You are on page 1of 13

Optik - International Journal for Light and Electron Optics 252 (2022) 168532

Contents lists available at ScienceDirect

Optik
journal homepage: www.elsevier.com/locate/ijleo

First-principles calculations to investigate ultra-wide bandgap


semiconductor behavior of NaMgF3 fluoro-perovskite with
external static isotropic pressure and its impact on
optical properties
Jalil Ur Rehman a, *, Muhammad Usman a, M. Bilal Tahir a, Abid Hussain a,
M. Awais Rehman a, Muhammad Sagir b, Hussein Alrobei c, Sami Ullah d,
Mohammed Ali Assiri d
a
Department of Physics, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan, Pakistan
b
Department of Chemical Engineering, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan, Pakistan
c
Department of Mechanical Engineering, College of Engineering, Prince Sattam Bin Abdulaziz University, Al Kharj, Saudi Arabia
d
Department of Chemistry, College of Science, King Khalid University, P.O.Box 9004, Abha 61413, Saudi Arabia

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

Keywords: The structural, electronic, and optical properties of cubic fluoro-perovskite NaMgF3 were inves­
Fluoro-perovskites tigated theoretically at pressures ranging from 0 to 100 GPa with a 10-step escalation, i.e., 0, 10,
Density functional theory 20, 30, 40, 50, 60, 70, 80, 90, and 100 GPa. The research is conducted by using density functional
Pressure-effects
theory (DFT) based CASTEP (Cambridge Serial Total Energy Package) code with ultra-soft
Elastic constants
Mechanical stability
pseudo-potential USP plane wave and Perdew Burke Ernzerhof (PBE) exchange-correlation
functional of Generalized Gradient Approximation (GGA). The band gap is found to be
increasing from 5.749 to 8.723 eV when the pressure is increased from 0 to 100 GPa. The nature
of the band gap remained indirect whether the pressure was applied or not. A significant incre­
ment was found in the band gap till 40 GPa as compared to higher pressure. Under the effect of
pressures ranging from 0 to 100 GPa, several optical properties like the refractive index, loss
function, reflectivity, and absorption coefficient have also been determined. The main peaks of
dielectric function (real and imaginary), reflectivity, loss function, and refractive index (n) are
observed to move towards higher values of energy with increasing pressure. A material that has a
high optical conductivity, absorption, and refractive index is ideal for photovoltaic applications.
Other optical technologies that use such material include waveguides, data storage medium, and
photonic crystals.

1. Introduction

Many materials with a similar crystalline structure have been discovered since the original discovery of perovskites in 1839, one of
the most common of which is CaTiO3. The ABX3 formula defines the fundamental structure of the cubic perovskites, in which cations

* Corresponding author.
E-mail address: jalil.rehman@kfueit.edu.pk (J. Ur Rehman).

https://doi.org/10.1016/j.ijleo.2021.168532
Received 7 June 2021; Received in revised form 23 December 2021; Accepted 23 December 2021
Available online 27 December 2021
0030-4026/© 2021 Elsevier GmbH. All rights reserved.
J. Ur Rehman et al. Optik 252 (2022) 168532

Fig. 1. 2 × 2 × 2 supercell of NaMgF3.

Table 1
The effect of pressure on lattice constants, volume and band gap of NaMgF3 after geometry optimization.
Pressure (GPa) (NaMgF3) Lattice constants (Å) (a = b = c) Volume (Å)3 Band gap (eV) Refs.

P=0 3.955 61.864 – [43]


3.833 56.314 – [44]
3.876 58.230 – [45]
3.836 56.446 – [46]
3.840 56.623 – [47]
P = 0 4.0009 64.043 5.749 Present Study
P = 10 3.837 56.490 6.335 Present Study
P = 20 3.729 51.853 6.787 Present Study
P = 30 3.655 48.827 7.155 Present Study
P = 40 3.595 46.462 7.416 Present Study
P = 50 3.545 44.550 7.692 Present Study
P = 60 3.498 42.802 7.926 Present Study
P = 70 3.458 41.350 8.154 Present Study
P = 80 3.424 40.142 8.357 Present Study
P = 90 3.392 39.027 8.551 Present Study
P = 100 3.363 38.035 8.723 Present Study

and anions are arranged in a specific order. Alkali metals or alkaline earth metals are the most common cations for A and B-site. In such
compounds, X is an anion in ABX3 and oxygen takes the place of X in oxide-perovskites [1–5]. On the other hand, the halogen pe­
rovskites in which X is replaced from a group VII-A element have attracted considerable attention in the recent times. These A and B
cations may be joined to the X in an esthetically pleasing but technically complex way. Using this stoichiometry, we can see how
temperature and pressure affect the electronic, magnetic, and optical characteristics [6–8]. Since fluoride perovskites have a wide
bandgap due to which these are the suitable candidates for use in a variety of optical applications, including lenses, ferroelectric and
antiferromagnetic systems, as well as optical coatings. Complex alkali metal fluorides have piqued scientists’ interest due to their
technical appeal for usage as fluorinating agents and catalysts in a variety of processes in organo-fluorochemical chemistry [9–12].
Photoluminescence is one of the many applications for fluoride-type perovskites such as photoluminescence [13–15],
high-temperature superconductivity [16], colossal magneto-resistivity (CMR) [17], and piezo-electricity [18]. These applications have
sparked a lot of interest in KCaF3, RbCaF3, BaLiF3, SrLiF3, and other members of this unusual class, but more theoretical and exper­
imental research is still needed [19]. Their magnetic properties, electronic structure, and optical characteristics are attracting a lot of
research since they offer promise in these areas. Superconductivity, piezoelectric, ferroelectricity, magneto resistance, and other
exceptionally favorable thermodynamic characteristics have all been explored in perovskites in order to find out how to use them in a
variety of applications [20–23]. A key feature of these materials is their wide bandgap, which allows them to be employed for different
applications, such as lenses, memory devices, sensors, and fuel cells. Such materials may also be used in LEDs and color TVs because of
their absorption properties. As long as perovskites are concerned, research has shown that by varying some parameters leads to a whole
set of varied properties. When it comes to perovskites, research has shown that changing a few factors can result in a wide range of
properties. We can learn how a material behaves by applying pressure or changing the temperature. When these parameters are

2
J. Ur Rehman et al. Optik 252 (2022) 168532

Fig. 2. (a) Band structure (b) TDOS of NaMgF3 at P = 0 GPa and (c) Band structure (d) TDOS of NaMgF3 at P = 10 GPa.

altered, we learn about changes in the elastic limit, sheer (shear) strength and bulk modulus as well as other features like as con­
ductivity, magnetic behavior, and energy band gap. A good example is the analysis of the compound KCaX3 (X = F, Cl) with the help of
linear density functional theory to explore the elastic, optical, and electronic properties of the compound. In KCaX3, an indirect
bandgap was discovered and the analysis also yielded a linear relationship between electronic bandgap and hydrostatic pressure [24].
The optimization approach for KCaF3 was used to get the bulk modulus, pressure, and lattice parameter values. The elastic moduli, as
well as the directional dependency of Young’s Moduli, appeared to be pressure-dependent, with pressures ranging from 0 to 40 GPa.
While a linear relationship between pressure and the Energy Band Gap was also shown in this study, several methodologies yielded
varying findings for the band gap value, but the major value is found 6.1 eV at R–Γ [25]. Another perovskite, BaZrO3, is well-known for
its various characteristics, which have been studied both with and without doping, as well as at high pressure [16]. When high pressure
was applied to BaZrO3, the cubic structure was converted into a tetragonal phase, according to an experimental study. The phase
transition was occurred at a pressure of 17.2 GPa [17]. Similarly, some other materials have also been studied and various properties of
the materials have been determined for different applications [26–30].
In the present study, we have investigated the structural parameters, electronic structure, and optical properties of NaMgF3 for the
first time under different pressure effects ranging from 0 to 100 GPa. The current study suggests that the material will show the same
behavior after 100 GPa. The altering values of lattice constants and band gap are observed with changing the pressure. The changes in
the conducting behavior and the optical properties of the compound are also observed under different pressures. The arrangement of
the article is as: Section 2 provides a brief description about the computational methodology; Section 3 is about the results and dis­
cussions while the Section 4 concludes the article.

2. Computational methodology

For the ternary compound NaMgF3, the cubic structure with the space group Pm3m was considered to investigate its properties.
The atomic configurations for the elements of NaMgF3 is as: 1s2 2s2, 2p6, 3s1 for Na; 1s2, 2 s2, 2p6, 3s2 for Mg and 1s2, 2s2, 2p5 for F. The
atomic positions were taken as (0.0, 0.0, 0.0), (0.5, 0.5, 0.5), and (0.5, 0.5, 0.0) for Na, Mg and F atoms, respectively. The Cambridge
Serial Total Energy Package (CASTEP) program [31] was utilized to determine the structural, electronic, and optical properties using
the ultra-soft pseudo-potential USP plane wave [32,33] and Perdew Burke Ernzerhof (PBE) exchange-correlation functional of
Generalized Gradient-Approximation (GGA) which are fast and more efficient than other approximations [34,35]. The Kohn-Sham
equations were used to do all of these calculations [36–38]. The creation of an ionic core is the result of nuclei interacting with
inner shell electrons. The valence electrons in the material and the ionic core then interact, causing the electron-ion potential to
converge. Integrations in the Brillion zone have been accomplished using customized k-points, with the k-points meshes built in the
order of 6 × 6 × 6 on the Monkhorst pack grid for optimizing the structure [39]. Throughout the calculations, the value of total energy
per atom is set as 2 × 10− 5 eV. Furthermore, during geometrical optimization, the convergence value of all the forces applied to the
atoms is kept 0.05 eV/Å. For the whole Brillouin zone, the value of the cutoff energy was set at 517.0 eV. The maximum ionic
displacement was kept in a range of 2 × 10− 3 Å during the geometry optimization. The BFGS algorithm was used for geometry
optimization and the equivalent hydrostatic pressure was increased from 0 to 100 GPa. Moreover, the Pulay density mixing scheme
was also used during the geometry optimization. At 0 GPa, the elastic constants were also calculated. Four number of steps for each
strain were taken during the calculation of elastic constants. Moreover, the maximum value of the strain amplitude was set to 0.003
and the total energy was taken as 4 × 10− 6 eV per atom. The value of maximum force was kept 1 × 10− 2 eV/Å and the maximum ionic

3
J. Ur Rehman et al.
4

Optik 252 (2022) 168532


Fig. 3. Band structures of NaMgF3 at pressures (GPa): (a) 20, (b) 30, (c) 40, (d) 50, (e) 60, (f) 70, (g) 80, (h) 90, (i) 100.
J. Ur Rehman et al. Optik 252 (2022) 168532

Fig. 4. Value of bandgaps of NaMgF3 with pressure showing an increasing trend.

displacement was set as 1 × 10− 4 Å for calculating the elastic constants. First of all, the geometry of the compound was optimized, and
then all properties of the material were investigated.

3. Results and discussions

3.1. Structure analysis

The 2 × 2 × 2 supercell of NaMgF3 is shown in Fig. 1. The geometry of the intended supercell of NaMgF3 was optimized first. The
overall energy of the crystal was maintained low by estimating the equilibrium lattice parameters with the help of Murnaghan state
equation, and the findings were obtained throughout a wide range of lattice constant values [40,41]. The total amount of energy is a
function of volume of the unit cell to the volume of the equilibrium cell V0. When the pressure is raised from 0 GPa to 100 GPa, the band
gap of NaMgF3 rises from 5.749 eV to 8.723 eV. The lattice parameters are found to have a decreasing value with increasing pressure,
which shows that the volume of the supercell will be decreased by increasing the pressure. A decline in the lattice parameter’s value
from 4.0009 Å to 3.363 Å occurs when pressure rises from zero to 100 GPa. The decrease in the lattice parameter with increasing the
pressure implies that there is present a strong atomic interaction. When the pressure is increased from 0 to 100 GPa, the volume of the
supercell drops from (64.043 Å)3 to (38.035 Å)3, indicating that the supercell has been sufficiently shrunk. Table 1 shows the impact of
pressure on the lattice parameter and volume. The determined lattice parameters, volume of the supercell and the band gap of the
material correspond well with the previous studies at 0 GPa. There is no theoretical or experimental study of this compound at high
pressure due to which the comparison of the results obtained at higher pressures is not possible. As a result, the future measurements
will corroborate all of the results calculated in this work. At 0 GPa, the thermodynamic and the mechanical stability of the compound is
determined with the help of formation energy and the elastic constants, respectively. The formation energy of the NaMgF3 compound
at 0 GPa is − 3.458 eV, which indicates that the material thermodynamically is stable. The values of the elastic constants are
determined at 0 GPa and found as: C11 = 104.26; C12 = 14.80; C44 = 36.31. The mechanical stability of the cubic crystal may be
determined with the help of Born stability criteria [42] as: C11 − C12 > 0; C11 + 2C12 > 0; C11 > 0; C44 > 0. We have C11 = 104.26 > 0;
C44 = 36.31 > 0; C11 − C12 = 89.46 > 0; C11 + 2C12 = 133.86 > 0. So, the NaMgF3 compound follows the Born stability criteria and
thus it is mechanically stable also.

3.2. Electronic band structure and density of states

The structure of electronic bands exposes the energy levels where electrons may exist (Energy band) as well as the energy levels in
which electrons are not accessible to occupy, i.e., there is zero availability of the electrons (band gap). There exist two types of the
energy bands: the conduction band (CB), which is located above the Fermi level (EF), and valence band (VB), which is located below
the EF. Because all of the calculations were done at 0 K, without taking into account the effects of finite temperature, therefore, the
highest point in the valence band is designated as the Fermi level. There is a direct band gap when the valence and conduction bands
are precisely aligned while it is indirect when these are not perfectly aligned. In our study, an indirect band gap is reported, no matter
how much the pressure is applied on the material. Fig. 2 depicts the band structure of NaMgF3 at 0 GPa and 10 GPa, with the total
density of states. In Fig. 3, the band structures are presented that were obtained at pressures ranging from 20 GPa to 100 GPa. The
minimum band gap can be observed in band structure under the pressure of 0 GPa. A rising trend in the band gap may be observed by

5
J. Ur Rehman et al. Optik 252 (2022) 168532

Fig. 5. TDOS of NaMgF3.

examining the band structures when the pressure increased from 0 GPa to 100 GPa. The largest band gap is shown in the band structure
at 100 GPa, whereas the smallest band gap is observed in the band structure at 0 GPa, demonstrating a direct relation between the
band gap and pressure. Due to a reduction in the lattice constant, the conduction band moves farther from the Fermi level which causes
the increment in the band gap of the material. Due to the higher value of the band gap, the material appears as a semiconductor with an

6
J. Ur Rehman et al. Optik 252 (2022) 168532

Fig. 6. PDOS of NaMgF3 (a) P = 0 GPa, (b) 10 GPa.

ultra-wide band gap with the indirect band gap nature. The total density of states (TDOS) and partial density of states (PDOS) may be
used to explain the increasing trend of the band gap. Fig. 4 illustrates the TDOS of NaMgF3 under pressure effects having a range from
0 to 100 GPa. When pressure is applied between 0 and 30 GPa, the major peak of TDOS exhibits a rising trend. At 0 GPa, its maximum
value is 12.33 and it reaches its highest value to 13.97 at 30 GPa. When pressure is increased further, its value goes on decreasing and
reaches its lowest value to 10.97 at 100 GPa. This main peak also slightly shifts towards the right side of the graph as pressure is
increased. At 0 GPa, this peak appears at − 20.18 eV and it appears at − 19.50 eV when temperature increases to 100 GPa. Fig. 5
shows the PDOS at 0 and 10 GPa while the Fig. 6 shows PDOS of NaMgF3 at 20–100 GPa. In Fig. 6, a vertical line represents the Fermi
level in PDOS at 0 and 10 GPa. The p-state is dominant in both below and above the Fermi level in PDOS shown in Fig. 6. Moreover, the
number of peaks in the p-state rises as the pressure increases. Therefore, the p-states provide substantially greater contributions than
the s- and d-states. By increasing the external applied pressure, the hybridization between the s- and p-states decreases, resulting in a
gradual increase in the band gap. When external pressure is applied, the valence band states move towards lower energy, while the
conduction band states shift towards higher energy, which is the fundamental cause for the widening of the bandgap. With the help of
above observations, we may conclude that the PDOS shifts between the higher and lower energies as a consequence of varying
pressures, leading to a rise in the band gap, which also impacted the optical properties described below.

3.3. Optical properties of NaMgF3 under static isotropic pressure

The electronic structure of the compounds may be explained with the help of its optical characteristics, such as energy loss function,
absorption coefficient, reflectivity, and index of refraction, as well as relative permittivity. These properties are very helpful in sug­
gesting the appropriateness and feasibility of the material in the nanoelectronics and optoelectronics [48]. All of these optical
properties are resulted due to the interaction between the waves (electromagnetic waves) and matter (material). As these properties
are interconnected to one another, therefore, the complex dielectric function is utilized which is given as:

ε(ω) = ε1(ω) + iε2(ω) (1)

The Eq. (1) represents the real and imaginary components of the dielectric function by ε1(ω) and iε2(ω), respectively. The dielectric
constant also refers to the relative permittivity of the material. If we take the word literally, it refers to how much an electric field is
permitted to enter the material. This basically refers that how much polarization a material will tolerate. As a result, a perfect electrical
conductor would have a value of 0 because no field can be formed within its boundaries [40]. The real component of the dielectric
function ε1(ω) gives information about the polarization intensity of the material. The main peak of the dielectric function at 0 GPa
appears at 6.35 eV. It seems to move towards a higher value of energy as the pressure is increased. It appears at 7.25, 7.59, 8.09, 8.49,
12.25, 13.97, 14.38, 14.68, 14.92, 15.27 eV when pressure was kept 10, 20, 30, 40, 50, 60, 70, 80, 90, 100 GPa, respectively. The
optical characteristics are connected to energy dissipation and are influenced by the complex component of the dielectric function
ε2(ω). The complex part of the dielectric function is shown in Fig. 7(b) an its value appears to be 0 at 0 GPa (Table 2).
The value of the complex component of the dielectric function also remains zero at all pressures from 0 to 100 GPa. Its main peak
appears at 16.86 eV at 0 GPa. As the pressure is steadily raised, this peak shifts slightly to the higher value and also towards the higher
energy. The value of the main peak is 1.75 at 0 GPa, 2.14 at 50 GPa, and 2.55 at 100 GPa. For this imaginary part of the dielectric
function, the main peak appears at 17.67, 18.42, 18.76, 19.00, 19.14, 19.48, 19.65, 19.84 20.05 and 20.38 eV when the pressure was
10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 GPa respectively. All peaks slightly move to the higher energy when the pressure rises
gradually and the reflectivity climbs in lockstep. The equations are given below [49], which are used to calculate the additional optical

7
J. Ur Rehman et al.
8

Optik 252 (2022) 168532


Fig. 7. (a–i): PDOS of NaMgF3 under pressure varying from 20 to 100 GPa.
J. Ur Rehman et al. Optik 252 (2022) 168532

Table 2
The static value of ε1 (ω) and n (ω) at 0 eV and major peaks of ε2 (ω), I(ω), L(ω) and R(ω).
Pressure Static value of Main peak, ε2 (ω) Absorption peak, I (ω) Static refractive index Loss function peak L Reflectivity peak, R (ω)
(GPa) ε1 (ω) (eV) (eV) n (ω) (ω) (eV) (eV)

0 1.7398 16.7140 17.5672 1.3199 21.9395 18.6336


10 1.7593 17.6471 18.7669 1.3253 22.4461 19.2468
20 1.7631 18.5269 19.8603 1.3265 24.3656 22.8593
30 1.7649 18.7669 20.6865 1.3283 25.2188 23.8858
40 1.7764 19.0068 21.3796 1.3332 25.5921 24.3923
50 1.7762 19.2468 21.9662 1.3332 26.4185 24.9788
60 1.7876 19.4867 22.4461 1.3367 26.8995 25.6987
70 1.7637 19.7267 22.8193 1.3263 26.9784 26.4587
80 1.7601 19.8602 24.4989 1.3282 27.0717 26.8984
90 1.7717 20.0733 24.8455 1.3300 27.4850 27.1584
100 1.7753 20.4192 25.1255 1.330 27.8792 27.7516

properties such as refractive index n(ω), energy loss L(ω), absorption coefficient I(ω) and reflectivity R(ω) and the associated results are
obtained for NaMgF3 with and without pressure ranging from 0 to 100 GPa (Figs 8 and 9).

n (ω) = [ε1(ω)/2 + {ε12(ω) + ε12(ω)}1/2/2]1/2 (2)

L (ω) = -Im(ε(ω)− 1) = ε2(ω)/ε1(ω)2 + ε2(ω)2 (3)


1/2 2
I (ω) = 2 ω[{ε1 (ω) + ε1 (ω)} 2 1/2
− ε1(ω)] 1/2
(4)

R (ω) = (n + ꜟ k − 1)/((n + ꜟ k + 1) (5)

The reflectivity of any compound can be used to analyze its surface behavior. The reflectivity spectrums for NaMgF3 at 0, 10, 20, 20,
30, 40, 50, 60, 70, 80, 90 and 100 GPa is shown in Fig. 7(f). At zero frequency, the value of reflectivity for NaMgF3 is equal to 0.0191,
0.0199, 0.0193, 0.0197, 0.0204, 0.0208, 0.0208, 0.0188, 0.0203, 0.0203 and 0.0203 when pressure was kept 0, 10, 20, 30, 40, 50,
60,70, 80, 90, 100 GPa respectively. The maximum reflectivity peaks are obtained at 18.42, 19.24, 22.92, 23.78, 24.43, 24.87, 25.73,
26.41, 26.89, 27.37 and 27.65 eV at 0, 10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 GPa respectively. These values show that the
reflectivity peak is slightly shifting towards the higher value of frequency as the pressure is increasing. The function of energy loss L(ω)
is an essential factor that indicates how much energy is lost by the electron as it moves through the material. The graph for this function
is shown in Fig. 7(e). The relevant frequency is referred to as the plasma frequency because the energy loss function L(ω) has the peaks
that accurately describe characteristics associated with plasma resonance. The value of its highest peak at 0 GPa is 1.84 and it appears
at 21.86 eV. As the pressure rises from 0 to 100 GPa, it also moves to a higher energy state. This main peak appears at 26.31 eV and at
27.89 eV when pressure is kept 50 and 100 GPa respectively. Under the influence of zero and higher pressures, the value of the static
refractive index is also computed. The value of n(ω) is 1.3199 at zero photon energy under the effect of zero pressure. Its value is 1.3317
and 1.3319 when pressure was 50 and 100 GPa respectively under the influence of zero photon energy. There is a minor shift in n(ω)
before and after applying pressure, which favors the semiconducting nature of the ternary compound. There will be no transmittance at
high energies (lower wavelengths) since all of the light will be absorbed. However, because there are no relevant electronic transitions
at low energies (longer wavelengths), transmission will be quite strong in this range [50]. A material with high optical conductivity,
absorption, and refractive index will perform well in photovoltaic systems. It is also utilized in waveguides, data storage medium,
photonic crystals, and other optical technologies [51].

4. Conclusion

Using the density functional theory (DFT)-based CASTEP code, this study examined the effects of externally applied pressure
ranging from 0 GPa to 100 GPa with a 10-step increment, i.e. 0, 10, 20, 30, 40, 50, 60, 70, 80, 90, and 100 GPa on the structural,
electronic, and optical properties of NaMgF3 using ultra-soft pseudopotential (USP) plane wave and the Perdew Burke Ernzerhof (PBE)
exchange-correlation functional of Generalized-Gradient-Approximation (GGA). After geometry optimization, the structural param­
eters and the cell volume are determined, which are equivalent to previously reported values at 0 GPa. As pressure is applied from 0 to
100 GPa, the electronic band gap rises from 5.749 to 8.723 eV, indicating an increase in the band gap. The significant rise in the band
gap is reported in a pressure range of 10–40 GPa. The band gap slightly changes with an increase in the applied external pressure above
40 GPa. Furthermore, an indirect band gap is reported for all values of applied external pressure. The TDOS and PDOS have been
calculated under different applied external pressures. Under the influence of external pressure within a range of 0–100 GPa, several
optical properties such as the energy-loss function, refractive index, reflectivity, and absorption function also have been determined.
The main peaks of dielectric function (real and imaginary), reflectivity, loss function, and refractive index (n) are observed to move
towards higher values of energy with increasing the pressure. The performance of photovoltaic systems will be enhanced by the use of a
material with high optical conductivity, absorption, and refractive index. It can also be used in the optical devices like waveguides,
data storage media, and photonic crystals.

9
J. Ur Rehman et al.
10

Fig. 8. Optical properties of NaMgF3 with inclusion and exclusion of a variety of pressure. (a, b) Dielectric function consisting of real and imaginary part, (c) absorption spectra, (d) refractive index, (e)
energy-loss function and (f) reflectivity.

Optik 252 (2022) 168532


J. Ur Rehman et al.
11

Fig. 9. Trend of different optical properties (a) ε1(ω), (b) ε2(ω), (c) I(ω), n(ω), (e) L(ω), and(f) R(ω) versus external isotropic pressure. The values of ε1(ω) and n(ω) are taken at 0 eV.

Optik 252 (2022) 168532


J. Ur Rehman et al. Optik 252 (2022) 168532

Declaration of Competing Interest

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

Acknowledgements

The author Dr. Mohammed Ali Assiri would like to acknowledge the efforts of King Khalid University, Saudi Arabia (Deanship of
Scientific Research) for support through the Research Groups Project under the Grant no. R.G.P.2/153/42.

References

[1] A.A. Mubarak, A.A. Mousa, The electronic and optical properties of the fluoroperovskite BaXF3 (X = Li, Na, K, and Rb) compounds, Comput. Mater. Sci. 59
(2012) 6–13.
[2] K. Bidai, M. Ameri, S. Amel, I. Ameri, Y. Al-Douri, D. Varshney, C.H. Voon, First-principles calculations of pressure and temperature dependence of
thermodynamic properties of anti-perovskite BiNBa3 compound, Chin. J. Phys. 55 (5) (2017) 2144–2155.
[3] K. Chakraborty, M.G. Choudhury, S. Paul, Numerical study of Cs2TiX6 (X = Br− , I− , F− and Cl− ) based perovskite solar cell using SCAPS-1D device simulation,
Sol. Energy 194 (2019) 886–892.
[4] N. Moulay, M. Ameri, Y. Azaz, A. Zenati, Y. Al-Douri, I. Ameri, Predictive study of structural, electronic, magnetic and thermodynamic properties of XFeO3
(X = Ag, Zr and Ru) multiferroic materials in cubic perovskite structure: first-principles calculations, Mater. Sci. Pol. 33 (2) (2015) 402–413.
[5] R. Arar, T. Ouahrani, D. Varshney, R. Khenata, G. Murtaza, D. Rached, A.H. Reshak, Structural, mechanical and electronic properties of sodium based
fluoroperovskites NaXF3 (X = Mg, Zn) from first-principles calculations, Mater. Sci. Semicond. Process. 33 (2015) 127–135.
[6] F. Litimein, R. Khenata, A. Bouhemadou, Y. Al-Douri, S.B. Omran, First-principle calculations to investigate the elastic and thermodynamic properties of R BRh3
(R = Sc, Y and La) perovskite compounds, Mol. Phys. 110 (2) (2012) 121–128.
[7] S. Feng, F. Guo, Y. Zhang, F. Miao, Z. Wang, C. Yuan, K. Yang, Structural evolution, lattice dynamics, electronic and thermal properties of VH2 under high
pressure, Solid State Commun. 330 (2021), 114287.
[8] L. Salik, A. Bouhemadou, K. Boudiaf, F.S. Saoud, S. Bin-Omran, R. Khenata, A.H. Reshak, Structural, elastic, electronic, magnetic, optical, and thermoelectric
properties of the diamond-like quaternary semiconductor CuMn2InSe4, J. Supercond. Nov. Magn. 33 (4) (2020) 1091–1102.
[9] J. Li, X. Zhang, Z. Fang, X. Cao, Y. Li, C. Sun, F. Yin, First-principles calculations to investigate electronic, elastic, and optical properties of one dimensional
electride Y5Si3, Results Phys. (2021), 104615.
[10] M. Bouchenafa, A. Benmakhlouf, M. Sidoumou, A. Bouhemadou, S. Maabed, M. Halit, Y. Al-Douri, Theoretical investigation of the structural, elastic, electronic,
and optical properties of the ternary tetragonal tellurides KBTe2 (B = Al, In), Mater. Sci. Semicond. Process. 114 (2020), 105085.
[11] S. Touam, R. Belghit, R. Mahdjoubi, Y. Megdoud, H. Meradji, M.S. Khan, Y. Al-Douri, First-principles computations of $$\hbox {Y} _ {x}\hbox {Ga} _ {1-{x}} $$
YxGa1-x As-ternary alloys: a study on structural, electronic, optical and elastic properties, Bull. Mater. Sci. 43 (1) (2020) 1–11.
[12] P. Sehgal, A.K. Narula, Improved optical, electrochemical and photovoltaic properties of dye-sensitized solar cell composed of rare earth-doped zinc oxide,
J. Mater. Sci.: Mater. Electron. (2021) 1–11.
[13] S. Xue, J. Wang, Q. Wu, L. Zhang, R. Dai, B. Tian, F. Zhang, The electronic and optical properties of Ni-doped Bi4O5I2: first-principles calculations, Results Phys.
19 (2020), 103596.
[14] Z. Cui, M. Wang, N. Lyu, S. Zhang, Y. Ding, K. Bai, Electronic, magnetism and optical properties of transition metals adsorbed puckered arsenene, Superlattices
Microstruct. 152 (2021), 106852.
[15] M.A. El-Morsy, A novel algorithm based on sub-fringe integration method for direct two-dimensional unwrapping phase reconstruction from the intensity of
one-shot two-beam interference fringes, Appl. Phys. B 125 (11) (2019) 1–16.
[16] Y. Al-Douri, N. Amrane, M.R. Johan, Annealing temperature effect on structural and optical investigations of Fe2O3 nanostructure, J. Mater. Res. Technol. 8 (2)
(2019) 2164–2169.
[17] Z. Souadia, A. Bouhemadou, S. Bin-Omran, R. Khenata, Y. Al-Douri, S. Al Essa, Electronic structure and optical properties of the dialkali metal monotelluride
compounds: ab initio study, J. Mol. Graph. Model. 90 (2019) 77–86.
[18] J. ur Rehman, M. Usman, M.B. Tahir, A. Hussain, M.A. Rehman, N. Ahmad, S. Muhammad, First-principles calculations to investigate structural, electronic and
optical properties of Na based fluoroperovskites NaXF3 (X = Sr, Zn), Solid State Commun. (2021), 114396.
[19] A. Bouhemadou, D. Allali, K. Boudiaf, B. Al Qarni, S. Bin-Omran, R. Khenata, Y. Al-Douri, Electronic, optical, elastic, thermoelectric and thermodynamic
properties of the spinel oxides ZnRh2O4 and CdRh2O4, J. Alloy. Compd. 774 (2019) 299–314.
[20] M.R. Kabli, J. ur Rehman, M.B. Tahir, M. Usman, A.M. Ali, K. Shahzad, Structural, electronics and optical properties of sodium based fluoroperovskites NaXF3
(X = Ca, Mg, Sr and Zn): first principles calculations, Phys. Lett. A (2021), 127574.
[21] J.U. Rehman, M.A. Rehman, M.B. Tahir, A. Hussain, T. Iqbal, M. Sagir, M. Alzaid, Electronic and optical properties of nitrogen and sulfur doped strontium
titanate as efficient photocatalyst for water splitting: a DFT study, Int. J. Hydrog. Energy (2021).
[22] A. Gherriche, A. Bouhemadou, Y. Al-Douri, S. Bin-Omran, R. Khenata, M.A. Hadi, Ab initio exploration of the structural, elastic, electronic and optical properties
of a new layered perovskite-type oxyfluoride: CsSrNb2O6F, Mater. Sci. Semicond. Process. 131 (2021) 105890–105900.
[23] A.H. Reshak, M.S. Abu-Jafar, Y. Al-Douri, Two symmetric n-type interfaces SrTiO3/LaAlO3 in perovskite: electronic properties from density functional theory,
J. Appl. Phys. 119 (24) (2016), 245303-205310.
[24] S. Alnujaim, A. Bouhemadou, A. Bedjaoui, S. Bin-Omran, Y. Al-Douri, R. Khenata, S. Maabed, Ab initio prediction of the elastic, electronic and optical properties
of a new family of diamond-like semiconductors, Li2HgMS4 (M = Si, Ge and Sn), J. Alloy. Compd. 843 (2020) 155991–156004.
[25] S. Hadji, A. Bouhemadou, K. Haddadi, D. Cherrad, R. Khenata, S. Bin-Omran, Y. Al-Douri, Elastic, electronic, optical and thermodynamic properties of
Ba3Ca2Si2N6 semiconductor: first-principles predictions, Phys. B: Condens. Matter 589 (2020) 412213–441221.
[26] M.A. Benali, H.T. Derraz, I. Ameri, A. Bourguig, A. Neffah, R. Miloua, Y. Al-Douri, Synthesis and analysis of SnO2/ZnO nanocomposites: structural studies and
optical investigations with Maxwell–Garnett model, Mater. Chem. Phys. 240 (2020) 122254–122260.
[27] Y. Al-Douri, M. Ameri, A. Bouhemadou, K.M. Batoo, First-principles calculations to investigate the refractive index and optical dielectric constant of Na3SbX4
(X = S, Se) ternary chalcogenides, Phys. Status Solidi (b) 256 (11) (2019) 1900131–1900134.
[28] Y. Al-Douri, A.A. Odeh, A.S. Ibraheam, Transition metals doped In2S3 nanostructure: structural and optical features, Mater. Res. Express 6 (12) (2020)
125914–125922.
[29] S. Sagadevan, S. Vennila, A.R. Marlinda, Y. Al-Douri, M.R. Johan, J.A. Lett, Synthesis and evaluation of the structural, optical, and antibacterial properties of
copper oxide nanoparticles, Appl. Phys. A 125 (8) (2019) 489–497.
[30] J.U. Rehman, M. Usman, M.B. Tahir, A. Hussain, M. Rashid, Investigation of structural, electronics, optical, mechanical and thermodynamic properties of
YRu2P2 compound for superconducting application, J. Supercond. Nov. Magn. (2021) 1–9.
[31] M.D. Segall, P.L.D. Lindan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J. Clark, M.C. Payne, First principles simulation: ideas, illustrations and the CASTEP code,
J. Phys. Condens. Matter 14 (2002) 2717.
[32] Georg K.H. Madsen, Peter Blaha, Efficient linearization of the augmented plane-wave method, Phys. Rev. B 64 (2001) 1951341–1951349.
[33] J.P. Perdew, K. Burke, M. Ernzerhof, generalized gradient approximation made simple, Phys. Rev. Lett. 77 (1996) 3865.

12
J. Ur Rehman et al. Optik 252 (2022) 168532

[34] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C. Fiolhais, Atoms, molecules, solids, and surfaces: applications of the
generalized gradient approximation for exchange and correlation, Phys. Rev. B 46 (1992) 6671.
[35] P. Hohenberg, W. Kohn, Inhomogeneous electron gas, Phys. Rev. 136 (1964) 864.
[36] W. Kohn, L.J. Sham, Self-consistent equations including exchange and correlation effects, Phys. Rev. A 140 (1965) 1133.
[37] W. Kohn, Nobel lectures: electronic structure of matter-wave functions and density functionals, Rev. Mod. Phys. 71 (1999) 1253.
[38] Hendrik J. Monkhorst, James D. Pack, Special points for Brillion-zone integrations, Phys. Rev. B 13 (1976) 5188–5192.
[39] F.D. Murnaghan, The compressibility of media under extreme pressures, Proc. Natl. Acad. Sci. USA 30 (9) (1944) 244–247.
[40] F. Birch, Finite elastic strain of cubic crystals, Phys. Rev. 71 (11) (1947) 809–824.
[41] F. Mouhat, F.X. Coudert, Necessary and sufficient elastic stability conditions in various crystal systems, Phys. Rev. B 90 (22) (2014), 224104.
[42] E.C.T. Chao, H.T. Evans Jr., B.J. Skinner, C. Milton, Neighborite, NaMgF3, a new mineral from the green river formation, South Ouray, Utah, Am. Mineral.: J.
Earth Planet. Mater. 46 (3–4_Part_1) (1961) 379–393.
[43] Y. Zhao, D.J. Weidner, J.B. Parise, D.E. Cox, Thermal expansion and structural distortion of perovskite—data for NaMgF3 perovskite. Part I, Phys. Earth Planet.
Inter. 76 (1–2) (1993) 1–16.
[44] J. Chen, H. Liu, C.D. Martin, J.B. Parise, D.J. Weidner, Crystal chemistry of NaMgF3 perovskite at high pressure and temperature, Am. Mineral. 90 (10) (2005)
1534–1539.
[45] V. Luana, A. Costales, A.M. Pendás, Ions in crystals: the topology of the electron density in ionic materials. II. The cubic alkali halide perovskites, Phys. Rev. B 55
(7) (1997) 4285.
[46] G.A. Geguzina, Ferroelectrics-magnetics with the perovskite structure, Ferroelectrics 568 (1) (2020) 85–94.
[47] N. Badi, Y. Al-Douri, S. Khasim, Effect of nitrogen doping on structural and optical properties of MgxZn1-xO ternary alloys, Opt. Mater. 89 (2019) 554–558.
[48] G. Hougham, G. Tesoro, A. Viehbeck, J.D. Chapple-Sokol, Polarization effects of fluorine on the relative permittivity in polyimides, Macromolecules 27 (21)
(1994) 5964–5971.
[49] R. Terki, H. Feraoun, G. Bertrand, H. Aourag, Full potential calculation of structural, elastic and electronic properties of BaZrO3 and SrZrO3, Phys. Status Solidi
(b) 242 (5) (2005) 1054–1062.
[50] M. Naser A., S. Zaliman, H. Uda, A.D. Yarub, Investigation of the absorption coefficient, refractive index, energy band gap, and film thickness for Al0.11Ga0.89N,
Al0.03Ga0.97N, and GaN by optical transmission method, Int. J. Nanoelectron. Mater. 2 (2009) (2009) 189–195.
[51] Jin Yang, Lijan Yang, Theoretical investigation of electronic structure, optical, elastic, hardness and thermodynamics properties of jadeite, Mater. Sci. Semicond.
Process. 31 (2015) 509–516.

13

You might also like