Home Properties of Cmc21-X2As2O (X = Si, Ge, and Sn) by First-Principles Calculations
Article Publicly Available

Properties of Cmc21-X2As2O (X = Si, Ge, and Sn) by First-Principles Calculations

  • Ruike Yang EMAIL logo , Yucan Ma , Qun Wei and Dongyun Zhang
Published/Copyright: July 31, 2018

Abstract

For the compounds Cmc21-X2As2O (X = Si, Ge, and Sn), the stabilities are verified by the elastic constants and the phonon dispersion spectra. The structural, mechanical, electronic, and optical properties are investigated by using density functional theory (DFT) calculations. For Cmc21-X2As2O, the mechanical strengths in the [100], [010], and [001] directions are studied. Young’s modulus for Cmc21-Ge2As2O is more anisotropic than that of Cmc21-Si2As2O and Cmc21-Sn2As2O. The band structures of Cmc21-Si2As2O and Cmc21-Sn2As2O show that they are indirect-bandgap semiconductors with bandgaps of 2.744 and 2.201 eV, by using the HSE06 hybrid functional. Cmc21-Ge2As2O is a direct narrow-bandgap semiconductor with a bandgap of 2.131 eV. The static dielectric constants of Cmc21-Si2As2O and Cmc21-Sn2As2O in the [001] direction are higher than those in the [100] and [010] directions. The static dielectric constant of Cmc21-Ge2As2O in the [001] direction is lower than those in the [100] and [010] directions.

1 Introduction

In recent years, orthorhombic-(Cmc21-) Si2N2O has attracted much attention as a kind of high-temperature ceramic material. Cmc21-Si2N2O has broad application prospects. For example, it can be used as a high-temperature electric insulator, as a moderator in nuclear reactors, and as material for solid electrolytes [1]. Because Cmc21-Si2N2O ceramics have excellent mechanical and dielectric properties at high temperature, they can be used as a new generation of radome materials [2]. Cmc21-Si2N2O has excellent chemical stability, good oxidation resistance, high bending resistance, and good thermodynamic stability at 1400 K [3]. Because Cmc21-Si2N2O has high ability to absorb microwaves in the wavelength range 8–13 μm, it holds promise as a radiative cooling material [3]. As a unique technical ceramic, the orthorhombic (Cmc21)-Ge2N2O also has excellent chemical and oxidation resistance in high-temperature environments [4].

Recently, a family of 2D crystals, derived from the group-VA layered materials (P, As, Sb, Bi), has emerged with increasing research interests owing to their significant fundamental bandgaps. The semiconducting behaviour of group-VA 2D materials can render them as potential candidates for next-generation electronic and optoelectronic devices [5], [6], [7]. The chemical properties of silicon, germanium, and tin are similar, and the properties of nitrogen and arsenic are similar, so according to the structures of orthorhombic-(Cmc21-)X2N2O (X = Si, Ge), those of orthorhombic-(Cmc21-)X2As2O (X = Si, Ge, and Sn) can be predicted. On going through the literature, it is seen that Cmc21-X2As2O (X = Si, Ge, and Sn) have not been studied. We therefore investigate the properties of Cmc21-X2As2O S densities of states (PHDOS), band structures, and electronic densities of states (DOS) are analysed. And their optical properties in the [100], [010], and [001] directions are discussed.

2 Computational Methods

Structural optimisation and evaluation of the elastic and Young’s moduli and optical properties of Cmc21-X2As2O were carried out by using density functional theory (DFT) with the generalised gradient approximation (GGA) parameterised by Perdew, Burke, and Ernzerrof (PBE) in the CASTEP code [8]. The band structures and electronic DOS were calculated by the HSE06 hybrid functional. The phonon spectra and PHDOS were calculated by using the detailed first-principles linear response method. The cut-off energy and the number of special k-points are the major parameters that influence the accuracy of the calculations [9]. The cut-off energy determines the number of plane waves, and the Brillouin zone (BZ) is separated by the special k-points [10]. The cut-off energy is set to 610 eV. The convergence test and energy cut-off analysis of k-point mesh samples show that the convergence of the BZ sampling and the kinetic energy cut-off are reliable and satisfy the computational requirements [11], [12]. The Broyden-Fletcher-Goldfarb-Shanno (BFGS) minimisation scheme was used in geometry optimisation. The convergent value of the total energy step difference was set within 5.0 × 10−6 eV/atom. The maximum Hellmann-Feynman force, the maximum stress, and the maximum displacement were set 0.01 eV/Å, 0.02 GPa, and 5.0 × 10−4 Å, respectively, in geometry optimisation. The Monkhorst-Pack k-points in the first irreducible BZ were given as 9 × 9 × 8, 9 × 9 × 8, and 8 × 8 × 8, for Cmc21-Si2As2O, Cmc21-Ge2As2O, and Cmc21-Sn2As2O, respectively. The elastic constants, Young’s moduli, phonon spectra, PHDOS, band structures, DOS, and optical properties of Cmc21-X2As2O were calculated according to the optimised crystal structures and parameters.

The stress-strain methods were used to calculate the elastic constants. The number of steps was set to 6 during strain calculations. The maximum strain amplitude was 0.003. The scissors operator was taken as 0 eV. In calculating the optical properties, the instrument smearing (that is, the Gaussian broadening specified for calculating the dielectric functions) was set to 5 eV in the [100], [010], and [001] directions. The pseudo-potential was set to be norm-conserving when calculating the phonon spectra, PHDOS, band structures, electronic DOS, and optical properties [9].

3 Results and Discussion

3.1 Structural Properties

The compounds X2As2O (X = Si, Ge, and Sn) have an orthorhombic structure with Cmc21 space group (No. 36). Figure 1 shows the crystal structure of Cmc21-X2As2O at 0 GPa. The compounds Cmc21-Si2As2O, Cmc21-Ge2As2O, and Cmc21-Sn2As2O are all centro-symmetric structures (0, 0, 0) with 10 atoms per unit cell. The lattice parameters and volumes of Cmc21-X2As2O at 0 GPa were calculated on the basis of the geometry optimisation and are presented in Table 1. The density values of Cmc21-X2As2O (X = Si, Ge, and Sn) are 3.56, 4.60, and 4.90 g/cm3, respectively.

Figure 1: Crystal structures of the orthorhombic-X2As2O (X = Si, Ge, and Sn) (Cmc21, No. 36).
Figure 1:

Crystal structures of the orthorhombic-X2As2O (X = Si, Ge, and Sn) (Cmc21, No. 36).

Table 1:

Lattice constants a, b, c (Å) and volume (Å3) of Cmc21-X2As2O.

StructureabcV
Si2As2OGGA10.4156.5036.12414.50
Ge2As2OGGA10.7046.6506.309449.05
Sn2As2OGGA11.4197.1226.721546.62

3.2 Mechanical Properties and Young’s Moduli

The elastic constants and other mechanical properties were calculated by the GGA-PBE method. The nine independent elastic constants for Cmc21-X2As2O were calculated and are presented in Table 2. The mechanical stabilities of Cmc21-X2As2O can be estimated by the elastic constants. In this work, the mechanical stabilities and mechanical moduli of Cmc21-X2As2O were approximated by the corresponding relationships for orthorhombic crystal class [4], [13], [14], [15]. The elastic stability criteria of the orthorhombic phase are as follows:

Table 2:

Calculated independent elastic constants of Cmc21-X2As2O (X = Si, Ge, and Sn).

SpeciesC11C12C13C22C23C33C44C55C66
Si2As2O153.6629.8917.5656.6010.41103.4129.0531.7029.08
Ge2As2O120.989.2718.6940.790.2849.3621.2419.7119.24
Sn2As2O106.5219.5013.1443.678.9257.3517.5015.5217.87
(1)Cij>0(i=1,2,3,4,5,6),C11+C12+C33+2(C12+C13+C23)>0,
(2)C11+C222C12>0,C11+C332C13>0,C22+C332C23>0.

Based on Table 2 and judging the stability, the crystal structures of Cmc21-X2As2O are stable. Cmc21-Si2As2O exhibits the largest elastic constants C11, C22, C33. For Si2As2O and Ge2As2O, the calculated C11 is larger than C33 and C22. Hence, the mechanical strength in the [100] direction is higher than that in the [001] and [010] directions. Moreover, C44, C55, and C66 denote the shear moduli in the (100), (010), and (001) crystal planes, respectively. From Table 2, we find that there is a slight difference in C44, C55, and C66, so the shear moduli of Si2As2O and Ge2As2O are almost isotropic [13].

For Cmc21-Sn2As2O, C11 is larger than C22and C33, and the C22 is smaller than C33. Hence, the mechanical strength in the [100] direction is higher than that in the [001] and [010] directions. The mechanical strength in the [010] direction is less than that in the [001] direction. The calculated C55 is smaller than C44 and C66.

C12, C13, and C23 are the mechanical moduli corresponding to the biaxial strain (stress); they are the shear moduli of the (110) plane in the [110] direction ((110)[110]), ((101)[101]), and ((011)[011]), respectively [16].

Based on the elastic constants, other mechanical parameters of Cmc21-X2As2O were calculated by using the Voigt-Reuss-Hill (VRH) method, such as the bulk modulus (B), shear modulus (G), Young’s modulus (E), and Poisson’s ratio (v). For Cmc21-X2As2O, based on the elastic constants, the Reuss shear modulus (GR) and the Voigt shear modulus (GV) are as follows [4]:

(3)GR=15/(4(S11+S22+S33)4(S12+S13+S23)+3(S44+S55+S66)),
(4)GV=(C11+C22+C33C12C13C23)/15+(C44+C55+C66)/5.

The Reuss bulk modulus (GR) and the Voigt bulk modulus (GV) are defined as

(5)BR=1/((S11+S22+S33)+2(S12+S13+S23)),
(6)BV=(C11+C22+C33)/9+2(C12+C13+C23)/9,

where Sij and Cij are related by

(7)[Sij]=[Cij]1.

In the above formulas, Cij represents the elastic stiffness matrix and Sij represents the elastic flexibility matrix. Hill proposed that the upper and lower limits of the actual polycrystalline modulus are represented by the arithmetic mean of the Voigt and Reuss moduli, which are expressed as

(8)B=(BV+BR)/2,
(9)G=(Gv+GR)/2.

Young’s modulus (E) and Poisson’s ratio (v) can be calculated by the equations

(10)E=9BG/(3B+G),
(11)v=(3B2G)/(2(3B+G)).

Based on (3)–(11), the calculated physical quantities are presented in Table 3.

Table 3:

Calculated mechanical moduli of Cmc21-X2As2O, including bulk modulus (Bv, BR, and B), shear modulus (Gv, GR, and G), Young’s modulus (E) and Poisson’s ratio (v), B/G, and HV (in GPa).

SpeciesBvBRGvGRBGEvB/GHV
Si2As2O47.7139.5935.0218.4743.6533.3579.740.201.318.36
Ge2As2O29.7321.7824.2321.7325.7622.9853.140.161.127.95
Sn2As2O32.3027.6621.2419.1129.9820.1849.440.231.454.30

From the calculation results, the bulk modulus (B), shear modulus (G), and Young’s modulus (E) of Cmc21-Si2As2O are significantly larger than those of Cmc21-Ge2As2O and Cmc21-Sn2As2O. Cmc21-Si2As2O has the largest value of bulk modulus among Cmc21-X2As2O, which indicates that Cmc21-Si2As2O has lower compressibility. The shear modulus of Cmc21-Sn2As2O is the smallest among Cmc21-X2As2O. Cmc21-Sn2As2O has the smallest Young’s modulus among Cmc21-X2As2O, which indicates that the hardness of Cmc21-Sn2As2O is less than that of Cmc21-Si2As2O and Cmc21-Ge2As2O.

Poisson’s ratio represents the stability of the shear strain of the crystal. For a typical metal, the value should be 0.33. For an ionic-covalent crystal, the value ranges from 0.2 to 0.3. Poisson’s ratio of a strong covalent crystal is relatively small, usually below 0.15 [4]. For Cmc21-Si2As2O and Cmc21-Sn2As2O, Poisson’s ratios are close to that of covalent-ionic crystals. For Cmc21-Ge2As2O, the calculated Poisson’s ratio shows that it is a covalent crystal.

The ratio B/G reflects the brittleness or ductility of a material, and the critical value is close to 1.75 [17]. Below 1.75, the material shows brittleness; otherwise it shows toughness. The calculated B/G values of Cmc21-Si2As2O, Cmc21-Ge2As2O, and Cmc21-Sn2As2O are 1.31, 1.12, and 1.45, respectively. It suggests that Cmc21-Ge2As2O is more brittle than the other two.

The anisotropy of the mechanical properties of a crystal structure can be characterised by many different ways, such as the universal anisotropic index (AU) and the percent anisotropy (AG and AB for shear and bulk moduli) [18]. In contrast to many anisotropic indexes or factors, the universal anisotropic indexes or factors and the universal anisotropic index and percent anisotropy can be applied to any of these crystal structures. The equations used for the calculations are

(12)AU=5GV/GR+BV/BR60,
(13)AB=(BVBR)/(BV+BR),
(14)AG=(GVGR)/(GV+GR).

The shear anisotropic factor for the (100) shear planes between the [011] and [010] directions is

(15)A1=(4C44)/(C11+C332C13),

for the (010) shear planes between [101] and [001] directions, it is

(16)A2=(4C55)/(C22+C332C23),

and for the (0 0 1) shear planes between [110] and [0 10] directions, it is

(17)A3=4C66/(C11+C222C12).

In Table 4, we show the values of some of the anisotropic parameters mentioned above. For an isotropic structure, we expect that AU = 0, AG = 0, and AB = 0.

Table 4:

Calculated shear anisotropic factors (A1, A2, and A3), universal anisotropic index (AU), and percent anisotropy (AB and AG) of Cmc21-X2As2O (X = Si, Ge, and Sn) compounds.

AUABAGA1A2A3
Si2As2O4.680.0930.310.5242.710.773
Ge2As2O0.940.150.0540.6390.880.627
Sn2As2O0.720.0770.0530.5090.7460.643

The anisotropic index and the shear anisotropic factors are shown in Table 4. Based on the values of AG, AB, and AU, Cmc21-Si2As2O shows stronger mechanical anisotropy than Cmc21-Ge2As2O and Cmc21-Sn2As2O. The calculated AG and AU values of the Cmc21-Si2As2O are higher than those of the other two structures, which indicates that the shear modulus of Cmc21-Si2As2O has stronger anisotropy than those of Cmc21-Ge2As2O and Cmc21-Sn2As2O. The calculated AB of Cmc21-Ge2As2O is larger than those of the other two structures, which indicates that the bulk modulus of Cmc21-Ge2As2O shows stronger anisotropy than those for Cmc21-Si2As2O and Cmc21-Sn2As2O.

The anisotropic 3D graph of Young’s modulus is an easier way to determine anisotropy. For the orthorhombic crystal class, the directional dependence of Young’s modulus (E) can be written as

(18)1E=S11l14+2S12l12l22+S22l24+2S23l22l32+S33l34+2S13l12l32+S44l22l32+S55l12l32+S66l12l22.

In (18), Sij is the compliance matrix; l1, l2, and l3 are the direction cosines (which are given as l1 = sinθcosφ, l2=sinθsinφ, and l3=cosφ in spherical coordinates) [4]. The directional dependences of Young’s moduli of Cmc21-X2As2O are shown in Figure 2, and they show strong anisotropy. For Cmc21-Si2As2O, the anisotropy is weaker than those of Cmc21-Ge2As2O and Cmc21-Sn2As2O.

Figure 2: Directional dependences of Young’s moduli: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.
Figure 2:

Directional dependences of Young’s moduli: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.

3.3 Phonon Spectra

The calculated phonon dispersion spectra of Cmc21-X2As2O at 0 GPa are shown in Figure 3. The absence of any imaginary phonon frequencies in the entire BZ confirms that Cmc21-X2As2O are dynamically stable [19].

Figure 3: Phonon band structures: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.
Figure 3:

Phonon band structures: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.

The PHDOS of Cmc21-X2As2O were calculated by CASTEP and are shown in Figure 4. In Figure 4a, there are five main regions of phonon bands for Cmc21-Si2As2O. When the frequency is in the range 0–6.21 THz, the PHDOS is mainly from the As atoms. When the frequency is in the ranges 6.21–9.0 and 25.10–27.97 THz, the PHDOS mainly originates from the O atoms. Within the ranges 10.64–14.05 and 18.37–19.36 THz, the PHDOS is mainly from the Si atoms. From Figure 4b, there are five main regions of phonon bands for Cmc21-Ge2As2O. Within the ranges 0–4.70 and 6.9–9.29 THz, the PHDOS is mainly from the Ge and As atoms. In the range 4.7–6.9 THz, the PHDOS mainly originates from the O and As atoms. Within the range 13.1–13.89 THz, the PHDOS is mainly from the O and Ge atoms. In the range 18.5–20.43 THz, the PHDOS mainly comes from the O atoms. From Figure 4c, there are four main regions of phonon bands for Cmc21-Sn2As2O. Because the Sn and As atoms are heavier than the O atom, the vibrational frequencies of the Sn and As atoms are obviously lower than that of the O atom [20].

Figure 4: Phonon DOS: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.
Figure 4:

Phonon DOS: (a) Cmc21-Si2As2O, (b) Cmc21-Ge2As2O, and (c) Cmc21-Sn2As2O.

3.4 Band Structures and Densities of States

The electronic properties of the Cmc21-X2As2O structures were analysed at 0 GPa. The DOS was calculated to get further insight into the bonding characteristics of Cmc21-X2As2O. The structural stability of a compound is related to its bonding electron orbits.

For Cmc21-Si2As2O, the valence band maximum is at the highly symmetric S-point and the conduction band minimum is at the Z-points, which indicates Cmc21-Si2As2O is an indirect bandgap semiconductor (Figure 5a); and the calculated bandgap is 2.024 eV. The total and partial DOS of Cmc21-Si2As2O are presented in Figure 5b. The main bonding peaks are in the range –8 to 8 eV. The total DOS in the valence band mainly comes from As p and O p orbitals and partially from Si p. The latter contributes most to the total DOS in the conduction band, with partial contributions from As p and Si s. In Figure 5b, the p states of the As atoms and the Si atoms have strong hybridisation. The result shows that interactions of covalent Si-As exist in Cmc21-Si2As2O.

Figure 5: Band structure and densities of states: (a) Cmc21-Si2As2O band structure and (b) DOS; (c) Cmc21-Ge2As2O band structure and (d) DOS; (e) Cmc21-Sn2As2O band structure and (f) DOS.
Figure 5:

Band structure and densities of states: (a) Cmc21-Si2As2O band structure and (b) DOS; (c) Cmc21-Ge2As2O band structure and (d) DOS; (e) Cmc21-Sn2As2O band structure and (f) DOS.

From Figure 5c, the valence band maximum and the conduction band minimum are near the Z-point, which indicates that Cmc21-Ge2As2O is a direct narrow-bandgap semiconductor. The bandgap is 1.487 eV. Figure 5d shows the total and partial DOS. Below the Fermi level, the total DOS is mainly from As p and partially from Ge p and O p. Above the Fermi level, the total DOS originates mainly from the Ge p and As p. The p states of As and Ge have a hybrid effect.

The band structure of Cmc21-Sn2As2O is shown in Figure 5e. In the band structure, the valence band maximum is located at the S-point and the conduction band minimum at the G-point. Cmc21-Sn2As2O is an indirect bandgap semiconductor. The bandgap of Cmc21-Sn2As2O is 1.512 eV. From Figure 5f, the total DOS in the valence band originates mainly from As s state. The total DOS in the conduction band originates mainly from O s state. From Figure 5f, some hybridisation from the s states of the As atoms and the O atoms can be found. The results show that there are covalent As-O interactions in Sn2As2O.

3.5 Optical Properties

Dielectric function is the most common characteristic of materials. It can characterise the response of materials to incident electromagnetic waves [21]. Optical properties of materials can usually be evaluated on the basis of a complex dielectric function, which depends on the frequency. The complex dielectric function is given by [22]

(19)ε(ω)=ε1(ω)+iε2(ω),

where ε1(ω) is the real part and ε2(ω) is the imaginary part. ε2(ω) is calculated on the basis of the momentum matrix elements between the occupied and unoccupied wave functions as follows [23]:

(20)ε2(ω)=(Ve2)/(2πm2ω2)d3k|kn|p|kn|2f(kn)×(1f(kn))δ(EknEknω),

where V is the volume of the unit cell, e is electronic charge, p is the momentum operator, |kn is for crystal wave function, f(kn) denotes the Fermi distribution function, and ω is the energy of the incident photon. The real part ε1(ω) is given by Kramers-Kroning relationship

(21)ε1(ω)=1+2πM0ε2(ω)ωω2ω2dω,

where M is the principal value of the integral. Generally, ε1(ω) is connected with the electric polarisation characteristics of the material.

The absorption coefficient A(ω) can be calculated on the basis of the complex dielectric function ε(ω), as follows [20]:

(22)A(ω)=2ωk(ω)/c.

The optical conductivity usually plays an important part in exploring the electronic structures of a compound and its response to external conditions such as temperature and pressure. So, it is necessary to calculate the optical conductivity. The optical conductivity σ(ω) is expressed by the complex dielectric function as follows:

(23)σ(ω)=iω4π[1ε(ω)].

Other optical constants, such as the absorption coefficient I(ω), the optical reflectivity R(ω), and energy-loss spectrum L(ω), can be calculated using ε(ω). These expressions are as follows: [24], [25], [26]:

(24)I(ω)=2ωk(ω)c,
(25)R(ω)=(n1)2+k2(n+1)2+k2,
(26)L(ω)=Im(1ε(ω))=ε2(ω)ε1(ω)+ε2(ω).

The optical properties were calculated and are presented in Figure 6 for the energy range up to 40 eV. In Figure 6a, the real parts of the calculated dielectric functions are shown for Cmc21-X2As2O in the [100], [010], and [001] directions. The static dielectric constants of Cmc21-X2As2O are presented in Table 5. The static dielectric constants of Cmc21-Si2As2O and Cmc21-Sn2As2O in the [001] polarisation direction are higher than those in the [100] and [010] directions. The static dielectric constant of Cmc21-Ge2As2O in the [001] polarisation direction is lower than those in the [100] and [010] directions. For Cmc21-X2As2O, the real parts of dielectric functions in the [001] direction increase with increasing photon energy and achieve the highest values (16.2, 16.9, and 14.9 F/m) at 2.64, 2.14, and 2.58 eV, respectively. The real parts of the dielectric functions of X2As2O in the [010] direction increase with the increasing photon energy and achieve the highest values (11.1, 15.2, and 11.6 F/m) at 2.15, 1.12, and 1.76 eV, respectively. The real parts of dielectric functions in the [100] direction achieve the highest values (11.1, 15.2, and 11.6 F/m) at 2.15, 1.12, and 1.76 eV, respectively.

Figure 6: (a) Real parts of dielectric function, (b) imaginary parts of dielectric function, (c) real parts of conductivity, (d) absorption, (e) reflectivity, and (f) loss function for Cmc21-X2As2O in the [100], [010], and [001] directions.
Figure 6:

(a) Real parts of dielectric function, (b) imaginary parts of dielectric function, (c) real parts of conductivity, (d) absorption, (e) reflectivity, and (f) loss function for Cmc21-X2As2O in the [100], [010], and [001] directions.

Table 5:

Calculated static dielectric constants (F/m) of Cmc21-X2As2O in the [100], [010], and [001] directions.

[100][010][001]
Si2As2O8.598.599.78
Ge2As2O12.8012.8011.00
Sn2As2O9.019.019.23

The curves of the imaginary parts of the dielectric functions in the [001], [010], and [100] directions are shown in Figure 6b. Cmc21-Si2As2O and Cmc21-Sn2As2O have only one peak in the [100], [010], and [001] directions. Cmc21-Ge2As2O has two peaks in the [100] and [010] directions and one peak in the [001] direction. Thus, Cmc21-Ge2As2O is more anisotropic than Cmc21-Si2As2O and Cmc21-Sn2As2O.

The curves of the real parts of the conductivities for Cmc21-X2As2O are plotted in Figure 6c. The real parts of the conductivities of Cmc21-Si2As2O in the [100], [010], and [001] directions increase with the increase in photon energy and reach the highest values (6.84, 11, and 11) at 5.91, 2.25, and 2.08 eV, respectively. For Cmc21-Ge2As2O, the real parts of the conductivities in the [100], [010], and [001] directions increase with the increase in photon energy and reach the highest values (6.72, 6.74, and 7.41) at 5.62, 5.71, and 4.43. For Cmc21-Sn2As2O, the real parts of the conductivities in the [100], [010], and [001] directions increase with the increase in photon energy and reach the highest values (6.15, 6.15, and 7.32) at 5.66, 5.77, and 4.31 eV.

The absorption coefficients in the [100], [010], and [001] directions are shown in Figure 6d. For Cmc21-Si2As2O, the peaks of the absorption spectra in the [100], [010], and [001] directions are located at 7.99, 8.06, and 5.62 eV, respectively. For Cmc21-Ge2As2O, the corresponding peaks are located at 6.8, 6.8, and 5.8 eV, respectively, and for Cmc21-Sn2As2O they are located at 6.83, 6.88, and 5.57 eV, respectively.

The curves of the reflectivity R(ω) of Cmc21-X2As2O are plotted in Figure 6e. For Cmc21-Si2As2O, the reflectivity spectra in the [100], [010], and [001] directions are characterised by two, two, and one peaks, respectively. In the [100] direction, these peaks are located at 8.4 and 11.9 eV, respectively. Similar peaks also appear in the [010] direction. In the [001] direction, the peak is located at 5.96 eV.

For Cmc21-Ge2As2O, the reflectivity spectra in the [100] direction are characterised by four peaks. These peaks are located at 1.63, 7.19, 11.4, and 14 eV, respectively. In the [010] direction, the four peaks are located at 1.72, 7.05, 11.6, and 14 eV, respectively. In the [001] direction, there is only one peak, which is located at 6.17 eV.

For Cmc21-Sn2As2O, the reflectivity spectra in the [100] direction are characterised by three peaks. These peaks are located at 2.2, 7.03, and 11.4 eV, respectively. In the [010] direction, the three peaks are located at 2.23, 7.06, and 11 eV, respectively. In the [001] direction, there is only one main peak, which is located at 13.4 eV.

The loss function L(ω) describes the energy loss of an electron passing through the material. For Cmc21-X2As2O, the calculated energy losses are shown in Figure 6f. We can see the characteristics of plasma oscillation from the peak of L(ω). And L(ω) also can describe the frequency of collective oscillation of the valence electrons. Above the plasma frequency, the material describes the dielectric behaviour, and below, the material shows the metallic property. The position of peak in the L(ω) spectrum indicates the change of the material from metallic to dielectric character [6]. For Cmc21-Si2As2O, in the [100], [010], and [001] directions, each peak of L(ω) is located at 17.7 eV. Similar are the situations in Cmc21-Ge2As2O in the [100], [010], and [001] directions. For Cmc21-Sn2As2O, the peaks of L(ω) in the [100], [010], and [001] directions are located at 16.4, 16.4, and 16.9 eV, respectively.

4 Conclusion

For the novel predicted materials Cmc21-X2As2O (X = Si, Ge, and Sn), the structural, mechanical, Young’s anisotropy, phonon spectra, band structures, electronic DOS, and optical properties were studied from first-principles calculations. For Cmc21-Si2As2O, the mechanical strength in the [100] direction is higher than that in the [001] and [010] directions. For Cmc21-Ge2As2O and Sn2As2O, the mechanical strength in the [010] direction is less than that in the [001] direction. Cmc21-X2As2O all show brittleness. Cmc21-Ge2As2O shows stronger anisotropy than Cmc21-Si2As2O and Cmc21-Sn2As2O. Phonon spectral calculations confirm that Cmc21-X2As2O (X=Si, Ge and Sn) are all dynamically stable. By analysing the PHDOS, we found that the vibrational frequency of the X (Si, Ge and Sn) atoms and As atoms were obviously lower than that of the O atoms. Calculation of the band structures showed they they are semiconductors. The electronic DOSs showed that there are covalent X-As interaction in Cmc21-X2As2O (X = Si, Ge, and Sn). The static dielectric constants of Cmc21-Si2As2O and Cmc21-Sn2As2O in the [001] polarisation direction are higher than those in the [100] and [010] directions. The static dielectric constant of Cmc21-Ge2As2O in the [001] direction is lower than that in the [100] and [010] directions.

Acknowledgement

This work was supported by the Natural Science Basic Research plan of Shanxi Province of China (2016JM1026) and by the 111 Project (B17035). It was also supported by the LEIHUA Electronic and Technology Research Institute, Aviation Industry Corporation of China.

References

W. Y. Ching and S.-Y. Ren, Phys. Rev. B 24, 5788 (1981).10.1103/PhysRevB.24.5788Search in Google Scholar

S. Lin, F. Ye, J. Ma, J. Ding, C. Yang, et al., Mater. Des. 97, 51 (2016).10.1016/j.matdes.2016.02.064Search in Google Scholar

Z.-L. Lv, H.-L. Cui, H. Wang, X.-H. Li, and G.-F. Ji, Ceram. Int. 43, 10006 (2017).10.1016/j.ceramint.2017.05.014Search in Google Scholar

Y. Ding, Phys. B Condens. Matter 407, 2190 (2012).10.1016/j.physb.2012.02.040Search in Google Scholar

S. Zhang, W. Zhou, Y. Ma, J. Ji, B. Cai, et al., Nano Lett. 17, 3434 (2017).10.1021/acs.nanolett.7b00297Search in Google Scholar PubMed

S. Zhang, Z. Yan, Y. Li, Z. Chen, and H. Zeng, Angew. Chem. Int. Ed. 54, 3112 (2015).10.1002/anie.201411246Search in Google Scholar PubMed

S. Zhang, S. Guo, Z. Chen, Y. Wang, H. Gao, et al., Chem. Soc. Rev. 47, 982 (2018).10.1039/C7CS00125HSearch in Google Scholar PubMed

Q. Fan, Q. Wei, H. Y. Yan, M. Zhang, Z. Zhang, et al., Comput. Mater. Sci. 85, 80 (2014).10.1016/j.commatsci.2013.12.045Search in Google Scholar

R. Yang, C. Zhu, Q. Wei, and Z. Du, J. Phys. Chem. Solids 104, 68 (2017).10.1016/j.jpcs.2016.12.032Search in Google Scholar

H. J. Monkhorst, Phys. Rev. B Condens. Matter 16, 1748 (2005).10.1103/PhysRevB.16.1748Search in Google Scholar

X. Zhang and H. F. Shi, Mater. Sci. Technol. 30, 732 (2014).10.1179/1743284713Y.0000000423Search in Google Scholar

R. Yang, C. Zhu, Q. Wei, and Z. Du, J. Phys. Chem. Solids 98, 10 (2016).10.1016/j.jpcs.2016.05.012Search in Google Scholar

J. P. Perdew and K. Burke, Phys. Rev. Lett. 77, 3865 (1996).10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed

Z.-J. Wu, E.-J. Zhao, H.-P. Xiang, X.-F. Hao, X.-J. Liu, et al., Phys. Rev. B 76, 054115 (2007).10.1103/PhysRevB.76.054115Search in Google Scholar

B. Xiao, J. Feng, C. T. Zhou, Y. H. Jiang, and R. Zhou, J. Appl. Phys. 109, 083521 (2011).10.1063/1.3596826Search in Google Scholar

Y. Ding, M. Chen, W. Wu, and M. Xu, J. Electron. Mater. 46, 1 (2017).10.1007/s11664-016-4861-2Search in Google Scholar

S. F. Pugh, XCII, Philos. Mag. Ser. 1 45, 823 (2009).10.1080/14786440808520496Search in Google Scholar

B. Xiao, J. Feng, C. T. Zhou, Y. H. Jiang, and R. Zhou, J. Appl. Phys. 109, 023507 (2011).10.1063/1.3532038Search in Google Scholar

R. Yang, C. Zhu, Q. Wei, and D. Zhang, Solid State Commun. 267, 23 (2017).10.1016/j.ssc.2017.09.008Search in Google Scholar

R. Yang, C. Zhu, Q. Wei, and Z. Du, Philos. Mag. 97, 3008 (2017).10.1080/14786435.2017.1364438Search in Google Scholar

Z. Y. Jiao, S.-H. Ma, and J.-F. Yang, Solid State Sci. 13, 331 (2011).10.1016/j.solidstatesciences.2010.11.030Search in Google Scholar

J. Feng, B. Xiao, J. C. Chen, and C. T. Zhou, Solid State Sci. 11, 259 (2009).10.1016/j.solidstatesciences.2008.04.015Search in Google Scholar

X. Li, H. L. Cui, and R. Z. Zhang, Sci. Rep. 6, 39790 (2016).10.1038/srep39790Search in Google Scholar PubMed PubMed Central

Z. G. Shao and Z. L. Sun, Physica E 74, 438 (2015).10.1016/j.physe.2015.07.011Search in Google Scholar

D. Misra and T. K. Kundu, Comput. Mater. Sci. 112, 113 (2016).10.1016/j.commatsci.2015.10.021Search in Google Scholar

J. Feng, B. Xiao, J. C. Chen, C. T. Zhou, Y. P. Du, et al., Solid State Commun. 149, 1569 (2009).10.1016/j.ssc.2009.05.042Search in Google Scholar

Received: 2018-03-23
Accepted: 2018-07-08
Published Online: 2018-07-31
Published in Print: 2018-10-25

©2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 8.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2018-0151/html
Scroll to top button