You are on page 1of 9

Superlattices and Microstructures 148 (2020) 106728

Contents lists available at ScienceDirect

Superlattices and Microstructures


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

Strain tunable ferroelectricity of SnSe/SnTe van der


Waals heterostructures
Hao Guo, Xiaobao Tian *, Haidong Fan, Wentao Jiang **
Department of Mechanics and Engineering, Sichuan University, Chengdu, Sichuan, 610065, China

A R T I C L E I N F O A B S T R A C T

Keywords: Two-dimensional (2D) van der Waals heterostructures (vdWHs) have attracted significant
IVA-monochalcogenides attention due to their potential applications in nanoscale electronic devices. Based on the first-
van der Waals heterostructures principles calculations, we design the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs with dynamical
Ferroelectricity
stability and thermal stability and investigate their mechanical and electronic properties. We find
Strain engineering
First-principles calculations
that the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs have better flexibility than other 2D vdWHs. The
SnSe/SnTe and SnSe/SnTe/SnSe vdWHs show reduced direct bandgap and indirect bandgap,
respectively. They can also realize the transition between direct and indirect bandgap, and even
exhibit the metallic property under external strains. Furthermore, the SnSe/SnTe and SnSe/SnTe/
SnSe vdWHs exhibit stable ferroelectric spontaneous polarizations with small potential barriers.
The ferroelectric spontaneous polarization of SnSe/SnTe vdWH increases rapidly as the biaxial
tensile strain (<1%) increases, and then decreases slightly. Under the action of 1% biaxial tensile
strain, the SnSe/SnTe vdWH has the largest ferroelectric spontaneous polarization 1.721 × 10− 10
C/m and a smaller potential barrier 7.593 meV. Our work confirms that IVA-monochalcogenides
vdWHs have tunable ferroelectric spontaneous polarization, which can provide new clues for the
study and application of novel 2D ferroelectric heterostructures.

1. Introduction

Ferroelectric materials exhibit spontaneous polarization that can be reversed by the application of an external electric field [1].
They have attracted significant attention due to their potential application for sensors, ferroelectric random access memories (FeR­
AMs) and dynamic random-access memory (DRAM) capacitors [2]. Given the demand for modern micro-nanoelectronics devices,
two-dimensional (2D) ferroelectric materials have attracted extensive research interest. However, the depolarization field of ferro­
electric materials gradually increases with the decreases of the thickness, which reduces their intrinsic ferroelectricity and limits their
application [3–5]. Previous studies have shown that the critical ferroelectric thickness of perovskite oxide is 20 Å or less [6–8]. Up to
now, many 2D ferroelectric materials, such as 1T-MoS2 [9], SbN and BiP monolayers [10], IVA- or IIIA-VA family [11],
IVA-monochalcogenides [12,13] and group-VA monolayers [14], have been reported. In particular, the IVA-monochalcogenides MXs
(M = Ge, Sn; X = S, Se, Te) have stable and robust ferroelectricity due to planar lattice distortion in their black phosphorus-like
wrinkled structures [15,16], which can also be adjusted by layers [17–19], strain [20–22] and charge doping [23]. In experiment,
the ferroelectricity has been discovered in monolayer SnTe with a Curie temperature close to room temperature [24]. Over the past two

* Corresponding author.
** Corresponding author.
E-mail addresses: xbtian@scu.edu.cn (X. Tian), scubme@aliyun.com (W. Jiang).

https://doi.org/10.1016/j.spmi.2020.106728
Received 15 May 2020; Received in revised form 19 August 2020; Accepted 19 October 2020
Available online 4 November 2020
0749-6036/© 2020 Elsevier Ltd. All rights reserved.
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

decades, different materials composed of chalcogen elements have also been widely studied and reported for a variety of applications,
which include the CdX (X = S, Se, Te) [25], GaOSe [26] and so on.
In recent years, the research hot spots have shifted to 2D vertical heterostructures, which can break the limitation of single 2D
material properties and are expected to lead the development of the next generation of nanoscale electronic devices [27–31]. The
vertical heterostructures are also known van der Waals heterostructures (vdWHs) because their layers are connected by vdW forces
[32–34]. Stacking 2D ferroelectric materials into 2D vdWHs not only preserves the properties of isolated compounds, but also dem­
onstrates many new properties, such as graphene/PbI2 [35], graphene/GaS [36], InGaAs/InP [37], and other 2D vdWHs [38,39].
There have also been some studies and reports of vdWHs based on IVA-monochalcogenides. For instance, the 2D GeSe/SnS2 (SnSe2)
vdWHs possess the broken-gap band alignment and can be adjusted by an external electric field [40]. The graphene/GeS vdWH shows
increased ferroelectric spontaneous polarization (186 μC/cm2), which can be adjusted by altering the interlayer spacing [41]. The
SnTe/Pt and SnTe/MgO vdWHs can also increase the ferroelectric spontaneous polarization with the bandgap unchanged [42]. The
Ni/SnTe vdWH realizes a possible scheme to obtain the multiferroics in 2D SnTe [42]. The GeSe/SnS vdWH has an intrinsic type-II
band alignment and their electronic structure and carrier effective masses can be effectively adjusted by strain [43,44]. Recently,
bilayer and three layers of IVA-monochalcogenides have been synthesized experimentally [19,45]. However, the exploration of
ferroelectricity in 2D IVA-monochalcogenides vdWHs is still rare.
In this letter, using first-principles calculations based on density functional theory (DFT) and modern theory of polarization, we
design the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs and calculate their lattice constants, formation energies, mechanical properties,
electronic structures, ferroelectric spontaneous polarizations, and potential barriers. Furthermore, taking SnSe/SnTe vdWH as an
example, we investigate the effects of external strains on its electronic properties and ferroelectricity. These results are an effective
supplement to the study and application of novel 2D ferroelectric vdWHs.

2. Computational details

All first-principles calculations are performed in the Vienna ab initio simulations package (VASP) [46,47]. The projector
augmented wave (PAW) pseudopotentials have been used [48]. The exchange-correlation functions are described within the general
gradient approximation (GGA) in the Perdew-Burke-Ernzerhof (PBE) functional [49]. The cutoff energy is set 500 eV. The first Bril­
louin Zone (BZ) is sampled using a Γ-centered k-points mesh of 11 × 11 × 1 [50]. A vacuum of 30 Å is set to eliminate the interaction
between adjacent films. The geometric structure optimizations are fully relaxed with the force and energy convergence criterions set as
0.01 eV/Å and 1.0 × 10− 5 eV, respectively. Considering the vdW interaction in vdWHs, the Grimm’s DFT-D3 method [51] is
implemented for structure optimization and self-consistent calculations. The modern theory of polarization based on the Berry phase
approximation is carried out to calculate the ferroelectric spontaneous polarization [52–54].
Phonon dispersion is calculated using the density functional perturbation theory (DFPT) [55] and implemented in the PHONOPY
package [56]. Ab initio molecular dynamics (AIMD) simulation is performed on a 3 × 6 × 1 supercell with 144 atoms to evaluate the
thermal stability of the SnSe/SnTe vdWH at 300 K for 2 ps.

3. Results and discussion

3.1. Lattice structures and stability

The ferroelectric spontaneous polarization of IVA-monochalcogenides MXs (M = Ge, Sn; X = S, Se, Te) is along the Armchair
direction [12,24] (See Fig. 1(a)), where the distortion distances are denoted by d1 and d2. The ferroelectric spontaneous polarization of
distorted ferroelectric phases B and B’ are represented by Ps and -Ps, respectively. Whereas, the paraelectric phase A has no ferroelectric
spontaneous polarization due to the center symmetry of undistorted structure (d1 = d2 = 0).
In order to design the vdWH based on monolayer SnTe, we first calculate the average lattice mismatches between monolayer SnTe

Fig. 1. (a) The side view distorted ferroelectric phases (B and B′ ) and paraelectric phase (A) of SnTe monolayer, where the red arrows represent the
ferroelectric spontaneous polarization along the Armchair direction. The side view of (b) SnSe/SnTe and (c) SnSe/SnTe/SnSe vdWHs. (d) The first
BZ with the high symmetry points.

2
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

and other IVA-monochalcogenides based on the lattice constants of MXs (M = Ge, Sn; X = S, Se, Te) monolayers reported by Xu et al.
[57] and Singh et al. [58] (See Table S1 in the supplemental material). The lattice mismatch is defined as: m = |a1 − a2 |/ am × 100%,
where a1 and a2 are the optimized lattice constants along the same direction of different MXs monolayers, respectively; am is the
maximum value of a1 and a2 . The average lattice mismatch is the average of the lattice mismatch along the Zigzag and Armchair
directions. The results show that the average lattice mismatches between monolayer SnTe and SnSe, PbSe, PbTe are 4.90%, 3.83% and
1.42%, respectively, which are less than 5% and acceptable. Considering the minimum lattice mismatch and environmental friend­
liness, we design the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs using monolayer SnTe and SnSe, as shown in Fig. 1(b) and (c),
respectively.
The optimized lattice constants and interlayer distances of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are listed in Table 1. The letters
a and b represent the optimized lattice constants along the Zigzag direction and Armchair direction, respectively. Due to the dual
effects of the interlayer vdW forces and the varying atomic radius, the optimized lattice constants of SnSe/SnTe vdWH are larger than
that of isolated monolayer SnSe (a = 4.293 Å, b = 4.389 Å) [57], but less than that of isolated monolayer SnTe (a = 4.559 Å, b = 4.570
Å) [57]. Similarly, the optimized lattice constants of SnSe/SnTe/SnSe vdWH are less than that of SnSe/SnTe vdWH. From the
interlayer distances dint of vdWHs, one can observe that the distance of Sn-Te is not equal to that of Sn-Se in SnSe/SnTe vdWH or
SnSe/SnTe/SnSe vdWH, indicating that optimized vdWHs are wrinkled structures. On the other hand, the interlayer distances of
SnSe/SnTe/SnSe vdWH are less than that of SnSe/SnTe vdWH due to the enhanced vdW forces.
To evaluate the stability, we calculate the formation energies ​ Ef of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs based on the for­
mula ​ Ef (eV) = ET − mESnSe − nESnTe , (See Table 1), where ET , ESnSe and ESnTe are the total energies of the vdWHs, isolated monolayer
SnSe and SnTe, respectively. The m and n are the number of monolayer SnSe and SnTe, respectively. The calculated formation energies
of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are − 0.661 eV and − 1.385 eV, respectively, which are comparable with that of GeSe/SnS
vdWH (~-0.530 eV) [44]. The negative formation energies indicate that the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are all ener­
getically stable and have the possibility of synthesis in experiments.
In addition, the dynamical stability and thermal stability of SnSe/SnTe vdWH are confirmed by the phonon dispersion and AIMD
with the canonical (NVT) ensemble, respectively. From the phonon dispersion (See Fig. S1 (a) in the supplemental material), one can
see that there is no imaginary frequency in the whole Brillouin zone, indicating that the SnSe/SnTe vdWH is dynamically stable.
Meanwhile, the fluctuation of free energy is less than 0.1 meV/atom, as shown in Fig. S1 (b) (See the supplemental material), meaning
that the SnSe/SnTe vdWH is also thermally stable at room temperature.

3.2. Mechanical properties

The mechanical properties provide information on the stability and stiffness of a material. Thus, we also calculate the elastic
constants Cij and Young’s moduli Y of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs (See Table 2). Firstly, one can see that the calculated
elastic constants of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs fulfill the mechanical stability criteria of 2D materials: C11C22 - C212 >
0 and C66 > 0 [59], indicating that both SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are mechanically stable. It is also obvious that the
Young’s modulus in Zigzag direction Y[100] is greater than that in Armchair direction Y[010], indicating that both SnSe/SnTe and
SnSe/SnTe/SnSe vdWHs are mechanical anisotropy, and the stiffness of Zigzag direction is greater than that of Armchair direction.
Furthermore, compared with the Young’s moduli of monolayer SnSe (Y[100] = 27.218 N/m, Y[010] = 14.801 N/m) and SnTe (Y[100]
= 26.328 N/m, Y[010] = 17.691 N/m), the Young’s moduli of vdWHs increase with the number of layers, indicating that the stiffness of
2D material increase with the number of layers. However,the Young’s module of SnSe/SnTe vdWH is less than that of BlackP/ZnO
(Y[100] = 143.60 N/m, Y[010] = 85.70 N/m) [60], BlackP/TiC2 (Y[100] = 140.80 N/m, Y[010] = 136.70 N/m) [61] and BlueP/ZnO
(Y[100] = 141.30 N/m, Y[010] = 139.00 N/m) [60], meaning that the SnSe/SnTe vdWH has better flexibility than other 2D vdWHs.

3.3. Electronic structures and ferroelectricities

From the band structures of SnSe/SnTe vdWH (See Fig. 2(a)), we find its conduction band minimum (CBM) and valence band
maximum (VBM) are located on Γ - Х path, showing a direct bandgap semiconductor. However, the CBM and VBM of SnSe/SnTe/SnSe
vdWH are located on Γ - Х path and Υ - Γ path, respectively (See Fig. 2(c)), demonstrating that it is an indirect bandgap semiconductor.
The bandgap values of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are 0.344 eV and 0.255 eV, respectively, which are both less than that
of monolayer SnSe (Indirect: 0.93 eV) and monolayer SnTe (Indirect: 0.72 eV). The calculated results indicate that stacking vdWHs not
only reduces the bandgap value but also realizes the transition between the indirect bandgap semiconductor and the direct bandgap
semiconductor. From the partial density of states (PDOS) (See the right panels in Fig. 2(a) and (c)), we find that the CBMs of SnSe/SnTe
and SnSe/SnTe/SnSe vdWHs are mainly contributed by the p orbital of Sn atoms, whereas their VBMs are formed by the hybridization

Table 1
Calculated lattice constants (a and b), interlayer distances (dint), and formation energies (Ef ) of the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs.
vdWHs Lattice constants (Å) Interlayer distances (Å) Formation energies (eV) Ef

a B dint (Sn-Te) dint (Sn-Se)

SnSe/SnTe 4.403 4.527 3.429 3.533 − 0.661


SnSe/SnTe/SnSe 4.345 4.472 3.327 3.296 − 1.385

3
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

Table 2
Calculated elastic constants (Cij) and Young’s moduli (Y) of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs. The standard Voigt notations are used: 1-xx, 2-
yy and 6-xy.
vdWHs Elastic constants (N/m) Young’s moduli (N/m)

C11 C22 C12 C66 Y[100] Y[010]

SnSe/SnTe 79.538 54.252 39.989 47.643 50.062 34.146


SnSe/SnTe/SnSe 124.239 84.444 63.780 68.042 76.066 51.702

Fig. 2. The band structures and PDOS of (a) SnSe/SnTe and (c) SnSe/SnTe/SnSe vdWHs. The arrows point from VBM to CBM. The weight of Sn, Te,
and Se atoms in band structures is in red, blue, and green, respectively. Double-well potentials of (b) the SnSe/SnTe vdWH and (d) SnSe/SnTe/SnSe
vdWH. Eb is the potential barrier.

Fig. 3. Berry phase polarization calculation of (a) SnSe/SnTe and (b) SnSe/SnTe/SnSe vdWHs. The distortion distance of the undistorted nonpolar
paraelectric phase is 0 Å. Ps and Pq represent the ferroelectric spontaneous polarization and polarization quantum, respectively.

4
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

of the p orbital of Te atoms and the s orbital of Sn atoms. The result can also be demonstrated by the weight of Sn, Te, and Se atoms in
the band structures.
Furthermore, we calculate a series of free energy of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs at different distortion distances (d1 =
d2 = d) along the Armchair direction and obtain the Double-well potentials by fitting Landau model (See Fig. 2(b) and (d)). It is obvious
that there are two lowest energy points, indicating that SnSe/SnTe and SnSe/SnTe/SnSe vdWHs have ferroelectric spontaneous po­
larization. The two lowest energy points represent the stable ferroelectric phase B and B’. In detail, the distortion distances of
ferroelectric in SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are d1 = d2 = ±0.032 Å and d1 = d2 = ±0.045 Å, respectively.
Meanwhile, we also obtain the free-energy contours of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs (See Fig. S2 (a) and (b) in the
supplemental material). The saddle point A between B and B′ corresponds to the paraelectric phase with a central symmetric structure.
In the process of polarization reversal, the reversal path strictly follows the dashed path of B→A→B′ , retaining the relation of d1 = d2.
The Eb in Fig. 2(b) and (d) are the potential barriers for the transformation from ferroelectric phase (B or B’) to paraelectric phase (A).
The potential barriers Eb of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are 0.591 meV and 0.790 meV, respectively. These small potential
barriers indicate that SnSe/SnTe and SnSe/SnTe/SnSe vdWHs can be reversed from the ferroelectric phase to the paraelectric phase
under the action of a smaller electric field.
Based on Berry phase approximation [52–54], we calculate the spontaneous polarization with distortion distance of SnSe/SnTe and
SnSe/SnTe/SnSe vdWHs. The maximum spontaneous polarization intensity is determined by the continuous evolution of the relation
between spontaneous polarization intensity and distortion distance function. The results show that the maximum spontaneous po­
larization of SnSe/SnTe vdWH is Ps = 0.661 × 10− 10 C/m (See Fig. 3(a)), which is smaller than that of monolayer SnSe (~1.51 × 10− 10
C/m) [12], monolayer SnTe (~2.326 × 10− 10 C/m) [24] and bilayer SnTe (~4.135 × 10− 10 C/m) [18]. The calculated ferroelectric
spontaneous polarization of SnSe/SnTe/SnSe vdWH is 1.385 × 10− 10 C/m (See Fig. 3(b)), which is comparable with that of three layers
of SnSe (0.617 × 10− 10 C/m) [12]. The possible cause of the results is the introduced Te atoms enhance the hybridization interaction
between atoms in the SnSe monolayers, and even the effect of Pauli exclusion on the interlayer force constant can be ignored. Finally,
we find the polarization quanta Pq of SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are 3.639 × 10− 10 C/m and 3.687 × 10− 10 C/m,
respectively, which are in good agreement with the theoretical formula 2eR/A, where e is the elemental charge, R is the lattice constant
along polarization direction (Armchair) and A is the unit cell’s area.

Fig. 4. The effects of external strains on (a) the total energy and (b) bandgaps of SnSe/SnTe vdWH. The positive and negative strain represent
tensile and compress strain, respectively. (c) The distortion distance, (d) the spontaneous polarization Ps and the potential barrier of SnSe/SnTe
vdWH under biaxial tensile strain.

5
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

3.4. Strain engineering

External strain is a feasible and effective strategy to modulate the physical properties of 2D materials. Taking the SnSe/SnTe vdWH
as an example, we investigate the effects of external strain on its total energy, bandgap, ferroelectric spontaneous polarization, and
potential barrier. Three types of strains have been considered: uniaxial strain along the Armchair direction, uniaxial strain along the
Zigzag direction, and biaxial strain (See Fig. S3 in the supplemental material). The positive and negative strains represent tensile strain
and compress strain, respectively.
The effects of uniaxial and biaxial strains on the total energy of SnSe/SnTe vdWH are shown in Fig. 4(a), where E’ and E0 are the
energy of SnSe/SnTe vdWH with and without strain, respectively. The results show that both tensile strain and compression strain can
increase the total energy of SnSe/SnTe vdWH, which is consistent with the energy variation of GeSe/SnS vdWH under strain [43]. One
can also observe that the effects of biaxial strains on the energy are greater than that of uniaxial strains. The possible reason is that the
distance between atoms in SnSe/SnTe vdWH decreases under biaxial compressive strain, leading to the atomic cluster state of atoms
and the metastable structure.
Fig. 4(b) shows the effects of uniaxial and biaxial strains on the bandgap of SnSe/SnTe vdWH. The result shows that the bandgap of
the SnSe/SnTe vdWH decreases with the increase of uniaxial and biaxial compression strains. The heterostructure exhibits the metallic
characteristics with 0 eV bandgap under the application of 6% compression strain along the Armchair direction, 4% compression strain
along the Zigzag direction, or 6% biaxial compression strain. Similarly, with the increase of uniaxial tensile strains along the Armchair
and Zigzag direction, the bandgap of SnSe/SnTe vdWH tends to decrease. However, the bandgap of the heterostructure gradually
increases with the increases of biaxial tensile strain.
The detailed changes of CBM and VBM of SnSe/SnTe vdWH are connected by red and blue dashed lines, respectively, as shown in
Fig. 5. From the fluctuation of the dashed lines, one can observe that the uniaxial strains along the Armchair and the Zigzag direction
have more effects on the VBM than that on the CBM, indicating that uniaxial strains mainly modulate the bandgap by changing its

Fig. 5. The evolutions of band structures of SnSe/SnTe vdWH under (a) uniaxial strain along Armchair direction, (b) uniaxial strain along Zigzag
direction, and (c) biaxial strain.

6
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

VBM. Moreover, the VBM and CBM of SnSe/SnTe vdWH constantly change between Γ - Х and Υ - Γ under uniaxial strain, leading the
transition between direct bandgap and indirect bandgap. Compared with uniaxial strains, the effects of biaxial tensile strain on the
VBM and CBM of SnSe/SnTe vdWH are more stable. The VBM and CBM of SnSe/SnTe vdWH always lie between Γ - Х under − 4%–4%
biaxial strain, indicating that the direct bandgap semiconductor performance is maintained. However, the SnSe/SnTe vdWH trans­
forms from direct bandgap to indirect bandgap when the biaxial tensile strain reaches 6%. In the following discussion, we only consider
the effects of biaxial tensile strain on the ferroelectricity of SnSe/SnTe vdWH.
Under different biaxial tensile strains, we first calculate the free energy of SnSe/SnTe vdWH and fit Double-well potentials (See
Fig. S4 in the supplemental material). The results show that the ferroelectric spontaneous polarization of SnSe/SnTe vdWH can be
maintained stably under 0%–10% biaxial tensile strains. Its distortion distance of the ferroelectric phase increases from 0.032 Å to
0.319 Å with the increases of the biaxial tensile strain from 0% to 10% (See Fig. 4(c)). In particular, the distortion distance increases
more obviously under 4%–6% biaxial tensile strain, which matches the mutation of the bandgap discussed above under 4%–6% biaxial
tensile strain. From the free-energy contours (See Fig. S5 in the supplemental material), one can see that the ferroelectric spontaneous
polarization reversal path from phase B to phase B′ strictly follows the path of B→A→B′ and maintains d1 = d2.
Furthermore, the ferroelectric spontaneous polarization and potential barrier of SnSe/SnTe vdWH under biaxial tensile strain have
been calculated (See Fig. 4(d)). The results show that the ferroelectric spontaneous polarization increases linearly to 1.721 × 10− 10 C/
m under the action of 0%–1% biaxial tensile strains. However, it gradually decreases until 0.208 × 10− 10 C/m as the biaxial tensile
strain continues to increase from 2% to 10%. Similarly, the spontaneous polarization has a large span drop under the action of 4%–6%
biaxial tensile strains, which is consistent with the increase sharply of the distortion distance. The variation trend of the influence of
biaxial tensile strain on ferroelectric spontaneous polarization of SnSe/SnTe vdWH is consistent with that of graphene/GeS vdWH
[41], proving that the ferroelectric spontaneous polarization of 2D vdWH can be increased by applying an appropriate external force. It
is worth noting that the potential barrier of SnSe/SnTe vdWH increases exponentially from 0.591 meV to 438.468 meV with the
increase of the biaxial tensile strain from 0% to 10%. However, when the spontaneous polarization of vdWH reaches the maximum, the
potential barrier is still at a small value of 7.593 meV, indicating that the spontaneous polarization of SnSe/SnTe vdWH can be reversed
by applying a small electric field.

4. Conclusion

In summary, based on the minimum lattice mismatch, we stack the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs. The small Young’s
moduli of the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs indicate that they show excellent flexibility. Using the first-principles calcu­
lation, we demonstrate that the SnSe/SnTe and SnSe/SnTe/SnSe vdWHs are direct bandgap semiconductor and indicate bandgap
semiconductor with small bandgaps, respectively, which can be transformed between direct bandgap, indirect bandgap and metal
under the action of external strains. Furthermore, based on the modern theory of polarization, we demonstrate that SnSe/SnTe and
SnSe/SnTe/SnSe vdWHs have stable ferroelectricity with small spontaneous polarizations and potential barriers. Particularly, the
ferroelectric spontaneous polarization of SnSe/SnTe vdWH increases to the maximum under the action of 1% biaxial tensile strain. The
results provide new clues for the study and application of 2D novel ferroelectric heterostructures.

Credit author statement

Hao Guo (The first author): Writing - Original Draft, Methodology, Software, Validation, Formal analysis, Data Curation. Xiaobao
Tian and Wentao Jiang (Corresponding authors): Conceptualization, Supervision, Funding acquisition, Writing - Review & Editing.
Haidong Fan: Software, Writing - Review & Editing. All authors of this manuscript agree to be accountable for all aspects of the work in
ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (No. 11602154), the Science and Technology
Planning Project of Sichuan Province (NO. 2019YJ0157).

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.spmi.2020.106728.

7
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

References

[1] M.E. Lines, A.M. Glass, Principles and Applications of Ferroelectrics and Related Materials, Oxford university press, 2001.
[2] J.F. Scott, Applications of modern ferroelectrics, Science 315 (5814) (2007) 954–959.
[3] M.W. Chu, I. Szafraniak, R. Scholz, et al., Impact of misfit dislocations on the polarization instability of epitaxial nanostructured ferroelectric perovskites, Nat.
Mater. 3 (2) (2004) 87–90.
[4] C. Lichtensteiger, J.M. Triscone, J. Junquera, et al., Ferroelectricity and tetragonality in ultrathin PbTiO3 films, Phys. Rev. Lett. 94 (4) (2005), 047603.
[5] M. Stengel, D. Vanderbilt, N.A. Spaldin, Enhancement of ferroelectricity at metal-oxide interfaces, Nat. Mater. 8 (5) (2009) 392–397.
[6] J. Junquera, P. Ghosez, Critical thickness for ferroelectricity in perovskite ultrathin films, Nature 422 (6931) (2003) 506–509.
[7] D.D. Fong, G.B. Stephenson, S.K. Streiffer, et al., Ferroelectricity in ultrathin perovskite films, Science 304 (5677) (2004) 1650–1653.
[8] L. Despont, C. Koitzsch, F. Clerc, et al., Direct evidence for ferroelectric polar distortion in ultrathin lead titanate perovskite films, Phys. Rev. B 73 (9) (2006),
094110.
[9] S.N. Shirodkar, U.V. Waghmare, Emergence of ferroelectricity at a metal-semiconductor transition in a 1T monolayer of MoS2, Phys. Rev. Lett. 112 (15) (2014)
157601.
[10] C. Liu, W.H. Wan, J. Ma, et al., Robust ferroelectricity in two-dimensional SbN and BiP, Nanoscale 10 (17) (2018) 7984–7990.
[11] D. Di Sante, A. Stroppa, P. Barone, et al., Emergence of ferroelectricity and spin-valley properties in two-dimensional honeycomb binary compounds, Phys. Rev.
B 91 (16) (2015), 161401(R).
[12] R. Fei, W. Kang, L. Yang, Ferroelectricity and phase transitions in monolayer group-IV monochalcogenides, Phys. Rev. Lett. 117 (9) (2016), 097601.
[13] H. Wang, X. Qian, Two-dimensional multiferroics in monolayer group IV monochalcogenides, 2D Mater. 4 (1) (2017), 015042.
[14] C. Xiao, F. Wang, S.A. Yang, et al., Elemental ferroelectricity and antiferroelectricity in group-V monolayer, Adv. Funct. Mater. 28 (17) (2018) 1707383.
[15] S. Raj Panday, B.M. Fregoso, Strong second harmonic generation in two-dimensional ferroelectric IV-monochalcogenides, J. Phys. Condens. Matter 29 (43)
(2017) 43LT01.
[16] S.P. Poudel, J.W. Villanova, S. Barraza-Lopez, Group-IV monochalcogenide monolayers: two-dimensional ferroelectrics with weak intralayer bonds and a
phosphorenelike monolayer dissociation energy, Phys. Rev. Mater. 3 (12) (2019) 124004.
[17] W. Wan, C. Liu, W. Xiao, et al., Promising ferroelectricity in 2D group IV tellurides: a first-principles study, Appl. Phys. Lett. 111 (13) (2017) 132904.
[18] T.P. Kaloni, K. Chang, B.J. Miller, et al., From an atomic layer to the bulk: low-temperature atomistic structure and ferroelectric and electronic properties of
SnTe films, Phys. Rev. B 99 (13) (2019) 134108.
[19] K. Chang, T.P. Kaloni, H. Lin, et al., Enhanced spontaneous polarization in ultrathin SnTe films with layered antipolar structure, Adv. Mater. 31 (3) (2019)
1804428.
[20] R. Fei, W. Li, J. Li, et al., Giant piezoelectricity of monolayer group IV monochalcogenides: SnSe, SnS, GeSe, and GeS, Appl. Phys. Lett. 107 (17) (2015) 173104.
[21] S. Barraza-Lopez, T.P. Kaloni, S.P. Poudel, et al., Tuning the ferroelectric-to-paraelectric transition temperature and dipole orientation of group-IV
monochalcogenide monolayers, Phys. Rev. B 97 (2) (2018), 024110.
[22] S. Guan, C. Liu, Y. Lu, et al., Tunable ferroelectricity and anisotropic electric transport in monolayer beta-GeSe, Phys. Rev. B 97 (14) (2018) 144104.
[23] L. Zhu, Y. Lu, L. Wang, Tuning ferroelectricity by charge doping in two-dimensional SnSe, J. Appl. Phys. 127 (1) (2020), 014101.
[24] K. Chang, J. Liu, H. Lin, et al., Discovery of robust in-plane ferroelectricity in atomic-thick SnTe, Science 353 (6296) (2016) 274–278.
[25] S.-H. Wei, S.B. Zhang, A. Zunger, First-principles calculation of band offsets, optical bowings, and defects in CdS, CdSe, CdTe, and their alloys, J. Appl. Phys. 87
(3) (2000) 1304–1311.
[26] X. Liu, C.-K. Tan, Structural and electronic properties of dilute-selenide gallium oxide, AIP Adv. 9 (12) (2019).
[27] M.Y. Li, C.H. Chen, Y.M. Shi, et al., Heterostructures based on two-dimensional layered materials and their potential applications, Mater. Today 19 (6) (2016)
322–335.
[28] K. Cheng, Y. Guo, N. Han, et al., Lateral heterostructures of monolayer group-IV monochalcogenides: band alignment and electronic properties, J. Mater. Chem.
C 5 (15) (2017) 3788–3795.
[29] W. Wu, Y. Liu, S. Li, et al., Nodal surface semimetals: theory and material realization, Phys. Rev. B 97 (11) (2018) 115125.
[30] Q. Li, X. Ma, L. Zhang, et al., Theoretical design of blue phosphorene/arsenene lateral heterostructures with superior electronic properties, J. Phys. D Appl. Phys.
51 (25) (2018) 1–10, 255304.
[31] K. Cheng, Y. Guo, N. Han, et al., 2D lateral heterostructures of group-III monochalcogenide: potential photovoltaic applications, Appl. Phys. Lett. 112 (14)
(2018) 1–5, 143902.
[32] C. Wang, L. Peng, Q. Qian, et al., Tuning the carrier confinement in GeS/Phosphorene van der Waals heterostructures, Small 14 (10) (2018) 1703536.
[33] Y.C. Huang, X. Chen, C. Wang, et al., Layer-dependent electronic properties of phosphorene-like materials and phosphorene-based van der Waals
heterostructures, Nanoscale 9 (25) (2017) 8616–8622.
[34] Y.J. Gong, J.H. Lin, X.L. Wang, et al., Vertical and in-plane heterostructures from WS2/MoS2 monolayers, Nat. Mater. 13 (12) (2014) 1135–1142.
[35] C.V. Nguyen, M. Idrees, H.V. Phuc, et al., Interlayer coupling and electric field controllable Schottky barriers and contact types in graphene/PbI2
heterostructures, Phys. Rev. B 101 (23) (2020).
[36] K.D. Pham, N.N. Hieu, H.V. Phuc, et al., Layered graphene/GaS van der Waals heterostructure: controlling the electronic properties and Schottky barrier by
vertical strain, Appl. Phys. Lett. 113 (17) (2018) 171605.
[37] M.A. Tito Patricio, R.R. LaPierre, Y.A. Pusep, Inter-valley phonon-assisted Auger recombination in InGaAs/InP quantum well, J. Appl. Phys. 125 (15) (2019).
[38] P.T.T. Le, N.N. Hieu, L.M. Bui, et al., Structural and electronic properties of a van der Waals heterostructure based on silicene and gallium selenide: effect of
strain and electric field, Phys. Chem. Chem. Phys. 20 (44) (2018) 27856–27864.
[39] H.V. Phuc, N.N. Hieu, B.D. Hoi, et al., Out-of-plane strain and electric field tunable electronic properties and Schottky contact of graphene/antimonene
heterostructure, Superlattice. Microst. 112 (2017) 554–560.
[40] T. Wang, Q. Zhang, J. Li, et al., 2D GeSe/SnS2(SnSe2) broken-gap heterostructures for tunnel field-effect transistors applications, J. Phys. D Appl. Phys. 52 (45)
(2019) 455103.
[41] Q.J. Wang, Q.H. Tan, Y.K. Liu, et al., Tunable electronic properties and giant spontaneous polarization in graphene/monolayer GeS van der Waals
heterostructure, Phys. Status Solidi B 256 (11) (2019) 1900194.
[42] Z. Fu, M. Liu, Z. Yang, Substrate effects on the in-plane ferroelectric polarization of two-dimensional SnTe, Phys. Rev. B 99 (20) (2019) 205425.
[43] S. Ahmed, T. Taher, R. Chakraborty, et al., First-principles study of strain engineered electronic properties of GeSe-SnS hetero-bilayer, J. Electron. Mater. 48 (10)
(2019) 6735–6741.
[44] C.X. Xia, J. Du, W.Q. Xiong, et al., A type-II GeSe/SnS heterobilayer with a suitable direct gap, superior optical absorption and broad spectrum for photovoltaic
applications, J. Mater. Chem. A 5 (26) (2017) 13400–13410.
[45] L. Li, Z. Chen, Y. Hu, et al., Single-Layer single-crystalline SnSe nanosheets, J. Am. Chem. Soc. 135 (4) (2013) 1213–1216.
[46] G. Kresse, J. Furthmuller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54 (16) (1996)
11169–11186.
[47] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B 47 (1) (1993) 558–561.
[48] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (24) (1994) 17953–17979.
[49] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett. 77 (18) (1996) 3865–3868.
[50] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B 13 (12) (1976) 5188–5192.
[51] S. Grimme, Semiempirical GGA-type density functional constructed with a long-range dispersion correction, J. Comput. Chem. 27 (15) (2006) 1787–1799.
[52] R.D. Kingsmith, D. Vanderbilt, Theory of polarization of crystalline solids, Phys. Rev. B 47 (3) (1993) 1651–1654.
[53] R. Resta, Macroscopic electric polarization as a geometric quantum phase, Europhys. Lett. 22 (2) (1993) 133–138.

8
H. Guo et al. Superlattices and Microstructures 148 (2020) 106728

[54] D. Vanderbilt, R.D. Kingsmith, Electric polarization as a bulk quantity and its relation to surface charge, Phys. Rev. B 48 (7) (1993) 4442–4455.
[55] S. Baroni, S. De Gironcoli, A. Dal Corso, et al., Phonons and related crystal properties from density-functional perturbation theory, Rev. Mod. Phys. 73 (2) (2001)
515.
[56] A. Togo, I. Tanaka, First principles phonon calculations in materials science, Scripta Mater. 108 (2015) 1–5.
[57] L. Xu, Ming Yang, Shi Jie Wang, Yuan Ping Feng, Electronic and optical properties of the monolayer group-IV monochalcogenides MX (M=Ge,Sn; X=S,Se,Te),
Phys. Rev. B 95 (23) (2017) 235434, 235434.
[58] A.K. Singh, R.G. Hennig, Computational prediction of two-dimensional group-IV mono-chalcogenides, Appl. Phys. Lett. 105 (4) (2014), 042103.
[59] F. Mouhat, F.-X. Coudert, Necessary and sufficient elastic stability conditions in various crystal systems, Phys. Rev. B 90 (22) (2014) 224104.
[60] X. Li, X. Wu, Z. Zhu, et al., On the elasticity and piezoelectricity of black(blue) phosphorus/ZnO van der Waals heterostructures, Comput. Mater. Sci. 169 (2019)
109134.
[61] Q. Peng, K. Hu, B. Sa, et al., Unexpected elastic isotropy in a black phosphorene/TiC2 van der Waals heterostructure with flexible Li-ion battery anode
applications, Nano Res. 10 (9) (2017) 3136–3150.

You might also like