Next Article in Journal
Preparation, Thermal Behavior, and Conformational Stability of HMX/Cyclopentanone Cocrystallization
Previous Article in Journal
An Enhanced Deep Learning-Based Pharmaceutical Crystal Detection with Regional Filtering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ca, Ba, Be, Mg, and Sr Substitution on Electronic and Optical Properties of XNb2Bi2O9 for Energy Conversion Application Using Generalized Gradient Approximation–Perdew–Burke–Ernzerhof

1
Department of Physics, The University of Lahore, Sargodha Campus, Sargodha 40100, Pakistan
2
Department of Physics, Fatima Jinnah Women University Rawalpindi, Rawalpindi 46000, Pakistan
3
Department of Physics, Division of Science and Technology, University of Education, Lahore 54000, Pakistan
4
Department of Chemistry, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(8), 710; https://doi.org/10.3390/cryst14080710
Submission received: 30 June 2024 / Revised: 13 July 2024 / Accepted: 5 August 2024 / Published: 7 August 2024
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
Bismuth layered structure ferroelectrics (BLSFs), also known as Aurivillius phase materials, are ideal for energy-efficient applications, particularly for solar cells. This work reports the first comprehensive detailed theoretical study on the fascinating structural, electronic, and optical properties of XNb2Bi2O9 (X = Ca, Ba, Be, Mg, Sr). The Perdew–Burke–Ernzerhof approach and generalized gradient approximation (GGA) are implemented to thoroughly investigate the structural, bandgap, optical, and electronic properties of the compounds. The optical conductivity, band topologies, dielectric function, bandgap values, absorption coefficient, reflectivity, extinction coefficient, refractive index, and partial and total densities of states are thoroughly investigated from a photovoltaics standpoint. The material exhibits distinct behaviors, including significant optical anisotropy and an elevated absorption coefficient > 105 cm−1 in the region of visible; ultraviolet energy is observed, the elevated transparency lies in the visible and infrared regions for the imaginary portion of the dielectric function, and peaks in the optical spectra caused by inter-band transitions are detected. According to the data reported, it becomes obvious that XNb2Bi2O9 (X = Ca, Ba, Be, Mg, and Sr) is a suitable candidate for implementation in solar energy conversion applications.

1. Introduction

In recent few years, research engineers and scientists have focused strongly on photovoltaic (PV) technology in terms of sustainability, scalability, and environmental impact to reduce the world’s reliance on fossil fuels for energy generation [1]. Photovoltaic technology supports power space satellites, electrical energy generation, large-scale utilities, and smaller scales like electric systems and home solar well. In solar desalination and solar car applications, various innovative inventions, including PV cells, have also been patented. Due to the technology’s proven benefits and numerous applications, scientists working in the field of renewable energy applications are encouraged to enhance and speed up the research on PV technology by seeking out new materials/methodologies to bring out the main advantages of the devices in the form of higher efficiency, while developing new structures/morphologies can also be beneficial to achieve interesting performances [2,3,4,5,6]. The emerging PV technology has attracted the interest of scientists, and are divided into the following families: perovskite solar cells [7], thin-film amorphous silicon solar cells [8], quantum dot solar cells [9], dye-sensitized solar cells [10], gallium copper indium selenide solar cells [11], cadmium telluride solar cells [12], organic solar cells [13], etc.
Recently, bismuth layered structural ferroelectrics (BLSFs) have exhibited strong activity under visible light, which has garnered significant attention, specifically for the Aurivillius family, which contains Bi as a component element. There are two factors that maintly express the importance of Bi existence in BLSFs [14]. Firstly, it increases the probability of visible light absorption by raising the valence band (VB) edge due to the hybridization of Bi-6s with O-2p orbitals. Secondly, the remarkable mobility of the photogenerated holes is caused by the widely dispersed Bi-6s and O-2p-hybridized VB [15].The Aurivillius family of oxides is renowned for being a type of layered perovskite that is created by a regular intergrowth of perovskite (An−1BnO3n+1)2− blocks and fluorite-like (Bi2O2)2+ layers; n is the number of corner-connected octahedral layers. Bi2WO6 and Bi2MoO6 are the prominent examples of n = 1 members of the Aurivillius family that have been thoroughly investigated as visible-light-driven photocatalysts. A recent article outlined the performances of the five-layer Aurivillius phase (Bi6Ti3Fe2O18) to examine the impact of La substitution on the structure, phase formation, and photocatalysis. The study concluded that single-semiconductor oxide photocatalysts without heterojunction formation are higher-ordered Aurivillius-layered perovskites [16,17,18,19]. Cui et al. computed the Aurivillius family’s members Bi4Ti3O12 (BTO) and Sr2Bi4Ti5O18 (SBT), which revealed that the BTO sample’s bandgap is 2.51 eV, whereas the SBT sample’s is 2.23 eV. These two materials enable the theoretical framework of the Aurivillius family to be utilized in sensor, optical, and random access memory applications [20]. Lardhi et al. showed a theoretical examination of compounds containing bismuth titanate (Bi12TiO20 and Bi4Ti3O12 materials). Bi12TiO20 offered lower effective masses, a greater absorption coefficient, a band edge, and improved charge transport properties compared to Bi4Ti3O12. Titanate is recognized for its strong optoelectronic properties, which should be used in the formation of perovskite-based photovoltaic or photocatalytic applications [21]. Werner et al. have reported structural and optical investigations of La2.1Bi2.9Ti2O11Cl. The electronic structure of La2.1Bi2.9Ti2O11Cl is comparable to that of other double-layered oxyhalides of the Sillén–Aurivillius type of compounds. Visible light sensitivity can be achieved by a narrow bandgap of 2.8 eV, which is mainly due to the contribution of Bi-6p orbitals to the conduction band. These orbitals are primarily located at the Bi site, and Bi is adjacent to the perovskite layer [22]. Liu et al. have designed the Aurivillius-type Sn-based halide perovskites Ba2X2[Csn−1SnnX3n+1] with X = I/Br/Cl, where the air’s oxygen is blocked by the [Ba2X2] layer, enhancing the crystals’ inoxidizability. The direct bandgaps (0.84–2.20 eV) of Ba2X2[Csn−1SnnX3n+1] meet the requirements for both single- and multi-junction perovskite solar cells (PSCs). With an ideal bandgap (1.26 eV), strong carrier mobility (135–173 cm2 V−1 s−1), and an ideal absorption coefficient (~105 cm−1), Ba2Br2[Cs2Sn3Br10] was reported to be the best candidate for the optical applications [23]. Xu et al. analyzed the electronic structure, ferroelectricity, and optical characteristics of CaBi2Ta2O9. Its indirect bandgap was reported to be 3.114 eV, and its numerous optical characteristics were outlined [24]. Stachiotti et al. reported the extremely accurate band structure calculations of SrBi2Ta2O9 with a paraelectric structure and presented the presence of ferroelectricity in it [25]. Given the above, it is evident that Aurivillius phase materials are gaining the interest of researchers due to their intriguing photovoltaic properties. To achieve calculations of the electronic and optical properties of BLSF’s family with reasonable accuracy, different approaches based on DFT, such as the local density approximation (LDA), LDA-1/2, full potential linearized augmented plane wave (FP-LAPW), modified Becke–Johnson (mBJ), GGA, and the Heyd–Scuseria–Ernzerhof (HSE06) function, can be employed [26,27,28].
Inspired by the above innovative concepts relating to BLSF’s family, with relatively few theoretical studies on it in the literature, XNb2Bi2O9 (X = Ca, Ba, Be, Mg, and Sr) has been selected to be comprehensively explored. Therefore, it is essential to thoroughly investigate the structural, electrical, and optical characteristics of XNb2Bi2O9 (X = Ca, Ba, Be, Mg, and Sr) compounds using the FP-LAPW approach with the DFT. The primary goal of this study is to use PBE-GGA exchange correlation potentials to carry out a complementary and comparative analysis of the ground state characteristics of the compounds. Having applications in fields of physics, material sciences, and also in chemistry, DFT-based calculations are among the cheapest, environmentally friendly, and time-saving computational approaches for evaluating the electronic and optical characteristics of different compounds.

2. Methodology

A DFT-based simulation was performed to examine the electronic properties [bandgap, band structure, the density of states (DOS), and partial density of states (PDOS)] and optical properties [reflectivity, refractive index, extinction coefficient, absorption, dielectric function (real and imaginary), optical conductivity (real and imaginary), and loss function] of the Aurivillius phase materials by using a framework named CASTEP (Cambridge Sequential Total Energy Package) [29], a first-principle plane-wave pseudopotential code for quantum mechanics-based simulations on BIOVIA Material Studio.
We settled the exchange correlation functional, as GGA provides better information regarding electronic subsystems, with PBE (Perdew–Burke–Ernzerhof) acting as the default exchange correlation functional [30,31,32,33]. For optimization convergence tolerance, the value of energy used to be 2.0 × 10−5 with a maximum force of 0.05 eV atom−1 and a maximum stress of 0.1 GPa, with a maximum displacement of 0.002 Å. For calculations, the OTFG ultrasoft pseudopotential for minimization of error concerning a fully converged all-electron, along with Koelling–Hamon as a relativistic treatment, were used. For the basic set, a parameter energy cut-off value of 1200 eV with 2 × 2 × 2 set of k-points of Brillouin zone was used, the SCF tolerance was set to be 2.0 × 10−6, with density mixing as an electronic minimizer being used for calculations of the electronic and optical properties of the given structures.

3. Results and Discussions

3.1. Structural Properties

The crystal structure of BLSFs possesses the conventional formula (Bi2O2)2+(An−1BnO3n+1)2−. From the side view of the structure, the layers of Bi atoms are visible, in between of which the A and B site atoms are present (Figure 1). There is a total of 4 A-site atoms which can be divalent, like Ca, Ba, Be, Mg, Sr, etc., and 8 B-site atoms which can be Nb and Bi atoms, and there is a total of 36 atoms of oxygen in this structure, with a total of 56 atom possessing the orthorhombic structure, with a = 5.64, b = 25.40, c = 5.68, and α = β = γ = 90°.

3.2. Electronic Properties

3.2.1. Band Structure and Density of State

The density of state (DOS) is used to acquire significant information regarding a material’s electronic movement within a certain orbital band, i.e., between the VB and conduction band (CB) orbits. The band structure and DOS determine the electronic orbital movement between energy levels for XNb2Bi2O9, which is displayed in Figure 2 and Figure 3 along the high-symmetry directions of the irreducible brillion zone (IBZ). These are obtained by employing the GGA approximation within the energy range of −6.0 eV to 6.0 eV. The dashed line at the center of Figure 2 presents the Fermi level (EF) at 0 eV. The VB is placed below EF, whereas the CB lies above EF. Sharp peaks are visible around the Fermi level on the VB’s edge. The band edges for VB and CB lie at the Y and G points, respectively. The results from the bandgap and TDOS clearly show the direct bandgap nature of these compounds. Direct bandgap semiconductors are advantageous for optical and electronic applications, as phonons are included in electronic transitions with electrons, as in indirect semiconductors. The values of the energy bandgaps for MgNb2Bi2O9, BaNb2Bi2O9, SrNb2Bi2O9, CaNb2Bi2O9, and BeNb2Bi2O9 are 2.501, 2.493, 2.493, 2.490, and 1.834 eV, respectively. Overall, all the compounds follow the same bandgap trend, which lies within the range of the semiconductor bandgap value. The bandgap investigation was performed using the CASTEP code governed by a set of Equations (1)–(3).
  ψ k i r = exp i k r f i r
  f i r = G G i , G exp i G . r
  G h 2 2 m k + G 2 δ G G + V i o n G G + V H G G C i , k + G = ε i C i , k + G
where, in Equations (1)–(3), ψ represents the pseudo wave function, while the wave vector is denoted by k; R represents the displacement function; δGG represents the Kronecker delta function; the unique constant density function is represented by G; Vion represents the ionic potential, while Vxc is the exchange correlation potential; and r is the radial distance in a spherical coordinate system for a point in space.

3.2.2. Partial Density of States (PDOS)

The PDOS of the compounds is determined by using BIOVIA and is presented in Figure 4 to illustrate the contribution of each atomic orbital in creating the band structure and the orbital contributions in adjusting the bandgap.
(a)
CaNb2Bi2O9
Figure 4a shows that the VB is mostly influenced by the O and Ca atoms, with a little minute influence by the Nb atoms. To be more precise, the O-2p4, Ca-4d4, and Nb-3d10 states have a significant influence in this sub-band, while the Bi-5p6 state is also somewhat present. Bi and Nb have a significant influence on the VB’s second sub-band, whereas Ca also has a slight presence. After that, there is no electronic state between 0 and 2.9 eV, which is the material’s energy bandgap. The Ca and O atoms contribute very little to the CB of CaNb2Bi2O9, which is mostly made up of Nb and Bi atoms. More precisely, the Bi-6s2 and Nb-3d10 states demonstrate their contributions, while the Ca-4d4 and Bi-5p6 states also show little presence.
(b)
BaNb2Bi2O9
From Figure 4b, it can be seen that VB mainly presents an outcome of the hybridization of the O-2p4, Ba-5p6, Ba-4d10, and Nb-4s2 states, but Nb-4p6 shows minute presence. The Ba, Nb, and O atoms have a significant influence on the VB. The compound’s energy gap is found to lie between 0 and 2.8 eV, where no electronic state of any type exists. Nb and Bi have an influence on the CB of BaNb2Bi2O9. Specifically, the states Bi-6s2 and Nb-3d10 exert significant effects, whereas the Ba-4d10 state also exhibits minimal presence.
(c)
BeNb2Bi2O9
Figure 4c shows that VB is significantly influenced by the Nb and O atoms. This sub-band is mostly composed of the O-2p4 state, with a minor proportion of Nb-4s2 orbitals, according to PDOS. The material’s energy bandgap lies between 0 and 2.0 eV, where no more electronic states exist. The atoms that influence the CB of BeNb2Bi2O9 are Bi, Be, and Nb. A tiny percentage of the Be-2s2 and Bi-5p6 states hybridize with the Bi-6s2, Nb-3d10, and Be-2p4 states to affect CB, according to the PDOS.
(d)
MgNb2Bi2O9
Figure 4d shows that the O and Nb atoms have a significant influence on the VB. Based on the PDOS study, this sub-band consists primarily of O-2p4 and Nb-4s2 states, with a small proportion of Bi-6s2 orbitals. After that, electronic states do not exist between 0 and 2.4 eV, which is the compound’s energy bandgap. The Nb, Bi, and Mg atoms are the main contributors to the CB. As to the PDOS, the formation of CB arises from the hybridization of Mg-3s2, Bi-6s2, and Nb-3d10 states, along with a small proportion of Bi-5p6 states.
(e)
SrNb2Bi2O9
Figure 4e shows that Sr, O, and Nb atoms demonstrate a significant influence on VB. According to the PDOS study, Sr-3d10, O-2p4, and Nb-3d10 states contribute the majority of this sub-band’s formation. After that, no electronic states appear between 0 and 2.4 eV, which is the compound’s energy bandgap. The CB is mostly attributed to the Sr, Bi, and Nb atoms. The PDOS result indicates that a minor amount of the Bi-5p6 state combines with the Sr-3d10, Bi-6s2, and Nb-3d10 states to generate CB.

3.3. Optical Properties

Significant and focused understanding about the refractive index n(ω) is required to assess the technical applications of optical materials in optical devices. Materials used in photovoltaic systems must have strong optical conductivity, high absorption coefficient, low emissivity, and a high refractive index. Using the relations provided for understanding the depth of the optical characteristics, the refractive index n(ω), reflectivity R(ω), absorption coefficient I(ω), and energy loss function L(ω) were computed by using BIOVIA Material Studio. Equations (4)–(9) are the expressions of these optical and electronic measurements:
n ω = ε 1 w 2 + ε 1 2 ω + ε 2 2 ω 2 1 2  
      ε 1 ω = 1 + 2 π P 0 ω ε 2 ω ω 2 ω 2 d ω
ε 2 ω = 4 π 2 e 2 m 2 ω 2 V B z ψ k v p i ψ k C 2 δ E ψ k c E ψ k v ħ ω
R ω = ε ω 1 ε ω + 1 2
I ω = ε 1 2 ω + ε 2 2 ω ε 1 w 1 2
L ω = Im 1 ε = ε 2 ω ε 1 2 ω + ε 2 2 ω
The expression n   ( ω ) = n   ( ω ) + i k   ( ω ) describes the complex refractive index. This case indicates that the degree of transparency of a material is determined by the incoming electromagnetic radiations. Figure 5a shows the refractive index plots of XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr, which have demonstrated static refractive indices (n) of 3.29, 3.31, 3.24, 3.33, and 3.29, respectively. The values of the refractive indices increase with the energy of the photon, and for XNb2Bi2O9, the largest peak is seen at 3.54 eV. The materials will be regarded as optically active if the value of n(ω) lies between 1.0 and 4.0 eV.
The threshold values of energy for the calculated spectra of K(ω) are presented in Figure 5b. The spectra of K(ω) initially show no peak, which is the known specific spectra for the threshold energy, and the values of this energy are approximately 2.48, 2.52, 2.57, 2.49, and 2.49 eV for XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr, respectively. The main cause of the prominent peaks in the spectra of K(ω) is the transition of electrons from VB to CB.
Equations (10) and (11) provide the intensity of light I(x) expression at the distance x from the surface of the material, because the absorption coefficient (α) indicates the part of the energy of light that is lost by the light wave as it passes through the material [30].
I x = I 0 exp α x
w h e r e   α = 2 k ω c
The penetration length of the incoming photon before its total absorption in the material is computed by using the absorption coefficient. The calculated spectra of α(ω) are shown in Figure 5c. All optical properties are measured for the energy range of 0–14 eV. Absorption is lowest for the material SrNb2Bi2O9, with a value of 199,031 cm−1, and greatest for CaNb2Bi2O9, with a value of 218,645 cm−1. The α(ω) spectrum demonstrates the highest peaks between 9.0 and 14.0 eV. These materials often behave as transparent in low-energy areas or in the infrared region; they may find suitable application in UV devices. The results show that these materials have the potential to be employed in solar devices.
The simulated reflectivity R(ω) spectra for XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr in the energy range of 0–14 eV are presented in Figure 5d. The reflectivity R(ω) is the measure of the fraction of incident photons that are reflected at the interface. The static values of R(ω) that correspond to the zero-frequency limit for XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr are 0.17, 0.187, 0.19, 0.20, and 0.18, respectively. Prominent peaks in the reflectivity spectra in the higher UV (2–6 eV) region indicate excited absorption. These materials can reflect more than 50% of incident light in the high UV range.
Figure 6a,b shows the Kramers–Kronig relation for the complex dielectric function ε ( ω ) = ε 1 ( ω ) + i ε 2 ( ω ) , which can be used to describe the dielectric properties. Here, ε1(ω) represents the real part that is useful for measuring light polarization and dispersion, and ε2(ω) denotes the imaginary part that characterizes the material’s absorptive qualities [34]. The maximum values of ε1(ω) for XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr occur at 3.4, 3.35, 3.16, 3.39, and 3.38 eV, respectively. Following that, one may see a decline in the spectrum with rising energy. Direct optical transitions take place between VBs and CBs because the electronic band topologies show the direct bandgap. For these compounds, the static dielectric constants (a real component of the dielectric constant at zero energy) are 6.35 for (ε1). These outcomes show how these materials develop an anisotropic dielectric function. The orthorhombic structure is responsible for the dielectric constants’ anisotropic nature. The total of all transitions from VBs to CBs is known as ε2(ω). At an energy of 4.5 eV, the XNb2Bi2O9 (X = Ca, Ba, Be, Mg, and Sr) structure exhibits a maximum value in ε2(ω), occurring inside the visible light spectrum.
The shift in optical conductivity σ(ω) with photon energy is presented in Figure 6c. Initially, the compounds act as semiconductors, since the conductivity σ(ω) values are zero. As seen in Figure 6c, the critical energy point of ~1.8 eV is where the conductivity σ(ω) begins. With a comparable value of 5.3 fs−1, the conductivity σ(ω) increases with an increase in photon energy up to ~4.8 eV in the visible range. The high absorption of the material is the cause of the increase in optical conductivity.
The electron energy loss (EEL) spectra of XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr in the energy range of 0–14 eV are presented in Figure 6d. When attempting to distinguish between plasma resonance occurrences and regular inter-band transitions, the EEL approach can be helpful. Sharp peaks in the energy loss function are known to coincide with the roots of Re(ε) crossing the x = 0 line, which is related to plasma oscillations. The EEL spectra show only one major peak that is connected to plasma oscillations. The plasmon energy has the greatest peak value at 6.7 eV. It is evident that the Re(ε) passes through 0 at these energies.

4. Conclusions

For the first time, a detailed theoretical investigation has been conducted on the layered perovskites XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr. The objective of this study is to analyze the optical and electronic characteristics of these materials to clarify their potential for optical and electronic applications. Since the observed bandgap lies between 1.8 and 2.6 eV for all compounds, these materials reveal a direct energy bandgap and are ideally suited for utilization in solar energy conversion applications. The optical parameters, such as the real part of the dielectric function, range from 10 to 10.6, which shows that these materials absorb the maximum number of incident photons in the visible region. The refractive index varies from 3.24 to 3.33, which proves that they are optically active. The optical conductivity, threshold energy of the extinction coefficient, absorption coefficient, and static value of reflectivity lie between 5.04 fs−1 and 5.3 fs−1, 2.4 and 2.6 eV, 192,000 cm−1 and 219,000 cm−1, and 0.33 and 0.35, respectively. Conversely, a low-range loss function of 0.351 to 0.545 has been reported. The above-discussed parameters show that XNb2Bi2O9 is a potential candidate for photovoltaic applications.

Author Contributions

Conceptualization, N.J. and A.H.; methodology, F.K. and A.Y.; software, A.Y. and N.U.H.; validation, A.H., F.K. and A.Y.; formal analysis, N.U.H.; investigation, A.Y.; resources, M.E.K.; data curation, F.K.; writing—original draft preparation, N.J. and A.H.; writing—review and editing, N.J. and M.E.K.; visualization, A.Y.; supervision, N.J.; project administration, A.H.; funding acquisition, M.E.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Taif University, Saudi Arabia, Project No. (TU-DSPP-2024-252). Also, authors are thankful to the University of Lahore, Pakistan for the ORIC-SRGP 17/2024 research fund.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors extend their appreciation to Taif University, Saudi Arabia, for supporting this work through project number (TU-DSPP-2024-252). Also, they are thankful to University of Lahore, Pakistan for ORIC-SRGP 17/2024 research fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kehoe, A.B.; Scanlon, D.O.; Watson, G.W. Modelling potential photovoltaic absorbers Cu3MCh4 (M = V, Nb, Ta; Ch = S, Se, Te) using density functional theory. J. Phys. Condens. Matter 2016, 28, 175801. [Google Scholar] [CrossRef]
  2. Andreotti, A.; Del Pizzo, A.; Rizzo, R.; Tricoli, P. An efficient architecture of a PV plant for ancillary service supplying. In Proceedings of the SPEEDAM 2010, Pisa, Italy, 14–16 June 2010; pp. 678–682. [Google Scholar]
  3. Abubakr, M.; Fatima, K.; Abbas, Z.; Hussain, A.; Jabeen, N.; Raza, H.H.; Chaib, Y.; Muhammad, S.; Siddeeg, S.M.; Gorczyca, I. Effect of S, Se and Te replacement on structural, optoelectronic and transport properties of SrXO4 (X = S, Se, Te) for energy applications: A first principles study. J. Solid State Chem. 2022, 305, 122689. [Google Scholar] [CrossRef]
  4. Piegari, L.; Rizzo, R.; Tricoli, P. Sizing criteria for stand alone power plants supplied by photovoltaic arrays. ACTA Electroteh. 2008, 2008, 241–246. [Google Scholar]
  5. Glaser, P.E.; Maynard, O.E.; Mackovciak, J.J.R.; Ralph, E.I. Feasibility Study of a Satellite Solar Power Station; (No. ADL-C-74830); NASA: Washington, DC, USA, 1974. [Google Scholar]
  6. Herold, D.; Horstmann, V.; Neskakis, A.; Plettner-Marliani, J.; Piernavieja, G.; Calero, R. Small scale photovoltaic desalination for rural water supply-demonstration plant in Gran Canaria. Renew. Energy 1998, 14, 293–298. [Google Scholar] [CrossRef]
  7. Liu, M.; Johnston, M.B.; Snaith, H.J. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature 2013, 501, 395–398. [Google Scholar] [CrossRef] [PubMed]
  8. Carlson, D.E.; Wronski, C.R. Amorphous silicon solar cell. Appl. Phys. Lett. 1976, 28, 671–673. [Google Scholar] [CrossRef]
  9. McDonald, S.A.; Konstantatos, G.; Zhang, S.; Cyr, P.W.; Klem, E.J.; Levina, L.; Sargent, E.H. Solution-processed PbS quantum dot infrared photodetectors and photovoltaics. Nat. Mater. 2005, 4, 138–142. [Google Scholar] [CrossRef] [PubMed]
  10. Burschka, J.; Pellet, N.; Moon, S.J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316–319. [Google Scholar] [CrossRef]
  11. Panthani, M.G.; Akhavan, V.; Goodfellow, B.; Schmidtke, J.P.; Dunn, L.; Dodabalapur, A.; Barbara, P.F.; Korgel, B.A. Synthesis of CuInS2, CuInSe2, and Cu(InxGa1−x)Se2 (CIGS) nanocrystal “inks” for printable photovoltaics. J. Am. Chem. Soc. 2008, 130, 16770–16777. [Google Scholar] [CrossRef]
  12. Boeer, K.W. p-type emitters covered with a thin CdS layer show a substantial improvement of Voc and FF. Sol. Energy Mater. Sol. Cells 2011, 95, 786–790. [Google Scholar] [CrossRef]
  13. Tang, C.W. Two-layer organic photovoltaic cell. Appl. Phys. Lett. 1986, 48, 183–185. [Google Scholar] [CrossRef]
  14. Sun, S.; Wang, W.; Xu, H.; Zhou, L.; Shang, M.; Zhang, L. Bi5FeTi3O15 hierarchical microflowers: Hydrothermal synthesis, growth mechanism, and associated visible-light-driven photocatalysis. J. Phys. Chem. C 2008, 112, 17835–17843. [Google Scholar] [CrossRef]
  15. Tang, J.; Zou, Z.; Ye, J. Efficient photocatalytic decomposition of organic contaminants over CaBi2O4 under visible-light irradiation. Angew. Chem. Int. Ed. 2004, 43, 4463–4466. [Google Scholar] [CrossRef] [PubMed]
  16. Naresh, G.; Malik, J.; Meena, V.; Mandal, T.K. pH-mediated collective and selective solar photocatalysis by a series of layered Aurivillius perovskites. ACS Omega 2018, 3, 11104–11116. [Google Scholar] [CrossRef] [PubMed]
  17. Kudo, A.; Hijii, S. H2 or O2 evolution from aqueous solutions on layered oxide photocatalysts consisting of Bi3+ with 6s2 configuration and d0 transition metal ions. Chem. Lett. 1999, 28, 1103–1104. [Google Scholar] [CrossRef]
  18. Shimodaira, Y.; Kato, H.; Kobayashi, H.; Kudo, A. Photophysical properties and photocatalytic activities of bismuth molybdates under visible light irradiation. J. Phys. Chem. B 2006, 110, 17790–17797. [Google Scholar] [CrossRef] [PubMed]
  19. Zhou, L.; Yu, M.; Yang, J.; Wang, Y.; Yu, C. Nanosheet-based Bi2MoxW1−xO6 solid solutions with adjustable band gaps and enhanced visible-light-driven photocatalytic activities. J. Phys. Chem. C 2010, 114, 18812–18818. [Google Scholar] [CrossRef]
  20. Cui, R.; Zhao, X.; Ye, Y.; Deng, C. Layer-effects on electrical and photovoltaic properties of Aurivillius-type Srn-3Bi4TinO3n+3 (n = 3, 5) films grown by pulsed laser deposition. J. Mater. Sci. Mater. Electron. 2022, 33, 25318–25328. [Google Scholar] [CrossRef]
  21. Lardhi, S.; Noureldine, D.; Harb, M.; Ziani, A.; Cavallo, L.; Takanabe, K. Determination of the electronic, dielectric, and optical properties of sillenite Bi12TiO20 and perovskite-like Bi4Ti3O12 materials from hybrid first-principle calculations. J. Chem. Phys. 2016, 144, 134702. [Google Scholar] [CrossRef]
  22. Werner, V.; Aschauer, U.; Redhammer, G.J.; Schoiber, J.; Zickler, G.A.; Pokrant, S. Synthesis and Structure of the Double-Layered Sillén–Aurivillius Perovskite Oxychloride La2.1Bi2.9Ti2O11Cl as a Potential Photocatalyst for Stable Visible Light Solar Water Splitting. Inorg. Chem. 2023, 62, 6649–6660. [Google Scholar] [CrossRef]
  23. Liu, S.M.; Zhong, H.X.; Liang, J.J.; Zhang, M.; Zhu, Y.H.; Du, J.; Guo, W.H.; He, Y.; Wang, X.; Shi, J.J. Lead-free layered Aurivillius-type Sn-based halide perovskite Ba2X2[Csn−1SnnX3n+1] (X = I/Br/Cl) with an optimal band gap of ∼1.26 eV and theoretical efficiency beyond 27% for photovoltaics. J. Mater. Chem. A 2022, 10, 10682–10691. [Google Scholar] [CrossRef]
  24. Xu, B.; Li, X.; Sun, J.; Yi, L. Electronic structure, ferroelectricity and optical properties of CaBi2Ta2O9. Eur. Phys. J. B 2008, 66, 483–487. [Google Scholar] [CrossRef]
  25. Xu, B.; Yi, L. First-principle study of the ferroelectricity and optical properties of the BaBi2Ta2O9. J. Alloys Compd. 2007, 438, 25–29. [Google Scholar] [CrossRef]
  26. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  27. Pela, R.R.; Hsiao, C.L.; Hultman, L.; Birch, J.; Gueorguiev, G.K. Electronic and optical properties of core–shell In AlN nanorods: A comparative study via LDA, LDA-1/2, mBJ, HSE06, G 0 W 0 and BSE methods. Phys. Chem. Chem. Phys. 2024, 26, 7504–7514. [Google Scholar] [CrossRef] [PubMed]
  28. Dos Santos, R.B.; Rivelino, R.; de Brito Mota, F.; Kakanakova-Georgieva, A.; Gueorguiev, G.K. Feasibility of novel (H3C)nX(SiH3)3−n compounds (X = B, Al, Ga, In): Structure, stability, reactivity, and Raman characterization from ab initio calculations. Dalton Trans. 2015, 44, 3356–3366. [Google Scholar] [CrossRef] [PubMed]
  29. Segall, M.D.; Lindan, P.J.; Probert, M.A.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717. [Google Scholar] [CrossRef]
  30. Haas, P.; Tran, F.; Blaha, P.; Schwarz, K.; Laskowski, R. Insight into the performance of GGA functionals for solid-state calculations. Phys. Rev. B—Condens. Matter Mater. Phys. 2009, 80, 195109. [Google Scholar] [CrossRef]
  31. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Für Krist-Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  32. Ziesche, P.; Kurth, S.; Perdew, J.P. Density functionals from LDA to GGA. Comput. Mater. Sci. 1998, 11, 122–127. [Google Scholar] [CrossRef]
  33. Mattsson, A.E.; Armiento, R.; Schultz, P.A.; Mattsson, T.R. Nonequivalence of the generalized gradient approximations PBE and PW91. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 73, 195123. [Google Scholar] [CrossRef]
  34. Hilal, M.; Rashid, B.; Khan, S.H.; Khan, A. Investigation of electro-optical properties of InSb under the influence of spin-orbit interaction at room temperature. Mater. Chem. Phys. 2016, 184, 41–48. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of bismuth layered structure ferroelectrics (BLSFs) with generalized formula (Bi2O2)2+(An−1BnO3n+1)2–.
Figure 1. Crystal structure of bismuth layered structure ferroelectrics (BLSFs) with generalized formula (Bi2O2)2+(An−1BnO3n+1)2–.
Crystals 14 00710 g001
Figure 2. Band structure of BLSFs family: (a) CaNb2Bi2O9; (b) BaNb2Bi2O9; (c) BeNb2Bi2O9; (d) MgNb2Bi2O9; and (e) SrNb2Bi2O9.
Figure 2. Band structure of BLSFs family: (a) CaNb2Bi2O9; (b) BaNb2Bi2O9; (c) BeNb2Bi2O9; (d) MgNb2Bi2O9; and (e) SrNb2Bi2O9.
Crystals 14 00710 g002
Figure 3. Density of states of (a) CaNb2Bi2O9, (b) BaNb2Bi2O9, (c) BeNb2Bi2O9, (d) MgNb2Bi2O9, and (e) SrNb2Bi2O9.
Figure 3. Density of states of (a) CaNb2Bi2O9, (b) BaNb2Bi2O9, (c) BeNb2Bi2O9, (d) MgNb2Bi2O9, and (e) SrNb2Bi2O9.
Crystals 14 00710 g003
Figure 4. Partial density of states (PDOS) of (a) CaNb2Bi2O9, (b) BaNb2Bi2O9, (c) BeNb2Bi2O9, (d) MgNb2Bi2O9, and (e) SrNb2Bi2O9.
Figure 4. Partial density of states (PDOS) of (a) CaNb2Bi2O9, (b) BaNb2Bi2O9, (c) BeNb2Bi2O9, (d) MgNb2Bi2O9, and (e) SrNb2Bi2O9.
Crystals 14 00710 g004
Figure 5. Optical properties of XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr, (a) refractive index, (b) extinction coefficient, (c) absorption (cm−1), and (d) reflectivity.
Figure 5. Optical properties of XNb2Bi2O9 with X = Ca, Ba, Be, Mg, and Sr, (a) refractive index, (b) extinction coefficient, (c) absorption (cm−1), and (d) reflectivity.
Crystals 14 00710 g005
Figure 6. (a) The real part of the dielectric function, (b) imaginary part of the dielectric function, (c) optical conductivity, and (d) loss function of CaNb2Bi2O9, BaNb2Bi2O9, BeNb2Bi2O9, MgNb2Bi2O9, and SrNb2Bi2O9.
Figure 6. (a) The real part of the dielectric function, (b) imaginary part of the dielectric function, (c) optical conductivity, and (d) loss function of CaNb2Bi2O9, BaNb2Bi2O9, BeNb2Bi2O9, MgNb2Bi2O9, and SrNb2Bi2O9.
Crystals 14 00710 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kainat, F.; Jabeen, N.; Yaqoob, A.; Hassan, N.U.; Hussain, A.; Khalifa, M.E. Effect of Ca, Ba, Be, Mg, and Sr Substitution on Electronic and Optical Properties of XNb2Bi2O9 for Energy Conversion Application Using Generalized Gradient Approximation–Perdew–Burke–Ernzerhof. Crystals 2024, 14, 710. https://doi.org/10.3390/cryst14080710

AMA Style

Kainat F, Jabeen N, Yaqoob A, Hassan NU, Hussain A, Khalifa ME. Effect of Ca, Ba, Be, Mg, and Sr Substitution on Electronic and Optical Properties of XNb2Bi2O9 for Energy Conversion Application Using Generalized Gradient Approximation–Perdew–Burke–Ernzerhof. Crystals. 2024; 14(8):710. https://doi.org/10.3390/cryst14080710

Chicago/Turabian Style

Kainat, Fatima, Nawishta Jabeen, Ali Yaqoob, Najam Ul Hassan, Ahmad Hussain, and Mohamed E. Khalifa. 2024. "Effect of Ca, Ba, Be, Mg, and Sr Substitution on Electronic and Optical Properties of XNb2Bi2O9 for Energy Conversion Application Using Generalized Gradient Approximation–Perdew–Burke–Ernzerhof" Crystals 14, no. 8: 710. https://doi.org/10.3390/cryst14080710

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

Article Metrics

Back to TopTop