Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Structural, electronic and optical characteristics of inorganic cubic perovskite Sr3AsI3

Open Access Open Access

Abstract

Inorganic perovskite materials have recently received significant consideration in the sector of solar technology because of their tremendous structural, optical, and electronic strengths. This research exhaustively inquired about the structural, optical, and electronic characteristics of the inorganic cubic perovskite Sr3AsI3 utilizing the first-principles density-functional theory (FP-DFT). The Sr3AsI3 molecule exhibits a direct bandgap of 1.265 eV value at Γ point. According to band characteristics, this component has a strong absorption capability in the region of visibility, as demonstrated by optical parameters including dielectric functions, absorption coefficient, reflectivity, and electron loss function. It is discovered that the spikes of the dielectric constant of Sr3AsI3 are visible in the photon energy range which are suitable for solar cells. As a result, the Sr3AsI3 perovskite is considered suitable for the application of energy production and light management in solar cells.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Organic-inorganic perovskites (OILHP) have recently gained a great deal of experimental interest in solar technology because of their fascinating qualities, such as a reasonable bandgap, widespread availability, amazing optical absorption capacity, low reflectivity and low manufacturing cost [13]. Despite the fact that CdTe, CIGS, Sb2Se3, CMTS, and FeSi2 solar cell materials have made significant strides in the PV field [49], researchers have begun to use perovskites materials due to their exceptional qualities [1013]. Perovskite materials can capture a greater number of photons while utilizing a layer with a thickness lower than one micrometer compared to other semiconductor materials [14]. The solar cells using perovskite-like CsPbI3 can attain a power conversion efficiency (PCE) of 17.9% [15] without hole transport layer (HTL) in 2022 and 19.06% [16] with HTL in 2023. OILHP-related solar cells can demonstrate high efficiency of power conversion (PCE), which, in 2022, amounted to 25.8% [17]. The enhancement of the perovskite materials’ characteristics during the previous ten years has accelerated this evolution. Inorganic perovskites made of metal halides have received a lot of intentness. Unfortunately, the OILHP’s long-term durability issue is currently a major worry. Since these perovskites are highly susceptible to moisture, wind, sunlight and temperature, when utilized in a real environment, their protracted stability will be considerably impacted [18,19]. Consequently, it has become particularly difficult to solve the instability issue with these perovskites in recent times.

It is expected that the OILHP solar cells’ difficulties with heat instability and optical instability can be significantly improved by switching out the organic cation for an inorganic cation [20,21]. A few inorganic halide perovskites currently outperform and stabilize OILHP in terms of performance. Zhu et al. [22] demonstrated that inorganic halide perovskites exhibit pretty much identical band edge carrier characteristics. Inorganic cubic perovskites have lately been thought to be direct bandgap substances with a large capacity for optical absorption, making them a prime contender for LED, semiconducting, and solar technologies [2325]. Hence, the drawbacks of OILHP are anticipated to be resolved with the successful implication and manufacture of inorganic halide perovskite components along with their photovoltaic cells. Recently, the A3BX3 type perovskite materials such as Sr3AsI3 perovskite has attracted a lot of attention in the field of solar technology due to their exceptional structural, optical, and electronic characteristics. We will study about it to calculate how effective it will be for solar cells and optoelectronics. However, there is still a lack of study about Sr3AsI3 perovskite. Therefore, the comprehensive study of Sr3AsI3 perovskite is very important. Since the bandgap is liable for both the production of particles and light absorption, it has a substantial impact on the PCE of solar cells. According to Shockley-Queisser theory, perovskite solar cells’ PCE might achieve up to 33% if their molecule's bandgap is set to be between 1.2-1.4 eV [26,27]. Despite being ideal materials for optoelectronic and photovoltaic systems, inorganic halide perovskites have a noticeable drawback: a marginally greater bandgap [28,29]. This makes the electronic bandgap controllability with undertaking approaches extremely important for getting the maximum PCE in inorganic halide perovskite solar cells. More nucleons are present in larger cations than in smaller ones due to their larger atomic sizes. The electronic band structure subsequently changes as a result of the fluctuation in atomic size

In this paper, the optical and electronic characteristics of Sr3AsI3 are investigated thoroughly using the density functional theory (DFT) methodology. We carefully examined the band structure and bandgap customizing procedure of Sr3AsI3. We also calculated the dielectric function, absorption coefficient, reflectivity and loss function of Sr3AsI3. By adjusting the optoelectronic characteristics of the Sr3AsI3 substance, one can customize it appropriately for optoelectronic as well as photovoltaic technologies.

2. Details of computational method

Sr3AsI3 perovskite structure was processed to the FP-DFT with norm-conserving (NC) pseudopotential [3032] and the Perdew-Burke-Ernzerhof (PBE) [33] exchange-correlation mechanism. The DFT was anticipated to be generated by the Quantum Espresso simulation program [3437]. The input data contained the necessary preliminary settings, including the Brillouin zone grid, crystal formations, lattice parameters, and kinetic cut-off energy. In order to optimize the structure and boost performance, the kinetic energy cut-off and charge density cut-off were adjusted to 30 Rydberg (Ry) and 220 Rydberg (Ry). The dimension of the k-point (kx,ky,kz) was set to (6 × 6x6) for the optimization of lattice utilizing the vc-relax computation. The self-consistency equation was calculated using ∼10−6 a.u. value of the convergence threshold with the maximum set of force tolerance at < 0.01 eV/Å [38]. The force convergence threshold of almost 10−3 a.u was taken into account throughout structural and ionic adjustment relaxation studies. We have not utilized adjusted PBE for metals in the latest research, despite the fact that there are certain methods for reducing this inaccuracy [39,40]. After achieving dynamical stabilization, the perovskite framework’s optical characteristics were inspected by the computation of their complicated dielectric functions that are photon energy sensitive. Utilizing the QE package's theory of time-dependent first-order perturbation, the optical characteristics were computed [41]. Therefore, the complicated dielectric component was examined to determine the energy spectrum of the photon (eV) at which it exhibits absorption peaks. While measuring the optical absorption coefficients, the complex dielectric function, ɛ(ω)= ɛ1(ω)+ jɛ2(ω) is regarded as the fundamental relationship.

3. Result and discussion

3.1. Structural properties

The Pm3m cubic foundation has been identified as the periodic pattern of Sr3AsI3 [42]. The structure's unit cell is made up of seven atoms. This substance contains an octahedral gap filled with I atoms and an arrangement of Sr and As atoms in a face-centered cubic lattice. The bond lengths of both Sr–As are 2.84 Å. The average Sr-l bond length is 2.84 Å. Six comparable Sr2+ atoms are linked to As3- to create corner-sharing AsSr6 octahedra. The octahedra that share a corner are not slanted. As, I, and Sr's fractional coordinates are (0,0,0), (0.5,0.5,0), and (0,0.5,0), respectively, utilizing the 1a, 3c, and 3d Wyckoff sites as shown in Fig. 1(a). Figure 1(b) depicts the first Brillouin zone's (BZ) k-path. Before computing the different properties of Sr3AsI3 perovskite, the computation of structural properties is necessary. Applying PBE, the structural characteristics were derived, such as the value of lattice constant a(Å) which is illustrated in Table 1. The most sustainable lattice constant for Sr3AsI3 has been discovered by evaluating the total amount of energy with consideration to the lattice parameter. The lattice constant for the Sr3AsI3 compound is estimated to be a = 6.58 Å. Moreover, a certain structure's conjunctive and production energy can be used as efficient instruments for confirming the durability of such a structure [4345].

 figure: Fig. 1.

Fig. 1. (a) The inorganic perovskite Sr3AsI3's optimum structure. (b) The first Brillouin zone's k-path in order to determine their electronic band structure.

Download Full Size | PDF

Tables Icon

Table 1. Sr3AsI3's lattice constant and energy bandgap were determined using experimental data and previous DFT calculations.

3.2. Charge density

An essential part of the analysis of a component's electronic characteristics is the study of electronic charge density. This characteristic produces a charge density structural map of the valence electrons concerning the overall concentration of charge in a structure's unit cell. The total electronic charge density map is examined to investigate the nature of chemical bonding inside a molecule. Generally, the charge density curve is made up of structural atoms that show how orbital electrons contribute to the electrical properties of atoms by accumulating charges. Then, the differentiated color density map introduces the charge contributions from the individual elements’ electronic DOS spectrum to correlate them. Figures 2(a) and 2(b) show the mapping image of the electron density in 2D view and bird's eye view. The colors red and blue, respectively, are used to convey high and low intensity. It can be observed, for all planes, the Sr element is the region where charge accumulates the most, while its depletion occurs close to the As atom. In another word, the intersection of outer electrons between these two components (Sr and As) suggests a covalent bond [47,48].

 figure: Fig. 2.

Fig. 2. The charge density of Sr3AsI3 structure (a) 2D view, (b) Bird’s eye view.

Download Full Size | PDF

The covalent bonding characteristic of the Sr-As atoms is largely supported by this charge distribution route. According to the researchers, the charge density around the atoms is almost spherical, which is an indication of ionic bonding that may be likened to perovskites that have already been published [49,50]. Here, the bonding between Sr and I atoms is an indication of an ionic bond. On the other side, the As-I bond has a negative population value, making it an antibonding feature.

3.3. Electronic properties

The electronic characteristics of a substance are primarily determined by its band structure, charge density, and density of states (DOS) [51]. The properties of the electronic band for Sr3AsI3 perovskite structures have been estimated, as well as the high equipoise directions. The Sr3AsI3 framework's electronic band configurations are depicted in Fig. 3(a). A zero adjustment has been made to the fermi levels in order to easily analyze the value of the bandgap. The cubic structure's Γ-X-M-R-Γ is considered along with the k-axis. The conduction band minimum (CBM) and valence band maximum (VBM), which are both located near the Γ (Gamma) point, are shown in Fig. 3(a). According to calculations made using the PBE/HSE function for Sr3AsI3, the Sr3AsI3 perovskite is predicted to have direct bandgap structures with values of around 1.265 eV/1.94 eV. This outcome is fairly consistent with the values that were previously published [52,53]. The bandgap value was clearly undervalued when the bandgap was evaluated utilizing the GGA approach, which is a very typical flaw in the GGA technique. In the (LDA)+U and LDA methods of approximation for local density, bandgap undervaluation was also discovered [54]. Several experts have provided a variety of techniques, including the GW methodology hybrid functional to avoid this kind of bandgap computation [55,56].

 figure: Fig. 3.

Fig. 3. The electronic (a) band structure and (b) refined structure of PDOS for inorganic Sr3AsI3 perovskite with the function of PBE/HSE.

Download Full Size | PDF

Generally, the partial density of states (PDOS) shows how the bandgap energy of the Sr3AsI3 structure is affected by individual atoms and their various forms. The dispersion of PDOS in Sr3AsI3 for the region of −3 to 3 eV is illustrated in Fig. 3(b). The forms of Sr and As hybridized with I in Sr3AsI3, are seen to expand across the full energy range while conserving the bandgap. It demonstrates that the fundamental kind of bonding between Sr-I and As-I is covalent. Sr3AsI3 additionally involved the transmission of electron charge from Sr and As to I (Fig. 3(b)), which could be explained by the sharp disparity between the atomic states. Near the vicinity of the Fermi level, the contribution of Sr atoms is essentially nonexistent. The I-2p orbital dominates our cubic phase article's analysis of the density of states near Sr3AsI3's valence band, where the As-2p orbital and a minor contribution from the Sr-3s orbitals both play a significant role in the conduction band.

3.4. Optical properties

An assessment of optical qualities includes a study of the complicated dielectric functions, electron loss function, absorption coefficient, and reflectivity to determine whether the materials under evaluation are suitable for optoelectronics and photovoltaic cell technologies. A thermodynamic method called biaxial strain can boost a substance's optical performance by modulating the lattice parameter [57]. In this research, various optical characteristics of Sr3AsI3 has been investigated. The dielectric function is denoted by the symbol ɛ(ω). It is calculated by the summation of two parts, one part is real which is denoted by ɛ1(ω) and another part is imaginary which is denoted by ɛ2(ω).

$$\varepsilon (\omega )= {\varepsilon _1}(\omega )\; + i{\varepsilon _2}(\omega )\; $$

The transformation of Kramers- Kronig [58] is implemented to obtain the real dielectric function and the components of the momentum matrix [59] are used to determine the imaginary part. The real part of Sr3AsI3's dielectric permittivity is depicted in Fig. 4(a) for photon energies up to 10 eV. The real component of the dielectric constant can be utilized to learn more about the polarization and dispersion impacts. The maximum frequency of zero, abbreviated as ɛ1(0), which refers to the electronic component of the real state of the dielectric function, is the finest fundamental factor in the part of the real state ɛ1(ω). Cubic Sr3AsI3's determined ɛ1(0) value comes out to 5.95 (Fig. 4(a)). When exposed to optical excitation, the quantity of ɛ1(ω) began to increase between ɛ1(0) and the maximum amount of ɛ1(ω), after which it abruptly decreased, indicating that the material has a significant capability for light absorption in this spectral region. Furthermore, the Sr3AsI3 perovskite has positive ɛ1(ω) values, indicating that it is highly refractive and semiconducting. In general, higher bandgap components display a smaller peak value of dielectric constant compared to the low band gap materials [38]. As a result, the Sr3AsI3 structure has a higher dielectric constant peak.

 figure: Fig. 4.

Fig. 4. The real (a) and imaginary (b) portion of dielectric function with a photon energy of Sr3AsI3 perovskite.

Download Full Size | PDF

Figure 4(b) indicate the characteristics of the dielectric function’s imaginary portion, ɛ2(ω). The imaginary part of the dielectric function plays a crucial role in the analysis of optical absorption and the crystal structure's energy storage potential caused by unbiased charge excitations [38]. Information on the electronic bandgap, which is clearly relevant to the energy of the inter-band transitions near the Fermi level, was contributed by the imaginary dielectric function ɛ2(ω). A huge portion of the absorption region was taken up by the ɛ2(ω) values of Sr3AsI3. For Sr3AsI3, ɛ2(ω)’s greatest maxima are observed at the 5.5 value of the optical position, which demonstrated the amount of energy for absorption photon almost 2.1 eV, according to Fig. 4(b). The carriers’ valence to conduction band movement is determined by these imaginary absorption peaks. We also discovered that when photon energy is beyond 8.3 eV, the imaginary dielectric portion becomes zero. The lack of ɛ2(ω) (above 8.3 eV) reveals the material's low optical absorption and better optical transparency.

The “electron loss function” is the quantity of energy that is dissipated by electrons, when passing through a dielectric substrate [57]. The term “light-induced behavior of a substance” L(ω) refers to the study of how a substance responds to light. As seen by the peak in the plot of L(ω) for Sr3AsI3 in Fig. 5(a), the energy loss is discovered when the emitted photon's energy exceeds the material's bandgap. We can observe the loss function utilizing a formula, L(ω) = j ($({ - 1} )/\mathrm{\varepsilon }(\mathrm{\omega } )\; $). The peaks of L(ω) for the cubic structure of the Sr3AsI3 developed between 7 to 10 eV as indicated in Fig. 5(a). At a value of 8.5 eV, the predicted electron loss function, L(ω), for Sr3AsI3 was quite high. The greatest and lowest peaks of loss, separately, were found at 2.5 eV and 8.5 eV. According to the insignificant presence of L(ω) peaks below 2 eV, the Sr3AsI3 component would function as an efficient optical absorption layer in the area of the visible photon spectra and IR. The loss function of the Sr3AsI3 perovskite is found to exist for photon energies up to 10 eV. Overall, the loss function of Sr3AsI3 has a significant impact on performance, which is an important consideration when designing and optimizing these materials for specific applications.

 figure: Fig. 5.

Fig. 5. The loss function (a) and absorption (b) as a photon energy function of Sr3AsI3 perovskite.

Download Full Size | PDF

One essential quality of Sr3AsI3 is its absorption coefficient, which expresses how strongly a substance absorbs light at various wavelengths. The Sr3AsI3 perovskite's absorption coefficient is influenced by a number of factors, including the crystal structure, material purity, and thickness. For any configuration, the optical absorption coefficient profile displays characteristics that are slightly similar to those of the imaginary part of the dielectric constant [38]. The visible portion of the electromagnetic spectrum, which normally includes the majority of the sun's radiation, typically has a higher absorption coefficient. The absorption coefficient of the Sr3AsI3 perovskite material is shown in Fig. 5(b) as a measurement of photon energy. The absorption peak is high between 6 to 7 eV photon energy. Overall, the absorption coefficient of Sr3AsI3 perovskite is an important parameter to consider when designing photovoltaic devices based on this material. The significant variations in the absorption coefficient of perovskite in the energy range of 1.3 -10 eV make it appropriate for solar cells.

Reflectivity is the measure of how much light Sr3AsI3 perovskite reflects when exposed to electromagnetic radiation, such as visible light. The overall reflectivity of perovskite materials can differ significantly depending on factors like composition, crystal structure, and surface shape. Furthermore, the reflectivity of Sr3AsI3 perovskite may be influenced by the wavelength and angle of incidence of the incident light. As a measurement of photon energy, the reflectivity of the Sr3AsI3 perovskite material is depicted in Fig. 6. The range of photon energy where reflectivity changes the most is 0 to 5 eV. At 0 eV photon energy, the reflectivity reaches its maximum.

 figure: Fig. 6.

Fig. 6. The reflectivity as a photon energy function of Sr3AsI3 perovskite.

Download Full Size | PDF

The results from this study on Sr3AsI3's optical characteristics suit the findings from other reports quite well [60]. Materials that perform better in implementations for visible light devices have bandgaps lower than 3.1 eV [38]. Sr3AsI3 perovskite is a potential candidate for use in a number of applications, including solar cells, optoelectronics, and optical sensors due to its unique optical characteristics.

4. Conclusions

In conclusion, First-principles DFT computations have been used to examine the electronic, optical, and structural properties of the inorganic perovskite Sr3AsI3. Our analysis of the optical properties leads us to the conclusion that the Sr3AsI3 exhibits the peaks of absorption in the vicinity of the ultra-violate to visual spectrum. The perovskite Sr3AsI3 material's direct bandgap was also measured to be 1.265 eV. Furthermore, the peak of the dielectric function of Sr3AsI3 transfers to greater energy of the photon. Our opinion is that our latest investigation will encourage further research to develop Sr3AsI3 as well-known perovskite for use in optoelectronic devices.

Acknowledgment

The authors are satisfied to Department of Electrical and Electronic Engineering Begum Rokeya University, Rangpur 5400, Bangladesh due to using Advanced Energy Materials and Solar Cell Research Laboratory.

Disclosures

The authors have no conflicts of interest.

Data availability

Data will be made available on request.

References

1. X. Huang, D. Ji, H. Fuchs, et al., “Recent Progress in Organic Phototransistors: Semiconductor Materials, Device Structures and Optoelectronic Applications,” ChemPhotoChem 4(1), 9–38 (2020). [CrossRef]  

2. T. W. Kelley, P. F. Baude, C. Gerlach, et al., “Recent Progress in Organic Electronics: Materials, Devices, and Processes,” Chem. Mater. 16(23), 4413–4422 (2004). [CrossRef]  

3. H. Wang, Z. Zeng, P. Xu, et al., “Recent progress in covalent organic framework thin films: fabrications, applications and perspectives,” Chem. Soc. Rev. 48(2), 488–516 (2019). [CrossRef]  

4. M. F. Rahman, M. M. Alam Moon, M. K. Hossain, et al., “Concurrent investigation of antimony chalcogenide (Sb2Se3 and Sb2S3)-based solar cells with a potential WS2 electron transport layer,” Heliyon 8(12), e12034 (2022). [CrossRef]  

5. M. M. A. Moon, M. H. Ali, M. F. Rahman, et al., “Design and simulation of FeSi 2-based novel heterojunction solar cells for harnessing visible and near-infrared light,” Phys. Status Solidi A 217(6), 1900921 (2020). [CrossRef]  

6. M. A. Moon, H. Ali, and F. Rahman, “Investigation of thin-film p-BaSi2/n-CdS heterostructure towards semiconducting silicide based high efficiency solar cell,” Phys. Scr. 95(3), 035506 (2020). [CrossRef]  

7. M. F. Rahman, M. J. A. Habib, M. H. Ali, et al., “Design and numerical investigation of cadmium telluride (CdTe) and iron silicide (FeSi2) based double absorber solar cells to enhance power conversion efficiency,” AIP Adv. 12(10), 1–11 (2022). [CrossRef]  

8. M. M. A. Moon, M. F. Rahman, M. Kamruzzaman, et al., “Unveiling the prospect of a novel chemical route for synthesizing solution-processed CdS/CdTe thin-film solar cells,” Energy Reports 7, 1742–1756 (2021). [CrossRef]  

9. A. Isha, A. Kowsar, A. Kuddus, et al., “High efficiency Cu2MnSnS4 thin film solar cells with SnS BSF and CdS ETL layers: A numerical simulation,” Heliyon 9(5), e15716 (2023). [CrossRef]  

10. M. K. Hossain, D. P. Samajdar, R. C. Das, et al., “Design and simulation of Cs2BiAgI6 double perovskite solar cells with different electron transport layers for efficiency enhancement,” Energy Fuels 37(5), 3957–3979 (2023). [CrossRef]  

11. S. Bhattarai, M. K. Hossain, R. Pandey, et al., “Perovskite solar cells with dual light absorber layers for performance efficiency exceeding 30%,” Energy Fuels 7(14), 10631–10641 (2023). [CrossRef]  

12. M. K. Hossain, M. K. A. Mohammed, R. Pandey, et al., “Numerical analysis in DFT and SCAPS-1D on the influence of different charge transport layers of CsPbBr3 perovskite solar cells,” Energy and Fuels 37(8), 6078–6098 (2023). [CrossRef]  

13. M. Khalid, E. Commission, A. Energy, et al., “Design Insights of La2NiMnO6-based perovskite solar cell employing different charge transport layers : DFT and SCAPS-1D frameworks,” Energy Fuels 37(17), 13377–13396 (2023). [CrossRef]  

14. A. Kojima, K. Teshima, Y. Shirai, et al., “Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells,” J. Am. Chem. Soc. 131(17), 6050–6051 (2009). [CrossRef]  

15. M. K. Hossain, M. H. K. Rubel, G. F. I. Toki, et al., “Effect of various electron and hole transport layers on the performance of CsPbI 3-based perovskite solar cells: a numerical investigation in DFT, SCAPS-1D, and wxAMPS frameworks,” ACS Omega 7(47), 43210–43230 (2022). [CrossRef]  

16. M. K. Hossain, G. F. I. Toki, I. Alam, et al., “Numerical simulation and optimization of CsPbI3-based perovskite solar cell to enhance the power conversion efficiency,” New J. Chem. 47(10), 4801–4817 (2023). [CrossRef]  

17. D. Saikia, A. Betal, J. Bera, et al., “Progress and challenges of halide perovskite-based solar cell- a brief review,” Mater. Sci. Semicond. Process. 150, 106953 (2022). [CrossRef]  

18. Y. Yuan, J. Chae, Y. Shao, et al., “Photovoltaic switching mechanism in lateral structure hybrid perovskite solar cells,” Adv. Energy Mater. 5(15), 1500615 (2015). [CrossRef]  

19. S. D. Stranks, G. E. Eperon, G. Grancini, et al., “Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber,” Science 342(6156), 341–344 (2013). [CrossRef]  

20. Q. A. Akkerman, M. Gandini, F. Di Stasio, et al., “Strongly emissive perovskite nanocrystal inks for high-voltage solar cells,” Nat. Energy 2(2), 16194 (2016). [CrossRef]  

21. M. Shirayama, H. Kadowaki, T. Miyadera, et al., “Optical transitions in hybrid perovskite solar cells: ellipsometry, density functional theory, and quantum efficiency analyses for CH3NH3,” Phys. Rev. Appl. 5(1), 014012 (2016). [CrossRef]  

22. H. Zhu, M. T. Trinh, J. Wang, et al., “Organic cations might not be essential to the remarkable properties of band edge carriers in lead halide perovskites,” Adv. Mater. 29(1), 1603072 (2017). [CrossRef]  

23. P. Pitriana, T. D. K. Wungu, R. Herman, et al., “The characteristics of band structures and crystal binding in all-inorganic perovskite APbBr3 studied by the first principle calculations using the Density Functional Theory (DFT) method,” Results Phys. 15, 102592 (2019). [CrossRef]  

24. H. Zitouni, N. Tahiri, O. El Bounagui, et al., “How the strain effects decreases the band gap energy in the CsPbX 3 perovskite compounds?” Phase Transitions 93(5), 455–469 (2020). [CrossRef]  

25. G. E. Eperon, G. M. Paternò, R. J. Sutton, et al., “Inorganic caesium lead iodide perovskite solar cells,” J. Mater. Chem. A 3(39), 19688–19695 (2015). [CrossRef]  

26. B. Ehrler, E. Alarcón-Lladó, S. W. Tabernig, et al., “Photovoltaics reaching for the Shockley–Queisser limit,” ACS Energy Lett. 5(9), 3029–3033 (2020). [CrossRef]  

27. M. K. Hossain, G. F. I. Toki, A. Kuddus, et al., “An extensive study on multiple ETL and HTL layers to design and simulation of high - performance lead - free CsSnCl 3 - based perovskite solar cells,” Sci. Rep. 13(1), 2521 (2023). [CrossRef]  

28. M. A. Green, Y. Jiang, A. M. Soufiani, et al., “Optical Properties of Photovoltaic Organic–Inorganic Lead Halide Perovskites,” J. Phys. Chem. Lett. 6(23), 4774–4785 (2015). [CrossRef]  

29. J. Deng, J. Li, Z. Yang, et al., “All-inorganic lead halide perovskites: a promising choice for photovoltaics and detectors,” J. Mater. Chem. C 7(40), 12415–12440 (2019). [CrossRef]  

30. D. R. Hamann, M. Schlüter, and C. Chiang, “Norm-Conserving Pseudopotentials,” Phys. Rev. Lett. 43(20), 1494–1497 (1979). [CrossRef]  

31. J. M. Smith, S. P. Jones, and L. D. White, “Rapid communication,” Gastroenterology 72(1), 193 (1977). [CrossRef]  

32. G. Kresse and J. Hafner, “Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements,” J. Phys.: Condens. Matter 6(40), 8245–8257 (1994). [CrossRef]  

33. J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized gradient approximation made simple,” Phys. Rev. Lett. 77(18), 3865–3868 (1996). [CrossRef]  

34. P. Giannozzi, O. Baseggio, P. Bonfà, et al., “Quantum ESPRESSO toward the exascale,” J. Chem. Phys. 152(15), 154105 (2020). [CrossRef]  

35. P. Giannozzi, O. Andreussi, T. Brumme, et al., “Advanced capabilities for materials modelling with Quantum ESPRESSO,” J. Phys.: Condens. Matter 29(46), 465901 (2017). [CrossRef]  

36. J. P. Perdew and A. Zunger, “Self-interaction correction to density-functional approximations for many-electron systems,” Phys. Rev. B 23(10), 5048–5079 (1981). [CrossRef]  

37. P. Giannozzi, S. Baroni, N. Bonini, et al., “QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials,” J. Phys.: Condens. Matter 21(39), 395502 (2009). [CrossRef]  

38. R. Islam, K. Liu, Z. Wang, et al., “Strain-induced electronic and optical properties of inorganic lead halide perovskites APbBr3 (A = Rb and Cs),” Mater. Today Commun. 31, 103305 (2022). [CrossRef]  

39. Z. Wu and R. E. Cohen, “More accurate generalized gradient approximation for solids,” Phys. Rev. B 73(23), 235116 (2006). [CrossRef]  

40. J. P. Perdew, A. Ruzsinszky, G. I. Csonka, et al., “Restoring the density-gradient expansion for exchange in solids and surfaces,” Phys. Rev. Lett. 100(13), 136406 (2008). [CrossRef]  

41. F. Karsch, A. Patkós, and P. Petreczky, “Screened perturbation theory,” Phys. Lett. B 401(1-2), 69–73 (1997). [CrossRef]  

42. M. Roknuzzaman, K. Ostrikov, H. Wang, et al., “Towards lead-free perovskite photovoltaics and optoelectronics by ab-initio simulations,” Sci. Rep. 7(1), 14025 (2017). [CrossRef]  

43. A. A. Emery and C. Wolverton, “High-throughput DFT calculations of formation energy, stability and oxygen vacancy formation energy of ABO3 perovskites,” Sci. Data 4(1), 170153 (2017). [CrossRef]  

44. A. Depeursinge, D. Racoceanu, J. Iavindrasana, et al., “Fusing visual and clinical information for lung tissue classification in HRCT data,” Artificial Intelligence in Medicine 50(1), 13–21 (2010). [CrossRef]  

45. T. Li, C. He, and W. Zhang, “Rational design of porous carbon allotropes as anchoring materials for lithium sulfur batteries,” J. Energy Chem. 52, 121–129 (2021). [CrossRef]  

46. H. J. Feng and Q. Zhang, “Predicting efficiencies >25% A3MX3photovoltaic materials and Cu ion implantation modification,” Appl. Phys. Lett. 118(11), 1 (2021). [CrossRef]  

47. M. H. K. Rubel, M. A. Hossain, M. K. Hossain, et al., “First-principles calculations to investigate structural, elastic, electronic, thermodynamic, and thermoelectric properties of CaPd3B4O12 (B = Ti, V) perovskites,” Results Phys. 42, 105977 (2022). [CrossRef]  

48. R. Islam, F. Rahman, S. Bhattarai, et al., “Photovoltaic performance investigation of Cs3Bi2I9-based perovskite solar cells with various charge transport channels using density functional theory and one-dimensional solar cell capacitance simulator frameworks,” Energy Fuels 37(10), 7380–7400 (2023). [CrossRef]  

49. D. G. Hinks, B. Dabrowski, J. D. Jorgensen, et al., “Synthesis, structure and superconductivity in the Ba1−xKxBiO3−y system,” Nature 333(6176), 836–838 (1988). [CrossRef]  

50. T. Nishio, J. Ahmad, and H. Uwe, “Spectroscopic Observation of Bipolaronic Point Defects in Ba1-xKxBiO3,” Phys. Rev. Lett. 95(17), 176403 (2005). [CrossRef]  

51. M. K. Hossain, A. A. Arnab, R. C. Das, et al., “Combined DFT, SCAPS-1D, and wxAMPS frameworks for design optimization of efficient Cs2BiAgI6-based perovskite solar cells with different charge transport layers,” RSC Adv. 12(54), 34850–34873 (2022). [CrossRef]  

52. P. Ramasamy, D.-H. Lim, B. Kim, et al., “All-inorganic cesium lead halide perovskite nanocrystals for photodetector applications,” Chem. Commun. 52(10), 2067–2070 (2016). [CrossRef]  

53. R. Rajeswarapalanichamy, A. Amudhavalli, R. Padmavathy, et al., “Band gap engineering in halide cubic perovskites CsPbBr3−yIy (y = 0, 1, 2, 3) – A DFT study,” Mater. Sci. Eng., B 258, 114560 (2020). [CrossRef]  

54. D. C. Langreth and M. J. Mehl, “Beyond the local-density approximation in calculations of ground-state electronic properties,” Phys. Rev. B 28(4), 1809–1834 (1983). [CrossRef]  

55. J. Yang, L. Z. Tan, and A. M. Rappe, “Hybrid functional pseudopotentials,” Phys. Rev. B 97(8), 085130 (2018). [CrossRef]  

56. X. Yang and W. A. Daoud, “Triboelectric and Piezoelectric Effects in a Combined Tribo-Piezoelectric Nanogenerator Based on an Interfacial ZnO Nanostructure,” Adv. Funct. Mater. 26(45), 8194–8201 (2016). [CrossRef]  

57. M. R. Islam, M. R. H. Mojumder, R. Moshwan, et al., “Strain-Driven Optical, Electronic, and Mechanical Properties of Inorganic Halide Perovskite CsGeBr 3,” ECS J. Solid State Sci. Technol. 11(3), 033001 (2022). [CrossRef]  

58. A. B. Kuzmenko, “Kramers–Kronig constrained variational analysis of optical spectra,” Rev. Sci. Instrum. 76(8), 083108 (2005). [CrossRef]  

59. Z. Xu, “The determination of the momentum matrix elements involved in calculating the dielectric constants of superlattices using the tight-binding method,” Solid State Commun. 76(9), 1143–1147 (1990). [CrossRef]  

60. K. M. Hossain, M. Z. Hasan, and M. L. Ali, “Narrowing bandgap and enhanced mechanical and optoelectronic properties of perovskite halides: Effects of metal doping,” AIP Adv. 11(1), 015052 (2021). [CrossRef]  

Data availability

Data will be made available on request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) The inorganic perovskite Sr3AsI3's optimum structure. (b) The first Brillouin zone's k-path in order to determine their electronic band structure.
Fig. 2.
Fig. 2. The charge density of Sr3AsI3 structure (a) 2D view, (b) Bird’s eye view.
Fig. 3.
Fig. 3. The electronic (a) band structure and (b) refined structure of PDOS for inorganic Sr3AsI3 perovskite with the function of PBE/HSE.
Fig. 4.
Fig. 4. The real (a) and imaginary (b) portion of dielectric function with a photon energy of Sr3AsI3 perovskite.
Fig. 5.
Fig. 5. The loss function (a) and absorption (b) as a photon energy function of Sr3AsI3 perovskite.
Fig. 6.
Fig. 6. The reflectivity as a photon energy function of Sr3AsI3 perovskite.

Tables (1)

Tables Icon

Table 1. Sr3AsI3's lattice constant and energy bandgap were determined using experimental data and previous DFT calculations.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

ε ( ω ) = ε 1 ( ω ) + i ε 2 ( ω )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.