Home Ab initio studies of the structural, electronic, and optical properties of quaternary BxAlyGa1–x–yN compounds
Article Publicly Available

Ab initio studies of the structural, electronic, and optical properties of quaternary BxAlyGa1–xyN compounds

  • M’hamed Larbi , Rabah Riane , Samir F. Matar EMAIL logo , Ahmed Abdiche , Mustapha Djermouni , Mohammed Ameri , Nacera Merabet and Allel Oualdine
Published/Copyright: January 7, 2016
Become an author with De Gruyter Brill

Abstract

Original first-principles calculations were performed to study the structural and electronic properties of quaternary BxAlyGa1–xyN compounds, using the non-relativistic full-potential linearized augmented plane wave method as employed in the Wien2k code. For the exchange-correlation potential, local density approximation and generalized gradient approximation have been used to calculate theoretical lattice parameters, bulk modulus, and their pressure derivatives. Non-linear variation with compositions x and y of the lattice parameter, bulk modulus, and direct and indirect band gaps have been found. The calculated bowing of the fundamental band gaps is in good agreement with the available experimental and theoretical values.

1 Introduction

III–V nitride semiconductors and their compounds have attracted great attention as some of the most important materials for optoelectronic and electronic applications. Gallium nitride (GaN), aluminum nitride (AlN), and related materials are of considerable current interest because of their applications in light-emitting devices operating in the visible and deep ultraviolet (UV) spectral regions. Boron nitride (BN) is a very good choice for protective coatings because of its hardness, high melting point, and large bulk modulus for the cubic variety. It also has features of high thermal conductivity suitable for applications in electronic devices [1].

Recently, many groups paid attention to studying the properties of the quaternary solid solution BxAlyGa1–xyN. There has also been considerable interest in the use of models and numerical methods in materials science, including the density functional theory (DFT) [2, 3] to study the structural and electronic properties of III–V semiconductors and their solid solutions.

In the present work, we have studied the structural and electronic properties of the binary zincblende (ZB) BN, AlN, GaN, and their ternaries BxGa1–xN, BxAl1–xN, and AlxGa1–xN as well as BxAlyGa1–xyN quaternary solid solutions over the entire composition range of x and y by performing first-principles calculations, based on the full-potential linearized augmented plane wave (FP-LAPW) method within DFT. To the best of our knowledge, this is the first quantitative theoretical investigation on BxAlyGa1–xyN quaternary solid solutions, and the results still await experimental confirmations.

We briefly describe the method of calculation, and then the results of our work are presented and discussed followed by conclusions.

2 Method of calculations

The DFT calculations on the structural and electronic properties of BN, AlN, and GaN binaries, their ternary compounds BxGa1–xN, BxAl1–xN, and AlxGa1–xN as well as the solid solution BxAlyGa1–xyN have been performed by the FP-LAPW method, which implements the Wien2k code [4, 5]. The calculation of the total energy was performed using two approximations in terms of exchange-correlation XC potentials, local density approximation (LDA) as parameterized by Perdew and Wang [6] and generalized gradient approximation (GGA) parameterized by Perdew et al. [7].

In this method, the primitive unit cell is divided into non-overlapping spheres around atoms and remaining interstitial regions, where the Kohn-Sham wave functions, charge density, and potential are treated differently in these regions of the unit cell. Inside the atomic spheres of radius RMT around each atom, radial solution of the Schrodinger equation times the spherical harmonic are used for expansion, and the plane wave basis set is used in the interstitial region. Muffin-tin radii RMT of B, N, Al, and Ga of 1.35, 1.4, 1.7, and 2.1 bohr (0.529 Å), respectively, are adopted.

For the wave function expansion, the maximum value of the secondary quantum number l was limited to lmax = 10 inside the atomic spheres. For plane wave cutoff, KmaxRMT = 7.0 was used to expand the wave functions in the interstitial region and for the charge density Fourier series were truncated at Gmax = 12. The integrals over the Brillouin zone (BZ) are performed up to 1000 k-points for the ZB binary compounds and 350 k-points for the quaternary ones in the irreducible BZ.

3 Results and discussion

3.1 Structural properties

The quaternary system, studied in this paper includes three binary compounds BN, AlN, GaN, and three ternary solid solutions BxGa1–xN, BxAl1–xN, and AlxGa1–xN. To calculate the structural properties of the binary and ternary solid solutions and then their quaternary solid solutions, we started calculations with the ZB structure and let the calculated forces move the atoms to their equilibrium positions. We have chosen eight atoms (in primitive P mode of fcc ZB) of 1 × 1 × 1 single cells for modeling the structure of the quaternary solid solutions BxAlyGa1–xyN for the considered structures and at different boron and aluminum concentrations x and y for compositions (x = 0.25, y = 0.25; x = 0.25, y = 0.50; and x = 0.50, y = 0.25). The structural properties were obtained by calculating the total energies for different volumes around the equilibrium cell volume V0 of the BN, AlN, and GaN binary compounds and their solid solutions. The calculated total energies were fitted to the Birch–Murnaghan equation of state [8] to determine the ground state properties such as the equilibrium lattice parameter a0, the bulk modulus B0, and the pressure derivative of the bulk modulus B′.

The calculated equilibrium parameters a0, B0, and B′ for the binary compounds are given in Table 1, which also contains results of previous calculations as well as the experimental data. There is a good agreement between our results and the reported theoretical investigations. In comparison with the experimental data, we found that GGA overestimates the lattice parameter and LDA underestimates it, whereas the bulk modulus is underestimated with GGA and overestimated with LDA. These observations are in concordance with the general trend of these approximations. Our calculated lattice parameters and bulk modulus at different compositions of ternary solid solution BxGa1–xN, BxAl1–xN, and AlxGa1–xN are depicted in Figs. 1 and 2. The lattice parameters vary almost linearly with concentration, with the trend following the Vegard’s law. The obtained results are fitted to the following equations:

Table 1

Lattice parameters a0, bulk modulus B0, and pressure derivative of bulk modulus B′ for ZB BN, AlN, and GaN binary compounds.

Binary compoundsThis workOther theoretical studiesExperimental
GGALDA
BN
a0, Å3.613.583.35a, 3.62b, 3.66c3.61d
B0, GPa383402401a, 368b, 353c382, 400e
B4.13.73.66a;3.32b3.0, 4.0e
AlN
a0, Å4.394.344.34a, 4.35f4.38g
B0, GPa196207211a, 209f
B4.103.953.90a, 3.89f
GaN
a0, Å4.544.474.46h, 4.51i4.49j, 4.50g
B0, GPa186209202h, 191i4.53k, 190l
B4.274.74.43h, 4.14i

References: a[9], b[10], c[11], d[12], e[13], f[14], g[15], h[16], i[17], j[18], k[19], and l[20].

Fig. 1: The calculated lattice parameters versus compositions for the ternary solid solutions using GGA.
Fig. 1:

The calculated lattice parameters versus compositions for the ternary solid solutions using GGA.

Fig. 2: The calculated bulk modulus B0versus compositions x for ternary solid solutions using GGA XC approximation.
Fig. 2:

The calculated bulk modulus B0versus compositions x for ternary solid solutions using GGA XC approximation.

For the lattice parameters:BxGa1xN: 4.53 0.5x0.37x2BxAl1xN: 4.38 0.45x0.322x2AlxGa1xN: 4.53 0.13x0.01x2For the bulk modulus:BxGa1xN: 187.65 + 18.34x+ 174.85x2BxAl1xN: 198.17 + 9.42x+ 172.57x2AlxGa1xN: 185.74 + 17.25x 6.85x2

The deviations observed from the linear concentration dependence can be explained mainly by the large mismatches of the lattice parameters of the binary constituents. From Figs. 1 and 2, we observe that the lattice parameter increases when the composition x is decreasing and the bulk modulus decreases the composition x is decreasing. Results obtained in our calculations agree well with other available data [16].

The structural and electronic properties of pseudo-binary (ternary) solid solutions BxGa1–xN, BxAl1–xN, and AlxGa1–xN were studied in the second step of our investigation. We used ordered structures described in terms of supercells with eight atoms per unit cell, for the compositions x = 0.25, 0.50, and 0.75. For the considered structures, we performed the structural optimization by minimizing the total energy with respect to the cell parameters and the atomic positions. The calculated lattice parameters at different compositions of BxGa1–xN, BxAl1–xN, and AlxGa1–xN as shown in Table 2.

Table 2

Lattice parameters a0, bulk modulus B0, and pressure derivative of bulk modulus B′ for ZB BxGa1–xN, BxAl1–xN and AlxGa1–xN solid solutions.

CompositionThis workOther theoretical studiesExperimental
GGALDA
B0.25Ga0.75N
a0, Å4.364.314.38a, 4.31b
B0, GPa206229198a, 226b
B4.584.534.19a, 4.10b
B0.50Ga0.50N
a0, Å4.174.134.19a, 4.13b
B0, GPa242266235a, 261b
B4.404.374.77a, 3.88b
B0.75Ga0.25N
a0, Å3.933.893.90a, 3.86b
B0, GPa295317302a, 287b
B4.334.153.15a, 3.8b
B0.25Al0.75N
a0, Å4.254.204.18c,4.20d
B0, GPa215230235d, 214e
B4.014.014.53d
B0.50Al0.50N
a0, Å4.084.043.97c,4.04d
B0, GPa248266266d, 245e
B4.024.173.56d
B0.75Al0.25N
a0, Å3.873.833.75c, 3.84d
B0, GPa296317318d, 296e
B3.923.872.95d
Al0.25Ga0.75N
a0, Å4.504.444.39f, 4.49g
B0, GPa189211203f
B4.624.504.39f
Al0.50Ga0.50N
a0, Å4.474.414.41f, 4.46g
B0, GPa193212205f
B4.364.154.46f
Al0.75Ga0.25N
a0, Å4.434.384.37f, 4.42g
B0, GPa195213207f
B4.164.134.53f

References: a[19], b[21], c[22], d[23], e[24], f[25], and g[26].

The calculated structural properties for the quaternary solid solutions are listed in Table 3. To the best of our knowledge, there are no experimental or theoretical data for the structural properties available in the literature for the studied solid solutions.

Table 3

Lattice parameters a0, bulk modulus B0, and pressure derivative of bulk modulus B′ for ZB BxAlyGa1–xyN solid solutions.

Composition (x,y)This work
GGALDA
B0.25Al0.25Ga0.50N
a0, Å4.334.27
B0, GPa210230
B4.384.29
B0.25Al0.50Ga0.25N
a0, Å4.294.24
B0, GPa213231
B4.144.07
B0.50Al0.25Ga0.25N
a0, Å4.134.08
B0, GPa245264
B4.604.18

Our results can serve as a reference for future investigations. Usually, in the treatment of solid solutions when experimental data are scarce, it is assumed that the atoms are located at the ideal positions and the lattice parameters vary linearly with concentration x according Vegard’s law [27]. However, violation of this linear law has been observed experimentally [28] and theoretically for semiconductor compounds. Assuming that Vegard’s law is valid, the lattice constants of BxAlyGa1–xyN quaternary alloys are generally expressed as a linear relation of boron composition x and aluminum composition y by the following expression [29]:

a(x,y)=x(BN)+y(AlN)+(1xy)(GaN)

We tested the validity of Vegard’s law for the BxAlyGa1–xyN quaternary solid solutions in the ZB structure with different boron and aluminum concentrations. a(BN), a(AlN), and a(GaN) are the equilibrium lattice parameter of BN, AlN, and GaN, respectively. a(x,y) represents the composition-dependent lattice constant of BxAlyGa1–xyN quaternary solid solutions.

3.2 Electronic structure properties

Accessing the energy band structures in semiconductors provides valuable information regarding their potential utility in the fabrication of electronic and optoelectronic devices. As group III nitrides are promising materials, the accurate knowledge of the band structures of the BxGa1–xN, BxAl1–xN, AlxGa1–xN, and BxAlyGa1–xyN compounds becomes essential. The band gap energies of BN, AlN, GaN, BxGa1–xN, BxAl1–xN, AlxGa1–xN, and BxAlyGa1–xyN were calculated for the equilibrium calculated lattice parameters [1].

Using our calculated values of lattice parameters with GGA, we computed the electronic band structure along some high symmetry directions in the BZ, with LDA and GGA XC. The obtained band structures for ternaries at x = 0.5 are presented in Fig. 3, and the density of states (DOS) of BxAlyGa1–xyN quaternary solid solutions for compositions (x = 0.25, y = 0.25; x = 0.25, y = 0.50, and x = 0.5, y = 0.25) are presented in Fig. 4.

Fig. 3: Non-relativistic band structure of ZB: B0.50Ga0.50N, B0.50Al0.50N, and Al0.50Ga0.50N ternary solid solutions using mBJ approximation.
Fig. 3:

Non-relativistic band structure of ZB: B0.50Ga0.50N, B0.50Al0.50N, and Al0.50Ga0.50N ternary solid solutions using mBJ approximation.

Fig. 4: Non-relativistic band structure of ZB-type BxAlyGa1–x–yN: B0.25Al0.25Ga0.50N, B0.25Al0.50Ga0.25N, and B0.50Al0.25Ga0.25N using mBJ approximation.
Fig. 4:

Non-relativistic band structure of ZB-type BxAlyGa1–xyN: B0.25Al0.25Ga0.50N, B0.25Al0.50Ga0.25N, and B0.50Al0.25Ga0.25N using mBJ approximation.

The numerical results are listed in Table 4 and compared with the available experimental and theoretical data. They show that the valence band maximum (VBM) occurs at the BZ center Γ point and the conduction band minimum (CBM) is located at the X point in both BN and AlN, thus resulting in indirect (Γ–X) band gaps for BN and AlN and direct (Γ–Γ) band gaps for GaN and ternaries solid solutions BxGa1–xN, BxAl1–xN, and AlxGa1–xN.

Table 4

Calculated direct (Γ–Γ) and indirect (Γ–X) band gap energies for ZB BN, AlN, GaN, BxGa1–xN, BxAl1–xN, and AlxGa1–xN compounds.

This workTheoretical calculationsExperimental
GGALDAmBJ
BN
E(Γ–Γ)8.828.8310.478.67a, 8.79b
E(Γ–X)4.254.385.889.09c6d
3.95a, 4.45b, 4.24c
AlN
E(Γ–Γ)3.954.395.483.92c, 4.58e
E(Γ–X)3.193.254.944.03f
3.28c, 3.29i
GaN
E(Γ–Γ)1.481.882.881.52g, 2.1h3.2i
E(Γ–X)3.173.254.663.22g
B0.25Ga0.75N
E(Γ–Γ)2.324.172.733.592.73e, 2.95j
E(Γ–X)4.315.583.48e, 4.45j
B0.50Ga0.50N
E(Γ–Γ)3.173.913.254.703.26e, 3.40j
E(Γ–X)4.015.344.03e, 5.36j
B0.75Ga0.25N
E(Γ–Γ)3.594.953.695.153.67e, 3.90j
E(Γ–X)5.066.504.27e, 6.30h
B0.25Al0.75N
E(Γ–Γ)3.425.273.485.163.7k, 3.45l
E(Γ–X)5.456.784.16m
B0.50Al0.50N
E(Γ–Γ)3.574.883.635.264.4k, 3.64l
E(Γ–X)5.036.384.35m
B0.75Al0.25N
E(Γ–Γ)3.965.804.065.624.9k, 4.05l
E(Γ–X)5.967.354.69m
Al0.25Ga0.75N
E(Γ–Γ)2.064.732.453.462.22o
E(Γ–X)4.946.084.80n
Al0.50Ga0.50N
E(Γ–Γ)2.595.053.004.012.78o
E(Γ–X)5.276.434.75n
Al0.75Ga0.25N
E(Γ–Γ)3.215.433.284.713.50o
E(Γ–X)5.636.874.65n

All energies are given in electron volts (eV).

References: a[30], b[31], c[32], d[16, 33], e[34], f[35], g[36], h[12], i[37], j[21], k[24], l[38], m[23], n[25], and o[26].

The calculated band gaps are in very good agreement with those of available theoretical calculations. It is worth mentioning here that the calculated band gap values listed in Table 4 are systemically underestimated in comparison with experimental data. Such underestimation of the band gaps is mainly because the simple forms of LDA and GGA do not take into account the quasi-particle self-energy correctly [36], which makes them insufficiently flexible to accurately reproduce both XC energy and its charge derivative.

The band gap energies of BxAlyGa1–xyN solid solutions have been computed using both GGA, LDA, and modified Bezcke–Johnson (mBJ) approximations [39]. We found that the quaternary solid solutions BxAlyGa1–xyN have a direct band gap. The resulting values for the different x and y concentrations are given in Table 5. It can be seen that the large value calculated for the quaternary solid solutions of the direct band gap using LDA approximation is 5.65 eV, which corresponds to x = 0.50, y = 0.25.

Table 5

Calculated direct (Γ–Γ) and indirect (Γ–X) band gaps for ZB-type BxAlyGa1–x–yN quaternary solid solutions.

Composition (x, y)This work
GGALDAmBJ
B0.25Al0.25Ga0.50N
E(Γ–Γ)2.893.374.13
E(Γ–X)4.504.685.90
B0.25Al0.50Ga0.25N
E(Γ–Γ)3.293.354.72
E(Γ–X)4.845.016.28
B0.50Al0.25Ga0.25N
E(Γ–Γ)3.393.474.96
E(Γ–X)5.525.657.09

All energies are in eV.

The band gap energy of an AxB1–xC compound can be depicted as a function of the A composition x and can be approximated using the following equation:

Eg(x)=xEg(AC)+(1x)Eg(BC)x(1x)b,

where Eg(x) is the band gap energy of the AxB1–xC solid solution, EgAC and EgBC are the band gap energies of the binary compounds AC and BC, respectively, and the quadratic term b is the bowing parameter of AxB1–xC [1].

The total bowing parameter is calculated by fitting the band gap energies of BxGa1–xN, BxAl1–xN, and AlxGa1–xN obtained with the equilibrium lattice parameters. The corresponding plots are provided in Figs. 57; the non-linear variation of the calculated direct and indirect band gaps in terms of concentration with a polynomial function is as follows:

Fig. 5: The energy band gaps of BxGa1–xN as a function of boron composition.
Fig. 5:

The energy band gaps of BxGa1–xN as a function of boron composition.

Fig. 6: The energy band gaps of BxAl1–xN as a function of boron composition.
Fig. 6:

The energy band gaps of BxAl1–xN as a function of boron composition.

Fig. 7: The energy band gaps of AlxGa1–xN as a function of aluminum composition.
Fig. 7:

The energy band gaps of AlxGa1–xN as a function of aluminum composition.

BxGa1xNE(ΓΓ)= 3.23  3.08x+ 9.78x2(direct gap)E(ΓX)= 4.68 + 3.26x 1.92x2(indirect gap)BxAl1xNE(ΓΓ)= 5.82  7.93x+ 12.11x2(direct gap)E(ΓX)= 5.02 + 6.98x 6x2(indirect gap)AlxGa1xNE(ΓΓ)= 2.89 + 1.97x+0.60x2(direct gap)E(ΓX)= 4.58 + 8.09x 7.55x2(indirect gap)

BxGa1–xN has a phase transition from direct to indirect gap for high boron contents (x > 0.71). For BxAl1–xN, a direct gap is found for 0.05 < x <0.75. As for AlxGa1–xN, its gap has been found entirely direct.

We have plotted the energy band gaps for BxAlyGa1–xyN with the compositional parameter y taken to be 0 ≤ y ≤ 1 as shown in Figs. 8 and 9 and with the compositional parameter x being in the 0 ≤ x ≤ 1 range. Both figures show that the quaternary compounds have a direct gap. The energy gap varies from 4.13 to 4.96 eV for the composition x = 0.25, y = 0.25 and x = 0.50, y = 0.25. It is clearly seen that the incorporation of boron affects considerably the band gap compared to the aluminum incorporation. We have not found any theoretical report or experimental data in the literature to compare and confirm our results for these solid solutions. Our work therefore can serve as a reference for future studies.

Fig. 8: The calculated direct E(Γ–Γ) band gaps versus composition (x, y) for quaternary BxAlyGa1–x–yN solid solutions.
Fig. 8:

The calculated direct E(Γ–Γ) band gaps versus composition (x, y) for quaternary BxAlyGa1–xyN solid solutions.

Fig. 9: The calculated indirect E(Γ–X) band gaps versus composition (x,y) for quaternary BxAlyGa1–x–yN solid solutions.
Fig. 9:

The calculated indirect E(Γ–X) band gaps versus composition (x,y) for quaternary BxAlyGa1–xyN solid solutions.

3.3 Density of states

To understand the calculated band structure in terms of the contributing atomic states, the total DOS is plotted in Fig. 10 for B0.25Al0.25Ga0.50N. The DOS is computed using GGA approximation with a k mesh of 400 special k-points for B0.25Al0.25Ga0.50N. In the calculation of total and partial atomic DOS for B0.25Al0.25Ga0.50N, we have distinguished the B (1s2), Al (1s22s22p6), N (1s2), and Ga (1s22s22p63s23p6) inner-shell electrons from the valence electrons of B (2s2p1), Al (3s23p1), N(2s22p3), and Ga (3d104s24p1) shells. For ZB B0.25Al0.25Ga0.50N, the total DOS presents three regions: two regions in the valence band (VB1) and (VB2) below the Fermi level considered as energy reference and one region (CB) above the Fermi level EF. The DOS results from the contributions of different orbitals in each region.

Fig. 10: Total and partial DOS for ZB-type solid solutions B0.25Al0.25Ga0.50N.
Fig. 10:

Total and partial DOS for ZB-type solid solutions B0.25Al0.25Ga0.50N.

The first region (VB1) located in the energy range [–14, –10 eV] is dominated by the aluminum Al-3s, Al-3p, gallium Ga-4s, Ga-4p, and nitrogen N-2s, N-2p states with little contributions of Ga-3d states.

The second region (VB2) expanded between [–8, 0.0 eV] is divided into two sub-regions. The first region at the left side results from the contributions of B-2s, Al-3s, Ga-4s, and N-2s states. The second sub region at the right is a mixture of B-2p, Al-3p, Ga-4p, and N-2p states. We remark that the deep levels are dominated by the “s” state and the higher energy levels in this band are dominated by “p” states. The third region (CB) is essentially dominated by B-2p, Al-3p, and Ga-4p empty states with a minor contribution of N-2p states.

3.4 Optical properties

The optical properties may be extracted from the dielectric function,

ε(ω)=ε1(ω)+iε2(ω),

which is determined mainly by the transition between the valence and the conduction bands. According to perturbation theory, the imaginary part ε2(ω) is expressed as

ε2(ω)=e2πm2ω2V,CBZ|MCV(k)|2δ[ωCV(k)ω]d3k.

The integral is over the first BZ. The momentum dipole elements,

MCV(k)=uck|δ|uvk,

where δ is the potential vector defining the electric field, are matrix elements for direct transitions between valence uvk(r) and conduction band uck(r) states, and the energy

ωcv(k)=EckEvk

is the corresponding transition energy.

The real part ε1(ω) can be derived from the imaginary part using the familiar Kramers–Kronig transformations

ε1(ω)=1+2πP0ωε2(ω)ω2ω2dω

where P implies the principal value of the integral. The knowledge of both real and imaginary parts of the frequency dependent dielectric function allows the calculation of important optical functions such as the refractive index n(ω) using the following expression:

n(ω)=[ε1(ω)2+ε12(ω)+ε22(ω)2]1/2.

We have also calculated the refractive index n by using some empirical models. The Moss formula based on an atomic model [40],

Egn4=k,

and Herve and Vandamme’s empirical relation [41],

n=1+(AEg+B)2,

with A = 13.6 eV and B = 3.4 eV.

Fig. 11 display the real and the imaginary parts of the dielectric function as well refractive index spectrum for B0.25Al0.25Ga0.50N. Our analysis of the imaginary part of the dielectric function shows that the first critical point occurs at 3.1 eV for B0.25Al0.25Ga0.50N. This point represents the direct optical transitions between the top of the valence band and the bottom of the conduction band at Γ point. We can also remark that the main peaks in the spectra are located at about 7.5 eV for B0.25Al0.25Ga0.50N.

Fig. 11: Calculated refractive index: real and imaginary parts of the dielectric function of B0.25Al0.25Ga0.50N.
Fig. 11:

Calculated refractive index: real and imaginary parts of the dielectric function of B0.25Al0.25Ga0.50N.

4 Conclusion

Employing the FP-LAPW method, we studied the structural and electronic properties of the quaternary solid solutions BxAlyGa1–xyN and their three pseudo-binary ones as a function of the compositions x and y. A non-linear behavior of the lattice parameters, bulk moduli, and band gap dependence on x and y has been observed. The structural and electronic properties of the binary and ternary compounds are in fair good agreement with the available theoretical results. Contrary to the parent binary compounds, the quaternary solid solutions are found to be direct band gap semiconductors. This essential characteristics indicate that those materials (BxGa1–xN, BxAl1–xN, AlxGa1–xN, and BxAlyGa1–xyN) can be useful for optoelectronic applications. The calculated band gap energy of BAlGaN is in the range of 3.29 to 5.12 eV, corresponding to the wavelength range 377–242 nm. Therefore, the BAlGaN system is a promising material for use in semiconductor lasers that operate in the UV spectral region.


Corresponding author: Samir F. Matar, Institut de Chimie de la Matière Condensée de Bordeaux CNRS, University of Bordeaux, 33600 Pessac, France, e-mail:

References

[1] L. Djoudi, A. Lachebi, B. Merabet, H. Abid, Acta Phys. Polon.A2012, 122, 13.Search in Google Scholar

[2] P. Hohenberg, W. Kohn, Phys. Rev.1964, 136, 864.10.1103/PhysRev.136.B864Search in Google Scholar

[3] W. Kohn, L. Sham, Phys. Rev.1965, 140, A1133.10.1103/PhysRev.140.A1133Search in Google Scholar

[4] K. Schwarz, P. Blaha, G. K. H. Madsen, Comput. Phys. Commun.2002, 147, 71.10.1016/S0010-4655(02)00206-0Search in Google Scholar

[5] K. Schwarz, P. Blaha, Comput. Mater. Sci.2003, 28, 259.10.1016/S0927-0256(03)00112-5Search in Google Scholar

[6] J. Perdew, Y. Wang, Phys. Rev. B1992, 45, 13244.10.1103/PhysRevB.45.13244Search in Google Scholar

[7] J. Perdew, S. Burke, M. Ernzerhof, Phys. Rev. Lett.1996, 77, 3865.10.1103/PhysRevLett.77.3865Search in Google Scholar

[8] F. D. Murnaghan, Proc. Natl. Acad. Sci. USA1994, 30, 244.10.1073/pnas.30.9.244Search in Google Scholar

[9] M. B. Kanoun, A. E. Merad, J. Cibert, H. Aourag, Solid State Electron.2004, 48, 1601.10.1016/j.sse.2004.03.007Search in Google Scholar

[10] A. Zaoui, F. El Haj Hassan, J. Phys. Condens. Matter2001, 13, 253.10.1088/0953-8984/13/2/303Search in Google Scholar

[11] S. Matar, V. Gonnet, G. Demazeau, J. Phys. I1994, 4, 335.10.1051/jp1:1994141Search in Google Scholar

[12] V. L. Solozhenho in Properties of Group III Nitrides (Ed.: J. H. Edgar), INSPEC Publications, London, 1994, p. 43.Search in Google Scholar

[13] M. Grimsditch, E. S. Zouboulis, A. Polian, J. Appl. Phys.1994, 76, 832.10.1063/1.357757Search in Google Scholar

[14] Z. Boussahl, B. Abbar, B. Bouhafs, A. Tadjer, J. Solid State Chem.2005, 178, 2117.10.1016/j.jssc.2005.03.047Search in Google Scholar

[15] A. Trampert, O. Brandt, K. H. Ploog in Gallium Nitride I, Semiconductors and Semimetals, Vol. 50 (Eds.: J. I. Pankove, T. D. Moustakas), Academic Press, San Diego, 1998, p. 167.10.1016/S0080-8784(08)63088-4Search in Google Scholar

[16] R. Riane, Z. Boussahl, A. Zaoui, L. Hammerelaine, S. F. Matar, Solid State Sci.2009, 11, 200.10.1016/j.solidstatesciences.2008.06.001Search in Google Scholar

[17] C. Stamp, C. G. Van de Walle, Phys. Rev. B1999, 59, 5521.10.1103/PhysRevB.59.5521Search in Google Scholar

[18] V. L. Solozhenho in Properties of Group III Nitrides, Electronics Materials Information Service (EMIS) Data Reviews Series (Ed.: J. H. Edgar), Institution of Electrical Engineers, London, 1994, p. 10.Search in Google Scholar

[19] D. Vogel, P. Kruger, J. Pollmann, Phys. Rev. B1997, 55, 12836.10.1103/PhysRevB.55.12836Search in Google Scholar

[20] M. E. Sherwin, T. J. Drummond, J. Appl. Phys.1987, 69, 8423.10.1063/1.347412Search in Google Scholar

[21] A. Lachebi, H. Abid, Turk, J. Phys.2008, 32, 157.10.1192/pb.bp.108.019893Search in Google Scholar

[22] L. K. Teles, L. M. R. Scolfaro, J. R. Leite, J. Furthmüller, F. Bechstedt, Appl. Phys. Lett.2002, 80, 11777.10.1063/1.1450261Search in Google Scholar

[23] R. Riane, Z. Boussahla, S. F. Matar, A. Zaoui, Z. Naturforsch. 2008, 63b, 1069.10.1515/znb-2008-0909Search in Google Scholar

[24] V. V. Ilyasov, T. P. Zhdanova, I. Ya. Nikiforov, Phys. Solid State2005, 47, 1618.10.1134/1.2045343Search in Google Scholar

[25] B.-T. Liou, Appl. Phys. A2007, 86, 539.10.1007/s00339-006-3810-ySearch in Google Scholar

[26] D. Li, X. Zhang, Z. Zhu, H. Zhang, Solid State Sci.2011, 13, 1731.10.1016/j.solidstatesciences.2011.06.027Search in Google Scholar

[27] L. Vegard, Z. Kristallogr.1928, 67, 239.10.1524/zkri.1928.67.1.239Search in Google Scholar

[28] J. P. Dismuckes, L. Ekstrom, R. J. Poff, J. Phys. Chem. 1996, 68, 3021.10.1021/j100792a049Search in Google Scholar

[29] T. Takano, M. Kurimoto, J. Yamamoto, H. Kawanishi, J. Cryst. Growth2002, 237–239, 972.10.1016/S0022-0248(01)02026-7Search in Google Scholar

[30] A. Abdiche, A. Abid, R. Riane, A. Bouaza, Acta Phys. Pol. A2010, 117, 921.10.12693/APhysPolA.117.921Search in Google Scholar

[31] R. M. Wentzcovitch, R. J. Chang, N. L. Cohen, Phys. Rev. B1986, 34, 1071.10.1103/PhysRevB.34.1071Search in Google Scholar PubMed

[32] K. Karch, F. Bachstedh, Phys. Rev. B1997, 56, 7404.10.1103/PhysRevB.56.7404Search in Google Scholar

[33] X. M. Meng Chan, S. T. Lee, Appl. Phys. Lett.2004, 85, 1344.10.1063/1.1784545Search in Google Scholar

[34] F. Bassani, M. Yochimine, Phys. Rev. B1963, 130, 20.10.1103/PhysRev.130.20Search in Google Scholar

[35] M. Buongiorno Nardelli, K. Rapcewicz, E. L. Briggs, C. Bungaro, J. Bernholc in III–V Nitrides (Eds.: F. A. Ponce, T. D. Moustakas, I. Akasaki, B. A. Monemar), Materials Research Society Proceedings No. 449, Materials Research Society, Pittsburgh, PA, 1997, p. 893.10.1557/PROC-449-893Search in Google Scholar

[36] T. Lei, T. D. Moustakas, R. J. Graham, Y. He, S. J. Berkowitz, J. Appl. Phys.1992, 71, 4933.10.1063/1.350642Search in Google Scholar

[37] L. E. Ramos, L. K. Teles, L. M. R. Scolfaro, J. L. P. Castineira, A. L. Rosa, J. R. Leite, Phys. Rev. B2001, 63, 165210.10.1103/PhysRevB.63.165210Search in Google Scholar

[38] J.-C. Zheng, H.-Q. Wang, C. H. A. Huan, A. T. S. Wee, J. Phys. Condens. Matter2001, 13, 5295.10.1088/0953-8984/13/22/322Search in Google Scholar

[39] D. Koller, F. Tran, P. Blaha, Phys Rev. B2011, 83, 195134.10.1103/PhysRevB.83.195134Search in Google Scholar

[40] C. Ambrosch-Draxl, J. O. Sofo, Comp. Phys. Commun. 2006, 175, 1.10.1016/j.cpc.2006.03.005Search in Google Scholar

[41] T. S. Moss, Proc. Phys. Soc. LondonB1950, 63, 167.10.1088/0370-1301/63/3/302Search in Google Scholar

Received: 2015-9-26
Accepted: 2015-10-20
Published Online: 2016-1-7
Published in Print: 2016-2-1

©2016 by De Gruyter

Downloaded on 24.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2015-0149/html
Scroll to top button