Next Article in Journal
An Enhanced Deep Learning-Based Pharmaceutical Crystal Detection with Regional Filtering
Previous Article in Journal
Influence of Graphene Oxide Concentration and Ultrasonication Energy on Fracture Behavior of Nano-Reinforced Cement Pastes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coexistence of Large In-Plane and Out-of-Plane Piezoelectric Response in Group III–VI XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) Monolayers

1
School of Physics and Optoelectronic Engineering, Guangdong University of Technology, Guangzhou 510006, China
2
Guangdong Provincial Key Laboratory of Information Photonics Technology, Guangdong University of Technology, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(8), 708; https://doi.org/10.3390/cryst14080708
Submission received: 13 July 2024 / Revised: 26 July 2024 / Accepted: 2 August 2024 / Published: 5 August 2024
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
Flexible materials with both in-plane and out-of-plane piezoelectric coefficients are needed in the development of advanced nanoelectromechanical systems. However, the challenge is to find flexible materials with the coexistence of in-plane and -out-of-plane piezoelectric responses, which hinders the progress of high-performance piezoelectric sensor development. In this paper, we propose the flexible XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) monolayers, which belong to the group III-VI XMAY2 family, which showcase notable in-plane and out-of-plane piezoelectric coefficients. The in-plane (d11) and out-of-plane (d31) piezoelectric coefficients of the XMAY2 monolayers vary from 5.20 to 7.04 pm/V and from −0.23 to 0.48 pm/V, respectively. The large in-plane and out-plane piezoelectric responses coexist (d11 = 7.04 pm/V; d31 = 0.48 pm/V) in the IZrGaS2 monolayer, which is larger than other materials in the XMAY2 family, such as SMoSiN2 (d11 = 2.51; d31 = 0.28 pm/V). In addition, the mechanical and transport properties of XMAY2 demonstrate its impressive flexibility characteristics as well as its efficient electrical conductivity. Due to inversion symmetry breaking in both atomic structure and charge distribution of XMAY2 monolayers, the group III-VI XMAY2 family exhibits a potentially rich scope of applications in the field of piezoelectricity.

1. Introduction

Flexible materials with large piezoelectric coefficients are commonly utilized in the advancement of advanced wearable nanosensors [1]. These high-performance piezoelectric materials efficiently transform mechanical energy into electrical energy. Consequently, they have found extensive application in nanomotors, actuation and energy harvesting sectors [2,3,4]. Since the discovery of piezoelectric properties in a MoS2 monolayer [5], various two-dimensional materials have captured attention, such as graphene [6,7,8], phosphorene [9,10,11], germanene [12,13,14], transition metal dichalcogenides [15,16,17], hexagonal boron nitride [18,19], group IV monochalcogenides [20,21,22,23] and other 2D materials [24,25,26,27,28]. The experimentally measured in-plane piezoelectric coefficient e11 of the MoS2 monolayer is in good agreement with the previous theoretical results, which shows the reliability of theoretical predictions [29]. Two-dimensional materials exhibit fine mechanical properties, with many flexible materials also demonstrating good piezoelectric properties. This suggests that mechanical properties play a crucial role in determining piezoelectric response. Two-dimensional materials possessing both in-plane and out-of-plane piezoelectric coefficients show promise for the development of energy harvesting and conversion devices. Despite this potential, the majority of 2D piezoelectric materials exhibit only in-plane piezoelectricity, restricting their usage to nanoscale devices with specific needs, like vertically integrated nanogenerator systems.
Therefore, 2D materials with out-of-plane piezoelectricity (as characterized by coefficient d31) are being explored, especially the coexistence of out-of-plane and in-plane piezoelectricity [30]. To better explore piezoelectric properties, Janus structures are constructed which can enhance both the in-plane and out-of-plane polarization of materials by breaking mirror symmetry and inversion symmetry. Experimental synthesis of the Janus MoSSe monolayer has validated the effectiveness of this approach [31]. Additionally, two-dimensional materials with out-of-plane piezoelectric coefficients may exhibit a negative longitudinal piezoelectric (NLP) response, indicating promising prospects for enhancing piezoelectric sensor technology [32,33,34,35].
Recently, a new Janus XMAY2 family material has been extensively researched for its properties [36,37,38,39]. Hong et al. synthesized a two-dimensional MoSi2N4 monolayer material using the chemical vapor deposition method and introduced the MA2Y4 family materials [40,41,42]. Sibator et al. [38] proposed, for the first time, asymmetric XMoSiN2 (X = S, Se, Te) monolayers by replacing the Si-N layer on one side of MoSi2N4 with an X (X = S, Se, Te) layer belonging to the XMAY2 family where letter X represents group VI element (S, Se), letter A represents group IV element (Si, Ge) and letter Y represents group V element (N, P, As). Liao et al. [43] reported the coexistence of in-plane (d11) and out-of-plane (d31) piezoelectric coefficients in XSSiN2 monolayers, indicating that the XMAY2 system may have excellent in-plane and out-of-plane piezoelectric properties.
In this paper, Janus XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) monolayers were proposed, which belong to the group III-VI XMAY2 family. Using first-principles calculations, we observed that IZrGaS2 monolayers exhibit the coexistence of the largest in-plane and out-of-plane piezoelectric coefficients among the XMAY2 family materials. X, A and Y represent group VII elements (iodine), group III elements (Al, Ga) and group VI elements (S, Se) in the group III-VI XMAY2 family, respectively. Compared to the previous XMAY2 family materials, the modifications in X, A and Y make an improvement in the out-of-plane piezoelectric coefficients. Furthermore, the stability, electrical, mechanical and transport properties of the XMAY2 monolayers were calculated and analyzed. The exceptional piezoelectric properties of XMAY2 monolayers indicate promising potential in the field of piezoelectricity.

2. Methods

This study utilizes density functional theory (DFT) stimulation with the projected argument wave (PAW) method implemented in the VASP package [44,45]. The wave function is expanded using a plane wave with a cutoff energy [46] of 500 eV. The Perdew–Burke–Ernzerhof (PBE) functional within the generalized gradient approximation (GGA) is employed for the exchange–correlation functional [47]. According to the intrinsic lattice feature of XMAY2 (M = Ti, Zr; A = Al, Ga; Y = S, Se) monolayers, the Brillouin region is divided into a grid of k-points using a 16 × 16 × 1 mesh, following the Monkhorst–Pack strategy. To ensure structural stability, the energy and force difference criteria of the plane wave are set at 10−8 eV and 0.01 eV/Å, respectively. Due to the thickness of the XAMY2 monolayers being about 6 Å with five atomic layers, it is necessary to consider the weak van der Waals interactions. Here, the semi-empirical DFT-D3 method is adopted [48]. Additionally, a vacuum layer with a minimum thickness of 30 Å in the z-direction is included to prevent interactions between neighboring layers. Dynamic stability is verified by expanding the primitive cell into a supercell using Phonopy [49] and calculating the phonon spectrum with density functional perturbation theory (DFPT) [50,51]. Band structure calculations are performed using both PBE and HSE06 functionals [52]. Piezoelectric coefficients and elastic stiffness coefficients are calculated using DFPT and energy–strain methods [53], respectively.

3. Results and Discussions

3.1. Crystal Lattice and Stability

The optimized XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) monolayers are depicted in Figure 1, with the side view shown in Figure 1a and the top view in Figure 1b. The blue and yellow areas represent the primitive cell and rectangle cell, respectively. The XMAY2 monolayer consists of four elements and five atomic layers, which belong to the P3m1 space group. The bond connection in the XMAY2 monolayers follows the pattern of Y-A-Y-M-X. Table 1 summarizes the optimized lattice constants, thickness and bond lengths. Additionally, the bond angle θ (Å) of XMAY2 monolayers is summarized in Table S1 and Figure S1.
In order to calculate the stability of XMAY2 monolayers, the cohesive energy Ecoh was calculated using the following formula:
E c o h = E X M A Y 2 E X E M E A 2 E Y N
where E X M A Y 2 is the total energy of XMAY2 monolayers, E X , E M , E A and E Y are the energies of single X, M, A and Y atoms, respectively. N is the number of atoms in the primitive cell of XMAY2 monolayers. The formation energy is calculated using the following formula:
E c o h = E X M A Y 2 E X E M E A 2 E Y N
where E X M A Y 2 is the total energy of the XMAY2 monolayers and μ X , μ M , μ A and μ Y are the energies of X, M, A and Y in the bulk stable phase. The cohesive energy Ecoh and the formation energy Eform value of XMAY2 monolayers are recorded in Table 1, respectively. The cohesive energy of XMAY2 monolayers is smaller than that of most previously reported two-dimensional monolayer materials, such as Al2TeSeS monolayer (−2.89 eV per atom), Cu2Si (−3.46 eV per atom), silene (−3.98 eV per atom) and germanene (−3.26 eV per atom) [54,55]. The negative values of formation energy and cohesive energy indicate that XMAY2 monolayers provide thermodynamic support for these materials compared to other materials in terms of energy.
Phonon spectrum calculations are performed to validate the dynamic stability of the XMAY2 monolayers. Figure 2 indicates the existence of a small virtual frequency near the Γ point, due to numerical error. These errors can be addressed using the Acoustic Summation Rule (ASR) and therefore can be ignored [56]. With five atoms in the XMAY2 monolayers, each phonon dispersion comprises 15 vibration branches, including 3 low-frequency acoustic branches and 12 high-frequency optical branches. Specifically, at the Γ point, the phonon mode of the XMAY2 monolayers can be decomposed into GC3v = 5A1 + 5E, where A1/E represent the non-degenerate/doubly degenerate phonon modes at the Γ point, where correspond to out-of-plane optical (ZO) modes/in-plane transverse (TO) and longitudinal optical (LO) phonon modes [57]. The coexistence of acoustic and optical branches in the phonon spectrum validates the dynamic stability of the XMAY2 monolayers.

3.2. Piezoelectric Properties

The elastic stiffness coefficient Cij is the crucial measure of a material’s response to external stress and strain, reflecting its mechanical characteristics. The relaxed-ion [5] elastic stiffness coefficient is determined by the atom completely relaxed when the strain is applied, underscoring the electronic and ionic common contributions. Conversely, the clamped-ion [58] elastic stiffness coefficient is calculated by maintaining atomic positions after applying strain, highlighting the electronic contributions. Piezoelectricity is the interaction between external mechanical stress and intrinsic electrical polarization. The piezoelectric coefficients of clamped-ion represent the contribution of electronic parts, while relaxed-ion accounts for the combined effect of ionic and electronic parts.
Following on existing research, we calculate the in-plane elastic stiffness coefficient Cij within the material as follows:
C i j = 1 S 0 E 2 ε i ε j  
where E , S0, ε i and ε j represent total energy, the area of the primitive cell and the strains along the x and y directions, respectively. The Young’s modulus and Poisson’s ratio of the XMAY2 monolayers can be calculated using the relaxed-ion elastic stiffness coefficient as follows [59]:
Y 2 D = C 11 2 C 12 2 C 11 , ν 2 D = C 12 C 11
In the case of 2D materials, the relaxed-ion piezoelectric coefficient directly reflects the efficiency of electro-mechanical coupling. The piezoelectric stress coefficient ejk and piezoelectric strain coefficient dij are essential coefficients for piezoelectric properties. According to the definition, the relaxed-ion piezoelectric stress coefficient is sum of electrons and ions. The piezoelectric strain coefficient dij can be calculated with the following formula:
e j k = d i j · C j k  
The clamped-ion piezoelectric stress coefficient e11 is the polarization value contributed by the electrons, while the relaxed-ion piezoelectric stress coefficient e11 is the polarization value contributed by the electrons and ions. The piezoelectric strain coefficient d11 and d31 can be expressed as follows:
d 11 = e 11 C 11 C 12 ,   d 31 = e 31 C 11 + C 12
According to eq6, we can obtain the relaxed-ion and the clamped-ion piezoelectric strain coefficient d11 and d31. Due to the symmetry of the XMAY2 monolayers e11 = −e12 and d11 = −d12. The calculation results are summarized in Table 2. For hexagonal 2D materials, the relaxed-ion elastic coefficients can better reflect the mechanical properties of the materials. The calculated elastic coefficients obey the Bonn–Huang stability criteria (C11C12 > 0 and C66 > 0).
For XMAY2 monolayers, the C11 values range from 73.51 to 106.58 N/m, comparable to MoS2 (130 N/m) [5] and graphene (336 N/m) [60]. The C11 value of the XMAY2 monolayers is equivalent to C22, while C12 is approximately a quarter of C11. As shown in Figure S3, the relationship between the Young’s modulus and Poisson’s ratio of XMAY2 monolayers exhibit isotropy. The Young’s modulus of XMAY2 monolayers varies from 74.88 to 104.63 N/m, with ITiAlS2 being the highest and IZrGaSe2 being the lowest. The elastic constants of XMAY2 monolayers are closely linked to the bond length dTi-S < dTi-Se < dZr-S < dZr-Se. A shorter bond length results in a stronger influence between nearby atoms, making the materials show stiffness. Conversely, longer bond lengths lead to more flexibility among atoms. The Young’s modulus of the XMAY2 monolayers is smaller than that of common two-dimensional materials, such as Janus MoSSe (Y2D = 113.61 N/m) monolayer, graphene (Y2D = 341 N/m) [61] and h-BN (Y2D = 275.9 N/m) [59]. XMAY2 monolayers exhibit good flexibility, making them suitable for applications in flexible skin materials. Furthermore, the Poisson’s ratio of the XMAY2 monolayers ranges from 0.21 to 0.25, which is similar to the previously reported XMoSiP2 (0.22−0.24) [62] monolayers with the same five-atom layers. Flexible mechanical properties indicate that the XMAY2 monolayers may possess favorable piezoelectric properties and hold significant potential in the development of wearable flexible sensor devices.
The in-plane relaxed-ion piezoelectric strain coefficients d11 in 2D materials directly indicate the distortion-coupling ability of materials. Table 2 illustrates that the relaxed-ion in-plane piezoelectric strain coefficients d11 values range from 5.20 to 7.04 pm/V and the relaxed-ion out-of-plane piezoelectric strain coefficients d31 values range from −0.23 to 0.48 pm/V. The clamped-ion elastic coefficient and piezoelectric coefficients are summarized in Table S3. For the XMAY2 monolayers, the clamped-ion and the relaxed-ion piezoelectric stress coefficient e11 values range from 431 to 562 pC/m and 291 to 440 pC/m, respectively. It is evident that the clamped-ion piezoelectric stress coefficients exceed the relaxed-ion piezoelectric stress coefficients for the same XMAY2 monolayers. This difference is attributed to the competition between ionic and electronic polarization, resulting in a smaller piezoelectric stress coefficient e11 for the XMAY2 monolayers. For a given M element, the values of e11 are related to the work function at the bottom, which is determined by the electrical properties of atoms A and Y. The larger the work function, the greater the values of e11. Additionally, the low elastic coefficients C11 and C12 of the XMAY2 monolayers contribute to its flexibility, ultimately leading to a larger piezoelectric coefficient for the XMAY2 monolayers. In the XMAY2 monolayers, the elastic stiffness coefficients (C11 − C12) and (C11 + C12) were analyzed, with values ranging from 52.50 to 80.09 N/m and from 92.52 to 133.07 N/m for C11 − C12 and C11 + C12, respectively. Due to the small elastic stiffness coefficients (C11 − C12) and (C11 + C12), as well as the large piezoelectric stress coefficient, large in-plane and expected out-of-plane piezoelectricity can be observed.
The piezoelectric stress coefficient e31 values of XMAY2 monolayers range from −23 to 52 pC/m. Upon inspecting the absolute values, it is evident that, except for the ITiAlS2 monolayer, all other XMAY2 monolayers demonstrate large piezoelectric stress coefficients (>20 pC/m). Notably, the IZrGaS2 monolayer stands out, with a value as high as 52 pC/m, which contributes to its large out-of-plane piezoelectric coefficients within the group III-VI XMAY2 family. In comparison to the IZrGaS2 and IZrAlS2 monolayers, both exhibit larger values of the piezoelectric stress coefficient e31. However, due to a larger dS-Ga compared to dS-Al, the elastic constants C11 + C12 of IZrGaS2 are smaller, ultimately leading to the piezoelectric strain coefficient d31 of IZrGaS2 being twice that of IZrAlS2. Additionally, despite the relatively small piezoelectric stress coefficients of IZrGaSe2 and IZrAlSe2 monolayers, their elastic coefficients C11 + C12 are also small, resulting in their out-of-plane piezoelectric coefficients (d31) being comparable to those of IZrAlS2. This highlights the significant influence of mechanical properties on the piezoelectric properties of materials, with flexible materials showing promise as good piezoelectric materials.
Among them, the IZrGaS2 monolayer exhibits the coexistence of large in-plane and out-of-plane piezoelectric coefficients at 7.03 pm/V and 0.48 pm/V, respectively. The values of d31 are similar to those of common two-dimensional materials like Janus HfSeTe (d31 = 0.41 pm/V) and Janus group III monochalcogenides (d31 = 0.07–0.46 pm/V) [63,64,65]. The large piezoelectric properties of the IZrGaS2 monolayer are attributed to several factors: first, the large potential gradient indicates a large difference between the work function between the bottom and top, forming a large built-in electric field inside the materials which contributes to the large piezoelectric strain coefficients d11 and d31. Furthermore, the IZrGaS2 monolayer exhibits a smaller elastic coefficient, indicating that the material generates greater strain when subjected to the same external stress, resulting in a larger piezoelectric coefficient. A higher piezoelectric coefficient means that converting mechanical energy into electrical energy promotes the application of piezoelectric materials in energy harvesting. Meanwhile, smaller elastic constants and larger piezoelectric strain coefficients also mean that when piezoelectric materials are used as pressure sensors, voltage will be generated, resulting in higher detection accuracy. At present, it is difficult for common piezoelectric materials to be suitable for use as flexible materials due to their brittleness. Therefore, piezoelectric materials with smaller elastic constants have a wider range of application scenarios. The IZrGaS2 monolayer positions as a highly promising two-dimensional piezoelectric material with the coexistence of large in-plane and out-of-plane piezoelectric coefficients in the XMAY2 family.

3.3. Electronic Properties

To further analyze XMAY2 monolayers, the electronic properties are calculated. The band structure is calculated using PBE and HSE06 functionals, as shown in Figure 3. The bandgap energy values of XMAY2 monolayers range from 0.15 eV to 0.60 eV using the PBE functional. It is widely recognized that the PBE functional generally underestimates the bandgap, while the bandgap calculated by the HSE06 functional has acceptable accuracy. The bandgap energy calculated using the HSE06 functional range from 0.38 to 0.67 eV. The bandgap results calculated using both PBE and HSE06 functionals are summarized in Table 3. For the HSE06 functional, it was observed that the bandgaps of the ITiAlS2, ITiAlSe2, ITiGaS2 and ITiGaSe2 monolayers are direct bandgap semiconductors and the VBM and CBM of the ITiAlS2, ITiAlSe2, ITiGaS2 and ITiGaSe2 monolayers are located at K point. The IZrAlS2, IZrAlSe2, IZrGaS2 and IZrGaSe2 monolayers are indirect band gap semiconductors. The VBM of the IZrAlS2, IZrGaS2 and IZrGaSe2 monolayers are located at M point, while the VBM of IZrAlSe2 were at point Γ. The CBMs of the IZrAlS2, IZrAlSe2, IZrGaS2 and IZrGaSe2 monolayers are situated at K point. The projected band structure of XMAY2 monolayers is in Supplementary Materials Figures S4 and S5, using the HSE06 functional. The CBM and VBM of the ITiAlS2, ITiAlSe2, ITiGaS2 and ITiGaSe2 monolayers represent the d-orbital contribution of the Ti element. The CBM and VBM of the ITiAlS2, ITiAlSe2, ITiGaS2 and ITiGaSe2 monolayers represent the p-orbital contribution of the Al/Ga element and the d-orbital contribution of the Ti element, respectively. It is worth noting that the band gap of the stable XMAY2 monolayer exceeds 0.1 eV, which is the basis of its piezoelectric properties.
Due to the difference in electronegativity between X, A and Y atoms in the Janus XMAY2 monolayer, the bond lengths between atoms are different, resulting in the breaking of mirror symmetry and the generation of an internal electric field Ein. This internal electric field creates potential differences along the z-axis direction, necessitating a dipole correction when calculating the planar average potential of the two-dimensional Janus structure [66,67]. As shown in Figure 4, the work function of the XMAY2 monolayers is denoted as Φ = EvacEfermi, where Evac represents the vacuum energy level and Efermi is the Fermi level. The work function and vacuum energy level difference between the top (Φ2) and bottom (Φ1) of the XMAY2 monolayers are summarized in Table 3. A vacuum energy level difference between the top and bottom ranging from −0.47 to 0.31 eV was observed, establishing an intrinsic built-in electric field in the XMAY2 monolayers that significantly affects the behavior of electrons and holes. ITiGaS2, ITiGaSe2, IZrGaS2 and ITiGaSe2 exhibit that the top side potential is higher than the bottom side, with the electric field direction from top to bottom promoting electron (hole) transition from the Y (X) layer to the X (Y) layer. Conversely, ITiAlS2, ITiAlSe2, IZrAlS2 and ITiAlSe2 exhibit that the bottom side potential is higher than the top side, with the electric field direction moving from the bottom to the top, facilitating electron (hole) transition from the X (Y) layer to the Y (X) layer. The highest built-in electric field Ein of the IZrGaS2 monolayer is 0.31eV, suggesting a possible correlation between the out-of-plane piezoelectric coefficient and the built-in electric field. Furthermore, charge density difference calculation was performed and analyzed using Bader Charge analysis [66]. (Supplementary Materials Table S2 and Figure S2). Both X and Y atoms acquire electrons, while M and A atoms lose electrons. The out-of-plane polarization is related to the values of electrons lost by the A atom and acquired by the Y atom.

3.4. Transport Properties

The carrier mobility, especially electronic, is vital for comprehending the transport properties of materials. The carrier mobilities of the XMAY2 monolayers were analyzed using the deformation potential (DP) theory [68]. The carrier mobility of two-dimensional materials was calculated as follows:
μ 2 D = e 3 C 2 D κ B T m * m ¯ E d 2
where e , 3 , C 2 D ,   κ B ,   T ,   m * ,   m ¯ and E d represent the electron charge, Planck constant, elastic modulus, Boltzmann constant, temperature, effective mass, average effective mass ( m ¯ = m x * m y * ) and deformation potential constant, respectively. The effective mass m* of electrons and holes in two-dimensional materials are important parameter for calculating carrier mobility, which can be obtained through the parabolic functions of CBM and VBM, respectively.
1 m * = 1 2 2 E ( k ) k 2
where E (k) is the energy corresponding to the wave vector k at the edge CBM and VBM positions. Therefore, the size of the effective mass m* is related to the slope near CBM and VBM. A uniaxial strain of −0.4%~0.4% is applied to the x and y directions of the rectangle lattice of the XMAY2 monolayers with an interval of 0.002 to fit the elastic modulus C2D and the deformation potential constant Ed is expressed as follows, respectively.
C 2 D = 1 S 0   2 E ε 2 ,   E d = E e d g e ε
where E is the total energy, S0 is the area of the optimized unit cell, ε indicates the uniaxial strain along the x and y transport directions and E e d g e is the energy change in band edge of electrons and holes. From the calculated m*, C 2 D and E d , the carrier mobility μ 2 D of electrons and holes, can be calculated using Equation (7).
Materials with high electronic carrier mobility are believed to be applied for application in nanoelectronics devices. As shown in Table 4, the values of the elastic modulus C2D are almost the same in the x and y directions and independent of the direction of uniaxial strain applied, which is attributed to the isotropy of the mechanical properties of the XMAY2 monolayers. The deformation potential constant of electrons in the XMAY2 monolayers is approximated in both the x and y directions and the effective mass of electrons shows differences in the x and y directions. Consequently, there is a notable difference in carrier mobility between electrons in the x and y directions, with higher mobility observed in the y direction. This is consistent with the lower effective mass of the structure along the y direction compared with the x direction.
Specifically, the high electronic carrier mobilities of ITiAlS2 and ITiAlSe2 in the y direction are 9546.23 and 9889.53 cm2 s−1 V−1, respectively. Additionally, the carrier mobilities of the IZrAlS2 and IZrAlSe2 monolayers in the y-direction are 4364.81 and 5728.33 cm2 s−1 V−1, respectively. Notably, the carrier mobility of the XMAY2 monolayers surpasses that of common two-dimensional materials by one or two orders of magnitude, such as MoS2 (~200 cm2 s−1 V−1) [69] and GeP3 (~190 cm2 s−1 V−1) [17]. The electronic carrier mobility of ITiAlS2 and ITiAlSe2 is comparable to the newly discovered five-atom layers structures (Al2STeSe ~10,536 cm2 s−1 V−1 and STiGeAs2 ~8175.66 cm2 s−1 V−1) [36,54]. The high electronic carrier mobility observed in XMAY2 monolayers, including ITiAlS2, ITiAlSe2, IZrAlS2, IZrAlSe2 and IZrGaSe2, makes them promising for application in the development of electronic devices.

4. Conclusions

In summary, we systematically studied the stability and the electrical properties of 2D asymmetric XMAY2 (X = I; M = Ti, Zr; A = Ga, In; Y = S, Se) monolayers. The cohesive energy, formation energy and phonon spectrum indicate that these structures are both thermodynamically and dynamically stable. All the XMAY2 monolayers demonstrate flexible mechanical properties, with Young’s modulus values ranging from 73.51 to 106.58 N/m. The in-plane (d11) and out-of-plane (d31) piezoelectric coefficients of XMAY2 monolayers range from 5.20 to 7.04 pm/V and from −0.23 to 0.48 pm/V, respectively. The coexistence of in-plane and out-of-plane piezoelectric properties in these systems may be attributed to differences in atomic electronegativity and symmetry breaking. Additionally, the band structure of XMAY2 monolayers is calculated using the HSE06 function, with bandgap values ranging from 0.34 to 0.67. ITiAlS2, ITiAlSe2, ITiGaS2 and ITiGaSe2 are direct bandgap semiconductors, while the others are indirect bandgap semiconductors. Furthermore, the XMAY2 monolayers showcase remarkable carrier mobility, particularly with the ITiAlS2 and ITiAlSe2 monolayers having high electronic carrier mobility (9546.23 and 9889.53 cm2 s−1 V−1). This study provides valuable insights into the potential applications of the XMAY2 monolayers in the nanopiezoelectric field of devices and sensors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14080708/s1, Figure S1: Side view of XMAY2 monolayers. Figure S2: Top and side views of charge density difference for XMAY2 monolayers (a-h). The red number are the results of Bader analysis. Figure S3: Young’s modulus (N/m) (a) and Poisson’s ratio (b) of XMAY2 monolayers. Figure S4: The projected band structure of ITiAlS2 (a), ITiAlSe2 (b), ITiGaS2 (c) and ITiGaSe2 (d) monolayers. The blue dots represent the d-orbital contribution of Ti element. Figure S5: The projected band structure of IZrAlS2 (a), IZrAlSe2 (b), IZrGaS2 (c) and IZrGaSe2 (d) monolayers. The blue dots and orange dots represent the d-orbital contribution of Ti element and the p-orbital contribution of Al/Ga element, respectively; Table S1: The bond angle θ (Å) of XMAY2 monolayers. Table S2: The results of Bader charge analysis for the charge transfer of XMAY2 monolayers. A negative (positive) value indicates that the mentioned atom loses (gains) electrons. Table S3. Elastic coefficient, piezoelectric stress coefficients and piezoelectric strain coefficients of clamped-ion XMAY2 monolayers.

Author Contributions

Conceptualization, Z.L.; data curation, Z.L. and Y.Z.; formal analysis, H.D.; funding acquisition, X.G.; investigation, Z.L. and Y.Z.; methodology, J.P.; project administration, H.D.; software, Z.L. and X.G.; supervision, H.D.; writing—original draft, Z.L.; writing—review and editing, Z.L., J.P. and H.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Grants No. 11604056 and 11804057).

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding author.

Acknowledgments

We thank the Center of Campus Network & Modern Educational Technology, Guangdong University of Technology, Guangdong, China for providing computational resources and technical support for this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yi, Z.; Liu, Z.; Li, W.; Ruan, T.; Chen, X.; Liu, J.; Yang, B.; Zhang, W. Piezoelectric Dynamics of Arterial Pulse for Wearable Continuous Blood Pressure Monitoring. Adv. Mater. 2022, 34, 2110291. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Z.L.; Song, J.H. Piezoelectric nanogenerators based on zinc oxide nanowire arrays. Science 2006, 312, 242–246. [Google Scholar] [CrossRef]
  3. Yin, H.B.; Zheng, G.P.; Wang, Y.X.; Yao, B.J. New monolayer ternary In-containing sesquichalcogenides BiInSe3, SbInSe3, BiInTe3, and SbInTe3 with high stability and extraordinary piezoelectric properties. Phys. Chem. Chem. Phys. 2018, 20, 19177–19187. [Google Scholar] [CrossRef] [PubMed]
  4. Hinchet, R.; Khan, U.; Falconi, C.; Kim, S.-W. Piezoelectric properties in two-dimensional materials: Simulations and experiments. Mater. Today 2018, 21, 611–630. [Google Scholar] [CrossRef]
  5. Duerloo, K.-A.N.; Ong, M.T.; Reed, E.J. Intrinsic Piezoelectricity in Two-Dimensional Materials. J. Phys. Chem. Lett. 2012, 3, 2871–2876. [Google Scholar] [CrossRef]
  6. 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. [Google Scholar] [CrossRef] [PubMed]
  7. Koskinen, P.; Malola, S.; Häkkinen, H. Self-passivating edge reconstructions of graphene. Phys. Rev. Lett. 2008, 101, 115502. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, Y.B.; Tan, Y.W.; Stormer, H.L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438, 201–204. [Google Scholar] [CrossRef]
  9. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black phosphorus field-effect transistors. Nat. Nanotechnol. 2014, 9, 372–377. [Google Scholar] [CrossRef]
  10. Liu, H.; Neal, A.T.; Zhu, Z.; Luo, Z.; Xu, X.; Tomanek, D.; Ye, P.D. Phosphorene: An Unexplored 2D Semiconductor with a High Hole Mobility. ACS Nano 2014, 8, 4033–4041. [Google Scholar] [CrossRef]
  11. Carvalho, A.; Wang, M.; Zhu, X.; Rodin, A.S.; Su, H.B.; Neto, A.H.C. Phosphorene: From theory to applications. Nat. Rev. Mater. 2016, 1, 16061. [Google Scholar] [CrossRef]
  12. Li, L.F.; Lu, S.Z.; Pan, J.B.; Qin, Z.H.; Wang, Y.Q.; Wang, Y.L.; Cao, G.Y.; Du, S.X.; Gao, H.J. Buckled Germanene Formation on Pt(111). Adv. Mater. 2014, 26, 4820–4824. [Google Scholar] [CrossRef] [PubMed]
  13. Dávila, M.E.; Xian, L.; Cahangirov, S.; Rubio, A.; Le Lay, G. Germanene: A novel two-dimensional germanium allotrope akin to graphene and silicene. New J. Phys. 2014, 16, 095002. [Google Scholar] [CrossRef]
  14. Chegel, R.; Behzad, S. Tunable Electronic, Optical, and Thermal Properties of two- dimensional Germanene via an external electric field. Sci. Rep. 2020, 10, 704. [Google Scholar] [CrossRef]
  15. Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O.V.; Kis, A. 2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. [Google Scholar] [CrossRef]
  16. Castro Neto, A.H. Charge density wave, superconductivity, and anomalous metallic behavior in 2D transition metal dichalcogenides. Phys. Rev. Lett. 2001, 86, 4382–4385. [Google Scholar] [CrossRef]
  17. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef]
  18. Hofmann, M.; Shin, Y.C.; Hsieh, Y.P.; Dresselhaus, M.S.; Kong, J. A facile tool for the characterization of two-dimensional materials grown by chemical vapor deposition. Nano Res. 2012, 5, 504–511. [Google Scholar] [CrossRef]
  19. Song, L.; Ci, L.J.; Lu, H.; Sorokin, P.B.; Jin, C.H.; Ni, J.; Kvashnin, A.G.; Kvashnin, D.G.; Lou, J.; Yakobson, B.I.; et al. Large Scale Growth and Characterization of Atomic Hexagonal Boron Nitride Layers. NANO Lett. 2010, 10, 3209–3215. [Google Scholar] [CrossRef]
  20. Gomes, L.C.; Carvalho, A.; Neto, A.H.C. Enhanced piezoelectricity and modified dielectric screening of two-dimensional group-IV monochalcogenides. Phys. Rev. B 2015, 92, 214103. [Google Scholar] [CrossRef]
  21. Gomes, L.C.; Carvalho, A. Phosphorene analogues: Isoelectronic two-dimensional group-IV monochalcogenides with orthorhombic structure. Phys. Rev. B 2015, 92, 085406. [Google Scholar] [CrossRef]
  22. Gomes, L.C.; Carvalho, A. Electronic and optical properties of low-dimensional group-IV monochalcogenides. J. Appl. Phys. 2020, 128, 121101. [Google Scholar] [CrossRef]
  23. Michel, K.H.; Çakir, D.; Sevik, C.; Peeters, F.M. Piezoelectricity in two-dimensional materials: Comparative study between lattice dynamics and ab initio calculations. Phys. Rev. B 2017, 95, 125415. [Google Scholar] [CrossRef]
  24. Haastrup, S.; Strange, M.; Pandey, M.; Deilmann, T.; Schmidt, P.S.; Hinsche, N.F.; Gjerding, M.N.; Torelli, D.; Larsen, P.M.; Riis-Jensen, A.C.; et al. The Computational 2D Materials Database: High-throughput modeling and discovery of atomically thin crystals. 2D Mater. 2018, 5, 042002. [Google Scholar] [CrossRef]
  25. Riis-Jensen, A.C.; Deilmann, T.; Olsen, T.; Thygesen, K.S. Classifying the Electronic and Optical Properties of Janus Monolayers. ACS Nano 2019, 13, 13354–13364. [Google Scholar] [CrossRef] [PubMed]
  26. Vo, D.D.; Vu, T.V.; Al-Qaisi, S.; Tong, H.D.; Le, T.S.; Nguyen, C.V.; Phuc, H.V.; Luong, H.L.; Jappor, H.R.; Obeid, M.M.; et al. Janus monolayer PtSSe under external electric field and strain: A first principles study on electronic structure and optical properties. Superlattices Microstruct. 2020, 147, 106683. [Google Scholar] [CrossRef]
  27. Almayyali, A.O.M.; Muhsen, H.O.; Merdan, M.; Obeid, M.M.; Jappor, H.R. Two-dimensional ZnI2 monolayer as a photocatalyst for water splitting and improvement its electronic and optical properties by strains. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 126, 114487. [Google Scholar] [CrossRef]
  28. Abdulameer, M.J.; Al-Abbas, S.S.A.; Jappor, H.R. Tuning optical and electronic properties of 2D ZnI2/CdS heterostructure by biaxial strains for optical nanodevices: A first-principles study. J. Appl. Phys. 2021, 129, 225104. [Google Scholar] [CrossRef]
  29. Xue, F.; Zhang, J.; Hu, W.; Hsu, W.-T.; Han, A.; Leung, S.-F.; Huang, J.-K.; Wan, Y.; Liu, S.; Zhang, J.; et al. Multidirection Piezoelectricity in Mono- and Multilayered Hexagonal α-In2Se3. ACS Nano 2018, 12, 4976–4983. [Google Scholar] [CrossRef]
  30. Liu, S.; Chen, W.; Liu, C.; Wang, B.; Yin, H. Coexistence of large out-of-plane and in-plane piezoelectricity in 2D monolayer Li-based ternary chalcogenides LiMX2. Results Phys. 2021, 26, 104398. [Google Scholar] [CrossRef]
  31. Zhang, J.; Jia, S.; Kholmanov, I.; Dong, L.; Er, D.; Chen, W.; Guo, H.; Jin, Z.; Shenoy, V.B.; Shi, L.; et al. Janus Monolayer Transition-Metal Dichalcogenides. ACS Nano 2017, 11, 8192–8198. [Google Scholar] [CrossRef]
  32. Liu, S.; Cohen, R.E. Origin of Negative Longitudinal Piezoelectric Effect. Phys. Rev. Lett. 2017, 119, 207601. [Google Scholar] [CrossRef] [PubMed]
  33. You, L.; Zhang, Y.; Zhou, S.; Chaturvedi, A.; Morris, S.A.; Liu, F.; Chang, L.; Ichinose, D.; Funakubo, H.; Hu, W.; et al. Origin of giant negative piezoelectricity in a layered van der Waals ferroelectric. Sci. Adv. 2019, 5, eaav3780. [Google Scholar] [CrossRef]
  34. Wang, Z.; Dong, S. Large in-plane negative piezoelectricity and giant nonlinear optical susceptibility in elementary ferroelectric monolayers. Phys. Rev. B 2023, 108, 235423. [Google Scholar] [CrossRef]
  35. Zhong, S.; Zhang, X.; Liu, S.; Yang, S.A.; Lu, Y. Giant and Nonanalytic Negative Piezoelectric Response in Elemental Group-Va Ferroelectric Monolayers. Phys. Rev. Lett. 2023, 131, 236801. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, Z.; He, X.; Li, W.; He, Y.; Xiong, K. First principles prediction of two-dimensional Janus STiXY2 (X = Si, Ge; Y = N, P, As) materials. Dalton Trans. 2023, 52, 8322–8331. [Google Scholar] [CrossRef]
  37. Ding, C.-H.; Duan, Z.-F.; Ding, Z.-K.; Pan, H.; Wang, J.; Xiao, W.-H.; Liu, W.-P.; Li, Q.-Q.; Luo, N.-N.; Zeng, J.; et al. XMoSiN2 (X = S, Se, Te): A novel 2D Janus semiconductor with ultra-high carrier mobility and excellent thermoelectric performance. Europhys. Lett. 2023, 143, 16002. [Google Scholar] [CrossRef]
  38. Sibatov, R.T.; Meftakhutdinov, R.M.; Kochaev, A.I. Asymmetric XMoSiN2 (X=S, Se, Te) monolayers as novel promising 2D materials for nanoelectronics and photovoltaics. Appl. Surf. Sci. 2022, 585, 152465. [Google Scholar] [CrossRef]
  39. Li, Z.; Luo, J.; Zhou, Y.; Chen, J.; Ling, H.; Zeng, J.; Yang, Y.; Dong, H. Asymmetric XMoGeY2 (X = S, Se, Te; Y = N, P, As) monolayers as potential flexible materials for nano piezoelectric devices and nanomedical sensors. Phys. Chem. Chem. Phys. 2024, 26, 12133–12141. [Google Scholar] [CrossRef]
  40. Hong, Y.L.; Liu, Z.B.; Wang, L.; Zhou, T.Y.; Ma, W.; Xu, C.; Feng, S.; Chen, L.; Chen, M.L.; Sun, D.M.; et al. Chemical vapor deposition of layered two-dimensional MoSi2N4 materials. Science 2020, 369, 670–674. [Google Scholar] [CrossRef]
  41. Wang, X.; Ju, W.; Wang, D.; Li, X.; Wan, J. Flexible MA2Z4 (M = Mo, W.; A = Si, Ge and Z = N, P, As) monolayers with outstanding mechanical, dynamical, electronic, and piezoelectric properties and anomalous dynamic polarization. Phys. Chem. Chem. Phys. 2023, 25, 18247–18258. [Google Scholar] [CrossRef] [PubMed]
  42. Vi, V.T.T.; Linh, T.P.T.; Nguyen, C.Q.; Hieu, N.N. Tunable Electronic Properties of Novel 2D Janus MSiGeN4 (M = Ti, Zr, Hf) Monolayers by Strain and External Electric Field. Adv. Theory Simul. 2022, 5, 2200499. [Google Scholar] [CrossRef]
  43. Liao, J.; Ma, X.; Yuan, G.; Xu, P.; Yuan, Z. Coexistence of in- and out-of-plane piezoelectricity in Janus XSSiN2 (X = Cr, Mo, W) monolayers. Appl. Surf. Sci. 2023, 610, 155586. [Google Scholar] [CrossRef]
  44. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  45. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  46. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B Condens. Matter 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  47. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B Condens. Matter 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  48. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  49. Giannozzi, P.; de Gironcoli, S.; Pavone, P.; Baroni, S. Ab initio calculation of phonon dispersions in semiconductors. Phys. Rev. B Condens. Matter 1991, 43, 7231–7242. [Google Scholar] [CrossRef]
  50. Baroni, S.; Giannozzi, P.; Testa, A. Green’s-function approach to linear response in solids. Phys. Rev. Lett. 1987, 58, 1861–1864. [Google Scholar] [CrossRef]
  51. Vanderbilt; King, S. Electric polarization as a bulk quantity and its relation to surface charge. Phys. Rev. B Condens. Matter 1993, 48, 4442–4455. [Google Scholar] [CrossRef] [PubMed]
  52. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  53. Wu, X.; Vanderbilt, D.; Hamann, D.R. Systematic treatment of displacements, strains, and electric fields in density-functional perturbation theory. Phys. Rev. B 2005, 72, 035105. [Google Scholar] [CrossRef]
  54. Qi, C.; Yan, C.; Li, Q.; Yang, T.; Qiu, S.; Cai, J. Two-dimensional Janus monolayers Al2XYZ (X/Y/Z = S, Se, Te, X ≠ Y ≠ Z): First-principles insight into the photocatalytic and highly adjustable piezoelectric properties. J. Mater. Chem. C 2023, 11, 3262–3274. [Google Scholar] [CrossRef]
  55. Yang, L.-M.; Bačić, V.; Popov, I.A.; Boldyrev, A.I.; Heine, T.; Frauenheim, T.; Ganz, E. Two-Dimensional Cu2Si Monolayer with Planar Hexacoordinate Copper and Silicon Bonding. J. Am. Chem. Soc. 2015, 137, 2757–2762. [Google Scholar] [CrossRef] [PubMed]
  56. Luo, W.; Xiang, H. Two-Dimensional Phosphorus Oxides as Energy and Information Materials. Angew. Chem. Int. Ed. 2016, 55, 8575–8580. [Google Scholar] [CrossRef] [PubMed]
  57. Hieu, N.N.; Phuc, H.V.; Kartamyshev, A.I.; Vu, T.V. Structural, electronic, and transport properties of quintuple atomic Janus monolayers Ga2XS2 (X= O, S, Se, Te): First-principles predictions. Phys. Rev. B 2022, 105, 075402. [Google Scholar] [CrossRef]
  58. Bernardini, F.; Fiorentini, V.; Vanderbilt, D. Spontaneous polarization and piezoelectric constants of III-V nitrides. Phys. Rev. B 1997, 56, R10024. [Google Scholar] [CrossRef]
  59. Andrew, R.C.; Mapasha, R.E.; Ukpong, A.M.; Chetty, N. Mechanical properties of graphene and boronitrene. Phys. Rev. B 2012, 85, 125428. [Google Scholar] [CrossRef]
  60. Leenaerts, O.; Peelaers, H.; Hernández-Nieves, A.D.; Partoens, B.; Peeters, F.M. First-principles investigation of graphene fluoride and graphane. Phys. Rev. B 2010, 82, 195436. [Google Scholar] [CrossRef]
  61. 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. [Google Scholar] [CrossRef] [PubMed]
  62. Hiep, N.T.; Nguyen, C.Q.; Poklonski, N.A.; Duque, C.A.; Phuc, H.V.; Lu, D.V.; Hieu, N.N. Structural, electronic, and transport properties of Janus XMoSiP2 (X = S, Se, Te) monolayers: A first-principles study. J. Phys. D Appl. Phys. 2023, 56, 385306. [Google Scholar] [CrossRef]
  63. Dimple; Jena, N.; Rawat, A.; Ahammed, R.; Mohanta, M.K.; De Sarkar, A. Emergence of high piezoelectricity along with robust electron mobility in Janus structures in semiconducting Group IVB dichalcogenide monolayers. J. Mater. Chem. A 2018, 6, 24885–24898. [Google Scholar] [CrossRef]
  64. Noor-A-Alam, M.; Kim, H.J.; Shin, Y.-H. Dipolar polarization and piezoelectricity of a hexagonal boron nitride sheet decorated with hydrogen and fluorine. Phys. Chem. Chem. Phys. 2014, 16, 6575–6582. [Google Scholar] [CrossRef] [PubMed]
  65. Guo, Y.; Zhou, S.; Bai, Y.; Zhao, J. Enhanced piezoelectric effect in Janus group-III chalcogenide monolayers. Appl. Phys. Lett. 2017, 110, 163102. [Google Scholar] [CrossRef]
  66. Fu, C.-F.; Sun, J.; Luo, Q.; Li, X.; Hu, W.; Yang, J. Intrinsic Electric Fields in Two-dimensional Materials Boost the Solar-to-Hydrogen Efficiency for Photocatalytic Water Splitting. Nano Lett. 2018, 18, 6312–6317. [Google Scholar] [CrossRef] [PubMed]
  67. Bengtsson, L. Dipole correction for surface supercell calculations. Phys. Rev. B 1999, 59, 12301–12304. [Google Scholar] [CrossRef]
  68. Bardeen, J.; Shockley, W. Deformation Potentials and Mobilities in Non-Polar Crystals. Phys. Rev. 1950, 80, 72–80. [Google Scholar] [CrossRef]
  69. Kaasbjerg, K.; Thygesen, K.S.; Jacobsen, K.W. Phonon-limited mobility in n-type single-layer MoS2 from first principles. Phys. Rev. B 2012, 85, 115317. [Google Scholar] [CrossRef]
Figure 1. Top view (a) and side view (b) of the optimized structure of XMAY2 monolayers. The primitive and orthogonal cells used are shown by the blue and yellow areas, respectively.
Figure 1. Top view (a) and side view (b) of the optimized structure of XMAY2 monolayers. The primitive and orthogonal cells used are shown by the blue and yellow areas, respectively.
Crystals 14 00708 g001
Figure 2. Phonon spectra of XMAY2 monolayers (ah).
Figure 2. Phonon spectra of XMAY2 monolayers (ah).
Crystals 14 00708 g002
Figure 3. The band structures of the XMAY2 monolayers (ah); the dashed blue and red lines are the results calculated using the PBE and HSE06 functionals, respectively.
Figure 3. The band structures of the XMAY2 monolayers (ah); the dashed blue and red lines are the results calculated using the PBE and HSE06 functionals, respectively.
Crystals 14 00708 g003
Figure 4. Plane-averaged electrostatic potential of the XMAY2 monolayers (ah).
Figure 4. Plane-averaged electrostatic potential of the XMAY2 monolayers (ah).
Crystals 14 00708 g004
Table 1. Lattice constant a (Å), bond length d (Å), thickness h (Å), cohesive energy Ecoh and formation energy Eform of XMAY2 monolayers.
Table 1. Lattice constant a (Å), bond length d (Å), thickness h (Å), cohesive energy Ecoh and formation energy Eform of XMAY2 monolayers.
PhaseahdI-MdM-YdA-Y(1)dA-Y(2)EcohEform
ITiAlS23.646.322.822.502.202.32−4.43−0.97
ITiAlSe23.776.782.852.632.352.45−3.95−0.81
ITiGaS23.696.492.832.502.262.34−3.98−0.79
ITiGaSe23.796.842.862.622.412.47−3.67−0.69
IZrAlS23.706.572.932.612.202.33−4.61−1.09
IZrAlSe23.736.702.962.732.352.46−4.31−1.00
IZrGaS23.736.712.952.612.272.36−4.27−0.92
IZrGaSe23.847.062.972.722.422.48−3.96−0.82
Table 2. Elastic coefficient C11 and C12 (N/m), Young’s modulus Y2D (N/m), Poisson’s ratio ν2D, piezoelectric stress coefficients and piezoelectric strain coefficients of XMAY2 monolayers.
Table 2. Elastic coefficient C11 and C12 (N/m), Young’s modulus Y2D (N/m), Poisson’s ratio ν2D, piezoelectric stress coefficients and piezoelectric strain coefficients of XMAY2 monolayers.
PhaseC11C12C11 − C12C11 + C12Y2Dv2De11d11e31d31
ITiAlS2106.5826.4980.09133.07104.630.234.165.20−0.03−0.02
ITiAlSe289.5622.7766.79112.3385.640.253.975.95−0.22−0.19
ITiGaS298.7923.8174.98122.6095.490.234.405.860.160.13
ITiGaSe279.4218.5060.9297.9280.000.244.176.84−0.23−0.23
IZrAlS296.7126.9269.79123.6398.440.213.735.350.330.26
IZrAlSe278.0624.3653.7102.4297.810.203.015.620.200.19
IZrGaS281.5726.7254.85108.2986.590.213.867.040.520.48
IZrGaSe273.5121.0152.5094.5274.880.192.915.540.190.20
Table 3. Band gap Eg with the PBE and HSE06 functionals, the fermi energy EF, vacuum level difference ΔΦ and top side Φ2 and bottom side Φ1 work function of XMAY2 monolayers. All the units are eV.
Table 3. Band gap Eg with the PBE and HSE06 functionals, the fermi energy EF, vacuum level difference ΔΦ and top side Φ2 and bottom side Φ1 work function of XMAY2 monolayers. All the units are eV.
E g P B E E g H S E EFΦ1Φ2ΔΦ
ITiAlS20.480.67−0.422.943.17−0.23
ITiAlSe20.380.48−0.702.522.99−0.47
ITiGaS20.330.65−0.803.072.860.21
ITiGaSe20.600.44−1.032.702.630.07
IZrAlS20.230.48−1.422.953.18−0.23
IZrAlSe20.150.38−0.422.592.99−0.40
IZrGaS20.340.50−0.343.132.820.31
IZrGaSe20.160.34−0.742.782.660.12
Table 4. The carrier effective mass m* (m0), elastic modulus C2D (N m−1), deformation potential constant Ed (eV) and carrier mobility μ 2 D (cm2 V−1s−1) along the transport x and y directions for the XMAY2 monolayers. m0 is the mass of a free electron.
Table 4. The carrier effective mass m* (m0), elastic modulus C2D (N m−1), deformation potential constant Ed (eV) and carrier mobility μ 2 D (cm2 V−1s−1) along the transport x and y directions for the XMAY2 monolayers. m0 is the mass of a free electron.
PhaseTypemxmyC2DxC2DyEdxEdyμxμy
ITiAlS2Electron0.990.24108.5107.229.421.4353.979546.23
Hole0.510.52108.5107.225.075.30331.77313.24
ITiAlSe2Electron1.170.2290.0790.308.491.3244.849889.53
Hole0.500.5190.0790.305.125.37284.89263.48
ITiGaS2Electron1.030.4198.4298.258.118.0847.61120.30
Hole0.600.6398.4298.255.695.88166.87164.36
ITiGaSe2Electron2.921.9087.0787.266.871.148.78207.93
Hole0.600.6087.0787.265.035.29204.05184.08
IZrAlS2Electron1.060.22102.42102.529.212.1750.244364.81
Hole0.370.40102.42102.524.935.24583.83558.15
IZrAlSe2Electron1.320.1985.9686.649.651.8429.745728.33
Hole0.350.3785.9686.644.944.98567.92586.13
IZrGaS2Electron1.320.3192.2391.244.924.9846.7545.49
Hole0.410.4592.2391.245.765.85303.03325.94
IZrGaSe2Electron2.441.7875.6775.246.391.4545.4946.79
Hole0.350.4275.6775.245.004.80 7.761241.03
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Z.; Zhou, Y.; Guo, X.; Peng, J.; Dong, H. Coexistence of Large In-Plane and Out-of-Plane Piezoelectric Response in Group III–VI XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) Monolayers. Crystals 2024, 14, 708. https://doi.org/10.3390/cryst14080708

AMA Style

Li Z, Zhou Y, Guo X, Peng J, Dong H. Coexistence of Large In-Plane and Out-of-Plane Piezoelectric Response in Group III–VI XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) Monolayers. Crystals. 2024; 14(8):708. https://doi.org/10.3390/cryst14080708

Chicago/Turabian Style

Li, Zujun, Yushan Zhou, Xiuping Guo, Junhao Peng, and Huafeng Dong. 2024. "Coexistence of Large In-Plane and Out-of-Plane Piezoelectric Response in Group III–VI XMAY2 (X = I; M = Ti, Zr; A = Al, Ga; Y = S, Se) Monolayers" Crystals 14, no. 8: 708. https://doi.org/10.3390/cryst14080708

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop