You are on page 1of 17

Journal Pre-proof

Janus structure derivatives SnP-InS, GeP-GaS and SiP-AlS Monolayers with in-
plane and out-of-plane piezoelectric performance

Yan-Zhen Zhao, Hao-Jun Jia, Sheng-Nan Zhao, Yu-Bin Wang, Heng-Yu Li, Zhen-
Lin Zhao, Yu-Xuan Wu, Xiao-Chun Wang

PII: S1386-9477(19)31487-0
DOI: https://doi.org/10.1016/j.physe.2019.113817
Reference: PHYSE 113817

To appear in: Physica E: Low-dimensional Systems and Nanostructures

Received Date: 29 September 2019


Accepted Date: 01 November 2019

Please cite this article as: Yan-Zhen Zhao, Hao-Jun Jia, Sheng-Nan Zhao, Yu-Bin Wang, Heng-Yu
Li, Zhen-Lin Zhao, Yu-Xuan Wu, Xiao-Chun Wang, Janus structure derivatives SnP-InS, GeP-GaS
and SiP-AlS Monolayers with in-plane and out-of-plane piezoelectric performance, Physica E: Low-
dimensional Systems and Nanostructures (2019), https://doi.org/10.1016/j.physe.2019.113817

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

Janus structure derivatives SnP-InS, GeP-GaS and SiP-AlS

Monolayers with in-plane and out-of-plane piezoelectric performance


Yan-Zhen Zhaoa,b†, Hao-Jun Jiac†, Sheng-Nan Zhaoa, Yu-Bin Wanga,b, Heng-Yu Lid,
Zhen-Lin Zhaoe, Yu-Xuan Wub, Xiao-Chun Wanga*

aInstitute of Atomic and Molecular Physics, Jilin University, Changchun, 130012,


China
bCollege of Physics, Jilin University, Changchun, 130012, China

cDepartment of Chemistry, Massachusetts Institute of Technology, Cambridge,


Massachusetts 02139, USA

dInstitute for Solid State Physics, The University of Tokyo, Kashiwa 277-8581, Japan

eDepartment of Materials Science and Engineering, Stanford University, Stanford,


CA, 94305 USA

†These authors contribute equally to this work


* Corresponding author E-mail address: wangxiaochun@jlu.edu.cn

Abstract
The piezoelectricity of two-dimensional (2D) materials has triggered the
development in a wide variety of fields. Since the symmetry of materials has a close
relationship with piezoelectricity, a series of derivatives of the Janus structure was
investigated on the piezoelectric effect of GeP-GaS, SiP-AlS, and SnP-InS. Density
functional theory (DFT) was employed to explore the stiffness and the piezoelectric
tensors. The charge density distribution was calculated to explain the polarization and
chemical bonds. Compared with monolayer group III monochalcogenides, such as
SiN and SnP who only possess the in-plane direction piezoelectricity, the results show
that all three systems have the out-of-plane piezoelectricity. The emergence of
out-of-plane piezoelectricity is caused by the unmirrored structure of GeP-GaS,
SiP-AlS, and SnP-InS. Simultaneously, the measured d31 of SnP-InS is 9.69 pm/V and
is higher than that of the other two systems. Therefore, the high energy conversion
efficiency between the electrical and the mechanical domain shows their great
1
Journal Pre-proof

potential in applications. Moreover, the intrinsic piezoelectricity of the Janus structure


in the out-of-plane direction also suggests that replacing the atoms is an effective
method to explore novel piezoelectric materials.

Keywords: piezoelectricity, Janus structure, 2D materials, first-principles


calculations

Introduction
The piezoelectric effect is a reversible physical process which can convert energy
from the mechanical domain to the electrical domain. This effect has become
increasingly important due to its various applications.1 For instance, materials
possessing the piezoelectric effect are used in applications at the nanoscale level, such
as actuators2, sensors3, motors4, locomotive devices5, energy harvesting devices6 and
medical devices7. Due to the increasing demand for miniaturized devices, the
dimension limitations on piezoelectric structures have gradually been recognized.
Inspired by Xin’s work8, in which a series of wurtzite-structured semiconductors was
analyzed in detail to explain the relation between the structure and its piezoelectric
properties, a series of materials with novel piezoelectric properties and structural
modulations was investigated.
Two-dimensional (2D) materials possess many charming characteristics, such as
on-site proton transfer9 and piezoelectricity10,11 Interestingly, the emergence of
piezoelectricity in 2D materials that are not piezoelectric in their bulk counterparts has
already been the subject of several theoretical predictions and experimental
characterizations.12 For instance, recently several predicted 2D materials, such as
metal dichalcogenides (MX2, where M = Cr, Mo, W, Nb, and Ta; X = S, Se, and Te)
and the group-IV monochalcogenides (MX, M =B, Ga, Al, and In; X = O, S, Se, and
Te) were suggested to be intrinsically piezoelectric.13-15 The search for piezoelectric
materials is an ongoing process in spite of the number of currently well-known
materials.16 Additionally, the centrosymmetry is a criteria to discern whether systems
are piezoelectric or not.17
The out-of-plane piezoelectricity is not common in 2D materials.18,19 Many
structurally broken inversion symmetrical 2D monolayers such as transition metal
dichalcogenides (TMDC) only show in-plane piezoelectricity. Therefore, applications
such as vertically integrated nanoelectromechanical systems20 and piezoelectric
cantilever and diaphragm devices (for example: loudspeaker) can be limited. To be
specific, the cantilever and diaphragm structures are easy to vibrate back and forth.
Thus, the bottom/top gate technologies can be restricted when the structures lack the
vertical piezoelectricity.21 In order to enhance the piezoelectricity performance, three
monolayer systems, GeP-GaS, SiP-AlS, and SnP-InS, were analyzed. Density
functional theory (DFT) was employed to calculate the piezoelectric effect of these
monolayers. This simulation method has been widely applied to calculate the
2
Journal Pre-proof

characteristics of various systems and to define the degree of agreement between the
theoretical and the experimental data. The measured piezoelectricity of GeP-GaS,
SiP-AlS, and SnP-InS were then compared to the mirrored monolayers, such as GaS.
The inversion symmetry of our models was broken. The asymmetric ion distribution
causes the out-of-plane piezoelectric effect. Moreover, these systems belong to the C3v
space group. Recent studies on the systems investigated in this paper have
demonstrated that the replacement of several atoms enhances the cohesive energy as
well as the electronic properties.22 Based on the aforementioned observations, the
exploration of piezoelectricity leads to a deep understanding of its mechanisms. The
identification of novel 2D piezoelectric materials can expand the family of
piezoelectric materials and foster the development of novel tunable electronic devices
and nanoelectromechanical systems (NEMS). 11,23

Computational Methods
A series of DFT24,25 calculations was performed using the Vienna Ab initio
Simulation Package (VASP)26,27 within the framework of the projector augmented
wave (PAW) method. The exchange correlation functionals were described via the
Perdew-Burke-Ernzerhof (PBE)28 form within the generalized gradient approximation
(GGA). A plane-wave basis set with an energy cut-off of 500 eV was used to expand
the wave functions. The Brillouin zone integration was obtained by setting a 11 × 18
× 1 k-point grid within the Monkhorst-Pack scheme for the nanosheets.19 For all the
calculations, a vacuum spacing of 20 Å sufficiently reduced the interactions between
neighboring layers. Each atom was fully relaxed until the force on each one reached
10-4 eV Å-1. Moreover, the convergence criteria for the total energy was set to 10-6 eV.
The schematic images of the charge density distribution and the crystal structures
were performed by employing VESTA.29 Furthermore, the elastic tensor coefficients
and the stress coefficients of the piezoelectric tensor were calculated by using density
functional perturbation theory (DFPT).30

Results and Discussion


The crystal structures SiP-AlS, GeP-GaS, and SnP-InS, consisting of a buckled
honeycomb monolayer, belong to the C3v space group. They show similarities with
the fundamental structure units of the monolayer group III monochalcogenides, such
as GaS. The results of the calculations show that lattice constant of GaS is 3.70 Å,
which is in agreement with previous theoretical studies.31 Due to the similar structure
of these three systems, only GeP-GaS is discussed in this section as an example. The
asymmetric structure of this compound (Fig. 1) shows several differences from the
group III monochalcogenides. The absence of the inversion symmetry in GeP-GaS is
responsible for the generation of the piezoelectric effect, especially along the direction
perpendicular to the plane of the monolayer. Several stable geometric structures of
GeP-GaS are presented in Fig. 1. The top and side views are regularly arranged in the
sequence S(1)-X(2)-M(3)-P(4) (X = Ga, Al, In; M= Ge, Si, Sn). By looking at the
sample from the top, the material shows a honeycomb structure. The sp3-like bonds
3
Journal Pre-proof

connecting S(1), X(2) and connecting M(3), P (4), and the sp2-like bonds connecting
X(2), and M(3) generate a sandwich-like structure where the X(2) and M(3) atoms are
located in the middle.
The optimized structural parameters of the three systems investigated in this work
are listed in Table 1. Here, a0 is the lattice constant; h is the sheet thickness; d
represents the distance between the middle of the two sublayers. As shown in the
Table 1, the structural parameters calculated with GGA-PBE are in perfect agreement
with previous works.15 In addition, the S(1)-X(2), X(2)-M(3), and M(3)-P(4) bonds (X
= Ga, Al, In; M= Ge, Si, Sn) provide a good measurements of the length. The
magnitudes of the S(1)-X(2) and M(3)-P(4) bonds are similar and smaller than
X(2)-M(3) for all three systems. The length of the In(2)-Sn(3) bond in the SnP-InS
system is larger than that in the other two systems, indicating that SnP-InS possesses a
flexible structure.
Analyzing the spatial charge density transfer process can be helpful to understand
the piezoelectric effect. Initially, the three-dimensional charge density difference was
demonstrated in Fig. 232 and this helps to visualize the donor and the acceptor atoms
involved in the electron transfer process. The charge transfer between each atom is
defined as Δρ(r) = ρ[MP-XS]−Σμρatom(r-Rμ).33 Here, ρ[MP-XS] represents the total
charge of the entire system and ρatom is calculated on each independent atom
separately by using the same atomic coordinates of the corresponding pristine
systems. Fig. 2 shows the three-dimensional (3D) charge density difference; the
yellow and the blue highlighted regions denote the accumulations and depletions of
electrons, individually.
To quantify the effective electron distribution and the variation in the MS-XP
dipole behavior, the Bader charge analysis34-36 was implemented. As depicted in Fig.
2, due to the stronger electronegativity of the P and S atoms, several electrons move
from the group V and VI atoms to the P and S atoms respectively. Taking SiP-AlS as
an example, the calculated effective charges of Si, Al, P, and S are expected to be
0.683 e, 1.987 e, -1.181 e, and -1.489 e separately. Since S is the element with the
highest electronegativity in each system, it accepts most of the charges. Moreover, the
S atoms accept more charges in the SiP-AlS monolayer than those in the
GeP-GaS(-0.758 e) and SnP-InS (-0.737 e).
According to the Bader analysis, the nature of the chemical bonds and of the
piezoelectricity can be explained. The chemical bonds’ properties are related to the
charge distribution. The accumulation region can be considered as partially negative
and the depletion region can be considered as partially positive. Therefore, the
original covalent bonds within the nanosheet possess an ionic character. Moreover,
compared with mirrored monolayer structures, such electron redistribution for MS-XP
without mirror symmetry can easily cause a polarization effect along the z-direction
while applying a strain. Thus, the material may be piezoelectric because of the
polarization effect.
Apart from the analysis of the 3D charge density difference, the two-dimensional
charge density difference of the three systems shown in Fig. 3 can be used to
investigate the interior atoms charge distribution.35,37 The red dashed line in Fig. 3 (a)
4
Journal Pre-proof

shows a cut of the primitive cell along the rhombic long half-axis. The electron
transfer between the S atoms and the M atoms is more intense than the one between
the P atoms and the X atoms, as shown in Fig. 3. For example, SiP-AlS shows the
strongest charge transfer effect, whereas a relatively low electron transfer is observed
in SnP-InS. The value of the charge density difference between the S and Al atoms
ranges from -0.161 to 0.191 e/Å3. This variation is larger than the transfer between the
S and In atoms (-0.161–0.103 e/Å3). This can be explained by the low
electronegativity value that the M atoms show in all the three system and by the
charge absorption process among the S atoms and the bottom sublayer. The variation
of charge intensity of the bottom sublayer can be offset by the charge transfer from
the M atoms. Moreover, the high electronegativity of the S atoms can enhance the
intensity of the transfer process of the top sublayer.
Further investigations are proposed to study the electronic properties of MP-XS.
In Fig. 4, the DOS images are almost identical. However, the peak shows small shifts
at the Fermi level while a force is applied.38 For instance, in the case of a -0.01
uniaxial strain, the peak near the Fermi level shifts towards lower energies in all the
three systems. Additionally, the -0.01 uniaxial strain leads the conduction band
minimum (CBM) to shift towards higher energies. For instance, the CBM of SnP-InS
shifts about 0.07 eV from its original state (1.31 eV), and this change is larger than
that in the other two systems. The band gaps become slightly wider in SnP-InS and
SiP-AlS. In conclusion, the change in the density of states of SnP-InS due to a strain
demonstrates that the charge transfer in the SnP-InS occurs more easily than that in
the other two systems.
The elastic constants are indispensable to calculate the piezoelectric coefficients
of the 2D materials. The values of the nonzero independent clamped-ion and
relaxed-ion elastic stiffness coefficients, C11 and C12 (C11 = C12 for hexagonal
structures) are listed in Table 2. The results are obtained via the finite differences
method implemented in the VASP codes. Each calculation is based on a 7 × 7 k-point
grid within the Monkhorst-Pack scheme. When the atomic positions are fully relaxed,
the so-called relaxed-ion coefficients can be generated and they are expected to be
experimentally measurable. The calculation of relaxed-ion condition needs to take
into account both the electronic and the ionic contributions. Moreover, the calculated
elastic coefficients are positive and satisfy the Born stability criteria for crystals,
which possesses a hexagonal symmetry.39,40 For the frozen case, however, the
so-called clamped-ion coefficients only consider the electronic contributions.
In addition to the clamped-ion and the relaxed-ion elastic coefficients, the
stiffness of the materials is also important for their corresponding piezoelectric
process. Therefore, the Poisson’s ratios and Young’s modulus are also considered in
this study. The Poisson‘s ratios (ν⊥ ) of SiP-AlS, GeP-GaS, and SnP-InS govern the
out-of-plane variation of h in response to an in-plane strain. The ratios can be
obtained from the relaxed structures for harmonic elastic deformations according to
the formula Δh/h = − v⊥ (ε11 + ε12). The formula can be simplified by defining v⊥ =
C11/ C12 based on H. Şahin’s work.41 Table 2 shows that the Poisson’s ratio of
SnP-InS (0.25) is higher than that of the other two systems. This means that SnP-InS
5
Journal Pre-proof

can undergo in-plane expansion more easily. Theoretically, the lower Young’s
modulus is, the more flexible is the structure in terms of elasticity and the higher is its
piezoelectric response. Based on the calculation of Young’s modulus (Y = (C112 –
C122)/C11), the structures investigated in this study are less stiff when compared to
graphene (Y = 341 N/m). SnP-InS possesses the lowest tensor value (Y = 64.4 N/m)
when compared to SiP-AlS (Y = 102.9 N/m) and GeP-GaS (Y = 91.1 N/m). In fact, the
magnitude of the SnP-InS tensor is close to one fifth of the value of graphene and one
fourth of the h-Bn monolayer (Y = 275.9 N/m). This may be the reason for SnP-InS’s
high piezoelectric strain coefficients, d31.
The linear piezoelectric strain coefficients (dijk) of the Janus structure derivatives
MS-XP are then calculated by evaluating the change in the polarization (eijk) upon
imposing uniaxial strains (εijk). These calculations are based on the modern
polarization theory.42 Both eijk and dijk are third-rank tensors and Pi is the polarization
vector which depends on the stress (σijk) and the strain (εijk).

∂𝑃𝑖
𝑒𝑖𝑗𝑘 = ∂𝜀𝑗𝑘 (1)
∂𝑃𝑖
𝑑𝑖𝑗𝑘 = ∂𝜎𝑗𝑘 (2)

The Voigt notation is employed in this work: i, j, and k subscripts correspond to


the x, y, and z directions respectively. The number of the non-zero and unique
piezoelectric coefficients is restricted by the symmetry elements of the crystal.
Because of the C3v point group symmetry, the values of e11, d11, e31, and d31 for the
systems investigated in this work are all unique piezoelectric coefficients. In other
words, the in-plane and out-of-plane piezoelectric responses coexist. These
parameters can also be related to the elastic stiffness coefficients as follows:

𝑒11 = 𝑑11(𝐶11 ― 𝐶12) (3)


𝑒31 = 𝑑31(𝐶11 + 𝐶12) (4)

The piezoelectric stress coefficients e11 and e31 were computed first by fitting
linearly Eqn. (1) and (2) with the Berry’s Phase method. The use of the Berry’s Phase
method to calculate the piezoelectricity has already been employed in many systems
such as MoS2 and GaS.12,31 By applying the gradient strains ε11 along the x-direction
of the orthorhombic supercell (Fig. 1), the changes in the polarization vector, ΔP1,
along x-direction, and in ΔP3 along z-direction were obtained. The atoms were either
relaxed or rigidly fixed to obtain the corresponding relaxed or clamped-ion
piezoelectric coefficients. Fig. 5 illustrates the linear fitting polarization change in ΔP1
as a function of ε11 and in ΔP3 as a function of ε11. The slopes of the fitted lines can be
used to derive the clamped-ion and relaxed-ion piezoelectric stress coefficients e11
and e31.
In Fig. 5, we use ε11 ranging from -0.01 to 0.01 with a step of 0.0025 along the
x-direction. By reducing the step, one can avoid the influence on the results caused by
6
Journal Pre-proof

extravagant displacements. The clamped-ion conditions are more linear than their
relaxed counterparts, since the impact of the ion contributions was neglected.43When
the atomic positions are relaxed, uniaxial strains eliminate the polarization of the
monolayer, especially in the case of e11. The clamped in-plane piezoelectric stress
coefficient, e11, of GeP-GaS, SiP-AlS, and SnP-InS are nearly four to nine times
larger than the ones of their relaxed structure. In addition, the negative relaxed-ion
piezoelectric stress coefficients, e31, of GeP-GaS and SiP-AlS show that the
polarization along the z-direction is inversely proportional to the strains along the
x-direction. This different behavior for similar structures illustrates that the
differences in the polarization direction may be caused by the different
electronegativity for each atom. As a result, stronger electron adsorption in the top or
the bottom layer during the shape change process can be detected.
According to Eqn. (3) and (4), the clamped-ion and relaxed-ion piezoelectric
strain coefficients, d11 and d31, together with their corresponding tensors were
eventually obtained. Table 3 lists their values for SiP-AlS, GeP-GaS, and SnP-InS.
The relaxed-ion e31 coefficients of these three systems are overall larger than their e11
counterparts. This means that a stronger piezoelectric response is generated
perpendicularly while applying a strain along the x-direction. One can notice that the
relaxed-ion strain coefficient d31 of SnP-InS (9.69 pm/V) possesses a larger absolute
value than that of GeP-GaS (-8.55 pm/V) and SiP-AlS (-4.61 pm/V). The flexible
structure and the intense electron transfer generated from In to S and from Sn to P
may explain the large value of the piezoelectric strain coefficient (d31) of SnP-InS.
Therefore, this structure is characterized by a more efficient energy conversion
process. However, SnP-InS is the compound which presents the lowest value of e11
(0.49×10-10 C·m-1) and d11 (0.95 pm·V-1), as shown in Table 3. The relaxed GeP-GaS
and SiP-AlS monolayers have a similarly large piezoelectric e11 coefficients (e11 =
1.14×10-10 C·m-1, d11 = 1.51 pm·V-1, and e11 = 1.14×10-10 C·m-1, and d11 = 1.34 pm·V-1,
respectively).

Conclusion
In this paper, the piezoelectric coefficients for a series of 2D materials, including
GeP-GaS, SiP-AlS, and SnP-InS, were investigated. Both the in-plane piezoelectric
strain coefficients and the out-of-plane piezoelectric strain coefficients were
measured. Moreover, the out-of-plane piezoelectric effect in each system performs
better than the in-plane counterpart. The coefficient d31 which does not exist in
mirrored monolayer Janus structure also appears in our Janus structure derivatives.
The d31 of SnP-InS is 9.69 pm/V which is higher than the value of GeP-GaS (-8.55
pm/V) and SiP-AlS (-4.61 pm/V). The high energy conversion efficiency for SnP-InS
can be attributed to the intense electron transfer and the flexible structure. Moreover,
the authors believe that many two-dimension mirrored structures can also become
piezoelectric along out-of-plane direction by replacing different atoms, such as by
replacing P with Bi in the systems. Finally, encouraged by several successfully
synthesis experiments of Janus structures, such as the 2D group-III
7
Journal Pre-proof

monochalcogenides, these systems can be experimentally fabricated in the near future


given their high stability and outstanding piezoelectricity.

Acknowledgements
This work was supported by the Natural Science Foundation of Jilin Province of
China under Grant no. 20170101154JC. Yan-Zhen Zhao and Hao-Jun Jia contributed
equally to this work.

References
1M. de Jong, W. Chen, H. Geerlings, M. Asta, and K. A. Persson, Sci Data 2, 150053 (2015).
2Z. L. Wang and J. H. Song, Science 312, 242 (2006).
3B. Morten, G. De Cicco, and M. Prudenziati, Sens. and Actuators A 31, 153 (1992).
4R. A. van Delden, M. K. J. ter Wiel, M. M. Pollard, J. Vicario, N. Koumura, and B. L. Feringa, Nature 437, 1337
(2005).
5Y. Shirai, A. J. Osgood, Y. M. Zhao, K. F. Kelly, and J. M. Tour, Nano Lett. 5, 2330 (2005).
6Y. Qin, X. Wang, and Z. L. Wang, Nature 451, 809 (2008).
7M. Yuan, L. Cheng, Q. Xu, W. Wu, S. Bai, L. Gu, Z. Wang, J. Lu, H. Li, Y. Qin, T. Jing, and Z. L. Wang, Adv.
Mater. 26, 7432 (2014).
8J. Xin, Y. Zheng, and E. Shi, Appl. Phys. Lett. 91 (2007).
9L. L. Liu, Y. Wang, C. P. Chen, H. X. Yu, L. S. Zhao, and X. C. Wang, Rsc Adv. 7, 40200 (2017).
10M. N. Blonsky, H. L. Zhuang, A. K. Singh, and R. G. Hennig, Acs Nano 9, 9885 (2015).
11M. T. Ong and E. J. Reed, Acs Nano 6, 1387 (2012).
12K. A. N. Duerloo, M. T. Ong, and E. J. Reed, J. Phys. Chem. Lett. 3, 2871 (2012).
13R. Fei, W. Li, J. Li, and L. Yang, Appl. Phys. Lett. 107 (2015).
14Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, Nat Nanotechnol 7, 699 (2012).
15H. L. Zhuang and R. G. Hennig, Chem. Mater. 25, 3232 (2013).
16H. J. Jia, H. M. Mu, J. P. Li, Y. Z. Zhao, Y. X. Wu, and X. C. Wang, Phys. Chem. Chem. Phys. 20, 26288
(2018).
17L. C. Gomes and A. Carvalho, Phys. Rev. B 92, 085406 (2015).
18L. Hu and X. Huang, Rsc Adv. 7, 55034 (2017).
19Y. Guo, S. Zhou, Y. Bai, and J. Zhao, Appl. Phys. Lett. 110 (2017).
20F. Xue, J. Zhang, W. Hu, W. T. Hsu, A. Han, S. F. Leung, J. K. Huang, Y. Wan, S. Liu, J. Zhang, J. H. He, W.
H. Chang, Z. L. Wang, X. Zhang, and L. J. Li, ACS Nano 12, 4976 (2018).
21J. Tan, Y. Wang, Z. Wang, X. He, Y. Liu, B. Wang, M. I. Katsnelson, and S. Yuan, Nano Energy 65, 104058
(2019).
22J. H. Lin, H. Zhang, X. L. Cheng, and Y. Miyamoto, Phys. Rev. B 96 (2017).
23H. G. Craighead, Science 290, 1532 (2000).
24P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).
25W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).
26G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).
27G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
28J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).

8
Journal Pre-proof

29K. Momma and F. Izumi, J. Appl. Crystallogr. 44, 1272 (2011).


30X. Wu, D. Vanderbilt, and D. R. Hamann, Phys. Rev. B 72, 035105 (2005).
31W. Li and J. Li, Nano Res. 8, 3796 (2015).
32C. Du, L. Yu, X. Liu, L. Liu, and C. Z. Wang, Sci Rep 7, 13152 (2017).
33J. P. Li, H. J. Jia, D. R. Zhu, X. C. Wang, F. C. Liu, and Y. J. Yang, Appl. Surf. Sci. 463, 918 (2019).
34L. L. Liu, C. P. Chen, L. S. Zhao, Y. Wang, and X. C. Wang, Carbon 115, 773 (2017).
35L. S. Zhao, C. P. Chen, L. L. Liu, H. X. Yu, Y. Chen, and X. C. Wang, Chin. Phys. B 27 (2018).
36W. Tang, E. Sanville, and G. Henkelman, Journal of Phys. Condens. Matter 21 (2009).
37X. Liu, C. Z. Wang, Y. X. Yao, W. C. Lu, M. Hupalo, M. C. Tringides, and K. M. Ho, Phys. Rev. B 83 (2011).
38L. Li, K. Jin, C. Du, and X. Liu, Rsc Adv. 9, 8253 (2019).
39M. M. Alyoruk, Y. Aierken, D. Cakir, F. M. Peeters, and C. Sevik, J. Phys. Chem. C 119, 23231 (2015).
40F. Mouhat and F.-X. Coudert, Phys. Rev. B 90, 224104 (2014).
41H. Şahin, S. Cahangirov, M. Topsakal, E. Bekaroglu, E. Akturk, R. T. Senger, and S. Ciraci, Phys. Rev. B 80,
155453 (2009).
42R. D. King-Smith and D. Vanderbilt, Phys. Rev. B 47, 1651 (1993).
43D. Vanderbilt, J. Phys. Chem. Solids 61, 147 (2000).

Fig. 1 (a) and (b) Top and side views of the GeP-GaS monolayer. The sulfur atoms are yellow, the gallium atoms
are green, the purple atoms are germanium, and the grey atoms are phosphorus. The rhombus primitive cell and the
orthorhombic unit cell, which are highlighted by the dashed lines, are used in the DFT calculations. The atoms of
the monolayer are in the sequence S(1)–X(2)–M(3)–P(4), where X = Ga, Al, In and M = Ge, Si, Sn.

Fig. 2 Top and side views of (a)-(b) SiP-AlS, (c)-(d) GeP-GaS, and (e)-(f) SnP-InS, respectively, showing the
three-dimensional (3D) charge density difference. By using the Bader charge analysis, the yellow and light blue
regions indicate the acceptor and donor areas in the material.

Fig. 3 Two-dimensional (2D) charge density difference of (b) SiP-AlS, (c) GeP-GaS, and (d) SnP-InS. Their
planes are shown along the red dashed line plotted in (a). The differences in color reflect the intensity of the
interaction between each atom and the charge density distribution of the entire system. Positive values correspond
to an increase in the charge density while negative values correspond to a loss of charge density.

Fig. 4 (a), (b), and (c) Electronic density of states (DOS) for SiP-AlS, GeP-GaS, SnP-InS, respectively. The

different colored lines correspond to the DOS under a uniaxial strain ranging from -0.01 to 0.01.

Fig. 5 (a), (c) and (b), (d) Change of polarization along the x-direction and the z-direction under the gradient strain
along the x-direction, for clamped-ion and the relaxed-ion respectively. The slopes of the fitted lines represent the
piezoelectric stress coefficients e11 and e31.

9
Journal Pre-proof

Table 1 Optimized lattice constants (a0), thickness (h), distance between middle two sublayers (d), bond lengths
between the S atoms and the X atoms (S(1)-X(2)), bond length between the X atoms and the M atoms (X(2)-M(3)),
and bond length between the M atoms and the P atoms (M(3)-P(4)) are listed.

S(1)-X(2) X(2)-M(3 M(3)-P(4


Structures a0(Å) h(Å) d(Å)
(Å) )(Å) )(Å)
GeP-GaS 3.655 4.646 2.487 2.370 2.487 2.370
SiP-AlS 3.555 4.551 2.467 2.313 2.467 2.291
SnP-InS 3.949 5.206 2.862 2.566 2.862 2.561
GaS 3.697 4.960 2.534 2.455 2.534 2.455

Table 2 Clamped-ion and relaxed-ion elastic stiffness coefficients, C11 and C12, calculated via the finite difference

method (in units of N/m), Poisson’s Ration v⊥,and Young’s modulus Y (in units of N/m).

Clamped-ion Relaxed-ion
Structures
C11 C12 C11 C12 v⊥ Y
GeP-GaS 119.27 28.76 95.34 20.09 0.21 91.11
SiP-AlS 133.24 30.45 107.55 22.34 0.21 102.91
SnP-InS 97.47 27.70 68.76 17.37 0.25 64.37

Table 3 Piezoelectric coefficients e11, e31, d11, and d31 are shown both in the clamped-ion case and in the
relaxed-ion situation. The values of e11 and e31 are expressed in units of 10-10 C m-1 and the d11 and d31 are in units
of pm V-1

Clamped-ion Relaxed-ion
Structures
e11 e31 d11 d31 e11 e31 d11 d31
GeP-GaS 4.31 -9.53 4.76 -6.44 1.14 -9.87 1.51 -8.55
SiP-AlS 4.14 3.77 4.03 2.30 1.14 -5.99 1.34 -4.61
SnP-InS 4.75 8.73 6.81 6.98 0.49 8.34 0.95 9.69

10
Journal Pre-proof

The enhancement of piezoelectric performance in mirrored monolayer Janus structure


along out-of-plane direction.

A new way to explore more two-dimension piezoelectric materials by sequencing


different atoms.

The intense charge transfer triggers the qualified piezoelectic effect.

You might also like