generation and Transport of Valley-polarized Current in
Transition-metal Dichalcogenides. Phys. Rev. B Condens. Matter Mater. Phys. 2014, 90, 195428.
50. Xie, Y.; Zhang, L.; Zhu, Y.; Liu, L.; Guo, H. Photogalvanic Effect in Monolayer Black Phosphorus. Nanotechnolo 2015, 26, 455202.
51. Henrickson, L.E. Nonequilibrium Photocurrent Modeling in Resonant Tunneling Photodetectors. J. Appl. Phys. 2002,
91, 6273–6281.
52. Pospischil, A.; Furchi, M.M.; Mueller, T. Solar-energy conversion and light emission in an atomic monolayer p-n diode. Nat.
Nanotechnol. 2014, 9, 257–261.
53. Zhao, P.; Yang, H.; Li, J.; Jin, H.; Wei, W.; LYu, Huang, B.; Dai, Y. Design of New Photovoltaic Systems Based on Two-dimensional
Group-IV Monochalcogenides for High Performance Solar Cells. J. Mater. Chem. A 2017, 5, 24145–24152.

You might also like