Recently, inorganic perovskite materials have been attracting increasing attention owing to their exceptional structural, electronic, and optical characteristics in photovoltaic technology. Ca3AsI3 is a semiconductor material that shares similarities with the group of inorganic metal halide perovskites. Ca3AsI3 possesses a perovskite crystal structure that is cubic, which is classified under the space group Pm-3m (no. 221). Our research aims to analyze how the optical and electronic properties of Ca3AsI3 are influenced by spin–orbit coupling (SOC) and strain using the first-principles density-functional theory. The inorganic Ca3AsI3 perovskite has an electronic band structure that possesses a direct bandgap of roughly 1.58 eV at the Γ(gamma)-point. However, when the SOC relativistic effect is introduced, this value decreases to around 1.27 eV. As the level of compressive strain is increased, the bandgap becomes narrower, whereas with increasing tensile strain, the bandgap becomes wider. It has been observed through analysis of the dielectric functions, absorption coefficient, and electron loss function of these materials that the optical properties give Ca3AsI3 the ability to effectively absorb visible light. According to the study, the dielectric constant peaks of Ca3AsI3 shift toward a lower photon energy (redshift) as the level of compressive strain increases. On the other hand, when subjected to increased tensile strain, these peaks have a tendency to shift toward a higher photon energy (blueshift), as per the same study. Modifying the energy gap of Ca3AsI3 perovskites to suit optoelectronic and solar cell needs could be achieved by using techniques involving the SOC effect and by applying strain. These approaches have the potential to enable utilization of Ca3AsI3 in such applications in the future.

The utilization of photovoltaic (PV) technology in the energy sector is notable due to its substantial capacity to displace fossil fuels by diminishing oil consumption, mitigating greenhouse gas (GHG) emission, and mitigating environmental degradation.1 Silicon (Si), GaAs, CdTe, and Cu(In,Ga)(S,Se)2 (CIGS) solar cell materials have made tremendous advances in the PV area.2–5 Because of its high power conversion efficiency (PCE), stability, and radiation hardness, crystalline silicon (Si) takes the lead in their commercialization.6–8 Traditional solar cells such as Si, however, have limited potential for advancement because of their indirect, suboptimal bandgaps and expensive fabrication processes.9 Halide perovskites were introduced in 2009.10 The remarkable properties of organic–inorganic halide perovskites (OILHPs), such as their reasonable bandgap, ample availability, impressive ability to absorb visible light, decent reflectivity, and affordable manufacturing costs, have generated a great deal of research interest in solar technology.10–16 Their PCE has increased quickly in the past thirteen years, from 3.8% to 25.8%.17–20 Unfortunately, the long-term durability issue of OILHPs is currently a major problem because of their susceptibility to moisture, wind, sunlight, and temperature when utilized in a real environment. Moreover, the application and development of perovskite cells on a broad scale are unfortunately restricted by Pb’s low stability, device hysteresis, and toxicity. The three significant obstacles that must be removed before it can be widely commercialized are material toxicity, device hysteresis, and perovskite material stability. Toxic lead needs to be replaced by other nontoxic materials because concerns about the environment arise from the use of such poisonous material.21–25 Recently, A3BX3 type perovskites have attracted great attention due to their exceptional properties such as being a Pb free and direct bandgap material and having mechanical stability and better electronic properties. To date, there have been no comprehensive reports on Ca3AsI3. Therefore, a comprehensive study of inorganic Ca3AsI3 perovskites is essential for future electronic and optoelectronic device applications.

Numerous investigations have been conducted in recent times using strain engineering, a technique that enables the improvement of specific characteristics of a substance.26–30 Strain engineering has been identified as a successful method that can modify the configurations and electronic characteristics of organic–inorganic halide perovskites, leading to a better comprehension of the correlation between the structure and properties.31–37 The study also involved an examination of how pressure affected the structure and bandgap of CsGeBr3 and CsGeCl3.38,39 One way to increase the photovoltaic performance of CaAsI3 is to restrict its bandgap to within the 1.2–1.4 eV range. If subjected to compressive strain, the inorganic cubic perovskite CsSnCl3 has the potential to demonstrate remarkable optoelectronic properties by transitioning from a semiconductor to a metallic state.40,41 Applying either compressive or tensile strain can effectively modify the properties of the bandgap and dielectric function of the CsGeI3 perovskite.42 Modulation of electronic properties in different materials can generally be achieved through the spin–orbit coupling (SOC) effect.43,44 While light elements exhibit minimal SOC, heavier atoms display a more prominent SOC effect.45 The generation of SOC is attributed to the relativistic effect, which causes the alignment of electron spins and the motion of electrons in their orbitals.46 As far as we know, no reports have been published to date regarding the electronic and optical characteristics of Ca3AsI3 subjected to both strain and SOC. Therefore, a comprehensive study of the effect of strain and SOC on inorganic Ca3AsI3 perovskites is essential for future electronic and optoelectronic device applications.

The objective of this study is to use first-principles density-functional theory (FP-DFT) calculations to investigate the structural, electronic, and optical characteristics of cubic Ca3AsI3 when subjected to varying levels of strain and SOC. Our investigation involves the use of two different approaches to examine the band structure of Ca3AsI3. In our research, we specifically examined the alterations in the bandgap resulting from changes in both strain and SOC, encompassing both increases and decreases. Hence, conducting this research would be advantageous in determining the SOC and strain-specific characteristics of Ca3AsI3 inorganic materials and their practical use in next-gen solar cells and energy storage devices.

The Quantum ESPRESSO simulation package is used to implement the DFT calculations.47 On the structures of the Ca3AsI3 perovskite, the FP-DFT with a norm-conserving (NC) pseudopotential48,49 and the Perdew–Burke–Ernzerhof (PBE)50 exchange-correlation function were used. The input data came pre-configured with the Brillouin zone grid, crystal formations, lattice parameters, and kinetic cut-off energy. To optimize the structure and perform self-consistent function (SCF) and non-self-consistent function (NSCF) calculations, a kinetic energy cut-off of 30 Rydberg (Ry) (∼410 eV) and a charge density cut-off of 220 Ry (∼2990 eV) were employed. A maximum force tolerance of less than 0.01 eV/Å and a convergence threshold of 10−6 a.u. were established for the SCF calculations. The force convergence threshold for ionic minimization in the relaxation calculations is set to 10−3 a.u. The Brillion zone’s 8 × 8 × 8 Monkhorst–Pack k-grid was used to calculate the band structure and partial density of states (PDOS). Local-density approximation (LDA) and generalized gradient approximation (GGA) functionals were employed to calculate specific organic halides,51 and it was shown that the absolute mean error is more than ten pm. Although there are certain ways to reduce this error, we have not used corrected PBE functions for metals in the most recent studies.47,48 By simultaneously modifying the lattice parameter in our investigation, it was also found that the level of biaxial strain, which involves both compression and tension, fluctuates between −6% and +6%, in increments of 1%. To determine the strain, we employed the following formula:31 
(1)
In the formula, arelaxed is regarded as the unstrained lattice constant, and the range of ε varies from −6% to +6% in units of 1%. By definition, arelaxed represents the unstrained lattice constant in this expression, while the value can shift to as much as +6% from −6% (in 1% increments). Therefore, the presence of compressive and tensile strains is indicated by negative and positive values, respectively.

The analysis of the dynamical stability of material structures and the investigation of optical properties involved the implementation of first-order time-dependent perturbation theory. In order to perform calculations related to the optical properties, a Monkhorst–Pack k-point grid consisting of an 8 × 8 × 8 Γ-centered configuration is employed to sample the Brillouin zone. After that, we analyzed the complex dielectric function to identify the energy range (in eV) in which it exhibits absorption peaks. The real and imaginary parts of the complex dielectric function are added together to obtain ε(ω) = ɛ1 (ω) + jɛ2 (ω).

The Ca3AsI3 material is depicted in Fig. 1(a) as being a member of the Pm-3m (no. 221) cubic space group.52  Figure 1(a) demonstrates that the unit cell of the structure is composed of seven atoms, with Ca atoms present at 3c (0.5, 0, 0.5) sites, As atoms are situated at 1b (0.5, 0.5, 0.5) sites, and I atoms occupy the 3d (0, 0.5, 0) sites. The first Brillouin zone’s (BZ) k-path is illustrated in Fig. 1(b), with emphasis on the high symmetrical BZ points (Γ-X-M-R-Γ) that are essential for the electronic band structure of Ca3AsI3. The band diagram’s recurrence across these points is observed throughout the entire structure. After optimization, the computed lattice parameter for Ca3AsI3 was determined to be 6.27 Å, which is considered the most favorable structure.

FIG. 1.

(a) Proposed optimized structure of the novel inorganic halide perovskite Ca3AsI3 and (b) the first Brillouin zone’s k-path in order to determine its electronic band structure.

FIG. 1.

(a) Proposed optimized structure of the novel inorganic halide perovskite Ca3AsI3 and (b) the first Brillouin zone’s k-path in order to determine its electronic band structure.

Close modal

The density of charges in Ca3AsI3 with respect to the (200) crystallographic plane is illustrated in Fig. 2(a) as a 2D map. The colors on the right-hand side of the image’s scale bar indicate the electron density intensity. In Fig. 2(b), there is a depiction of the charge density from a top-down perspective, similar to a bird’s eye view.

FIG. 2.

Arrangement of density of charges in Ca3AsI3 (a) with respect to the (200) plane in 2D, (b) in a bird’s eye view, and (c) in 3D view.

FIG. 2.

Arrangement of density of charges in Ca3AsI3 (a) with respect to the (200) plane in 2D, (b) in a bird’s eye view, and (c) in 3D view.

Close modal

This vantage point allows for a clear comprehension of how the charges disperse around the atoms with varying levels of concentration. The charge distribution is illustrated from a 3D perspective in Fig. 2(c). It is evident that the positively charged particles tend to aggregate near the Ca (calcium) atom and the arrangement of charges appears to be uniform. The I (iodine) atom was encircled by a considerable number of negative charges.

As a result of the overlap between the charges of the Ca (calcium) atom and the As (arsenic) atom, it is not feasible to distinguish the As atom individually in the charge density diagram. Consequently, the Ca and As ions developed a covalent bond.53,54 The charge distribution map provided further evidence that the Ca (calcium) and I (iodine) ions have formed an ionic bond as their charge contours were found to be non-coincidental.19,54 The ionic bonding mechanism reinforces the bond strength within the structural unit, while the covalent bond between Ca and As ions is leveraged to create a less robust bond. Conversely, As (arsenic) and I (iodine) exhibit antibonding characteristics.

After structural optimization, we computed the electronic band structure and highly symmetric point direction of Ca3AsI3. Typically, for a system to be a good candidate for application in optoelectronics devices, a direct bandgap is required.55–57  Figure 3(a) illustrates the band structures of unstrained Ca3AsI3 perovskite formations. In order to interpret the bandgap value from the figure, the Fermi level is fixed at zero. The Γ-X-M-R-Γ path of the cubic structure is being analyzed with respect to the k-axis, and Fig. 3(a) depicts the conduction band minimum (CBM) and valence band maximum (VBM), both of which are situated at the Γ-point.

FIG. 3.

(a) Electronic band structure and (b) partial density of states (PDOS) using the PBE function of the proposed structure of Ca3AsI3.

FIG. 3.

(a) Electronic band structure and (b) partial density of states (PDOS) using the PBE function of the proposed structure of Ca3AsI3.

Close modal

Based on the results obtained using the PBE function, it can be concluded that the Ca3AsI3 material being studied has a direct bandgap structure, with a value of ∼1.58 eV. These findings demonstrate a high level of conformity with the values that were previously documented.32 The evaluation of bandgap using the GGA approach often results in underestimation of its value, which is a well-known drawback of this method. In addition, the LDA+U and LDA methods also found that the bandgap value was underestimated.33 To overcome this kind of bandgap computation, some researchers have provided a variety of approaches, including the GW method and hybrid functional.34,35 Furthermore, a direct bandgap is a necessary characteristic that a crystalline substance must possess to be appropriate for advanced photo-thermal and renewable energy purposes. These compounds are considered ideal for efficient solar cells and cells used in photovoltaic applications because of their high value for direct bandgap.

The partial density of states (PDOS), which often depicts the influence of individual atoms and their various states on the bandgap energy of Ca3AsI3 formation, usually reflects this. The spread of PDOS for Ca3AsI3 structures is shown in Fig. 3(b) in the region of −3 to 3 eV. In Ca3AsI3, the places of Ca and As that have been hybridized with I extend across the full energy level while sparing the bandgap. It shows that the primary kind of bonding for Ca–I and As–I is covalent. Moreover, due to significant atomic position variations in Ca3Asl3 [Fig. 3(b)] formation, the electron charge was transferred from Ca and As to I. According to the analysis of the cubic phase, it can be said that I-5p orbitals play a significant role in referring to the valence band of every perovskite structure. The main contribution to the conduction band is from the As-4p orbital, whereas the Ca-4s orbital has very little effect.

To consider the impression of SOC on the electronic features of the Ca3AsI3 perovskite, we factored in the Hamiltonian equation that involves SOC,58 
(2)
Here, Hsoc refers to the Hamiltonian operator for SOC, where represents the reduced Planck’s constant, p represents the orbital angular momentum, F refers to the potential energy or force, m0 denotes the mass of free electrons, and s indicates the spin angular momentum.

To perform relativistic calculations, we utilized the PBE functional method to detect the size of the relativistic effect and assess how the heavy element Ca as well as As influences the SOC-included electronic structures of Ca3AsI3 perovskites. The impact of SOC on the conduction and valence band regions was significant, leading to changes in the positions of the CBM and VBM, as illustrated in Figs. 4(a) and 4(b). The VBM exhibited significant alterations in its upward shift, while the CBM underwent a downward shift toward the Fermi level. However, the bandgap energy value of Ca3Asl3 caused by the SOC is 1.27 eV. Ca3AsI3 perovskite structures exhibit a decrease in the bandgap value as a result of the SOC effect. This change in the bandgap is illustrated in Fig. 4(a).

FIG. 4.

(a) Band structure and (b) PDOS using the PBE function of the proposed structure of Ca3AsI3 with the SOC effect.

FIG. 4.

(a) Band structure and (b) PDOS using the PBE function of the proposed structure of Ca3AsI3 with the SOC effect.

Close modal

One could obtain a more thorough comprehension of the band construction of cubic Ca3AsI3 by investigating the PDOS while considering the influence of the SOC. The inclusion of the SOC effect results in alterations to the conduction and valence bands as a result of the distinct electron influence compared to the PDOS when SOC is not considered. The PDOS illustrated in Fig. 4(b) demonstrates that there is no consistent pattern observed in the impression of the Ca atom when SOC is included. Ca and I atoms have a slight effect on the conduction-band area between 0.635 and 1.75 eV, but the influence of I-5p(j = 0.5) and I-5p(j = 1.5) is noticeable in the valence-band region between −3 and −0.635 eV. The As-4p(j = 0.5) and As-4p(j = 1.5) atoms have a significant impact on the conduction band area between 0.635 and 1.75 eV. However, Ca and I have a smaller effect on the valence-band region between −3 and −0.635 eV.

Investigations were carried out on how strain affected the Ca3AsI3 bandgap when incorporating the SOC effect and without it. Ca3AsI3 bandgaps scaled between 1.14 and 1.84 eV after the strain was applied in the effective range of −6% to 6% with a 1% space between each value. Figure 5 shows how the Ca3AsI3 band topologies change under biaxial compressive and tensile strain. During compressive strain, it is discovered that the VBM and CBM are pushed toward the Fermi level (−6% to 0%) for Ca3AsI3 [Figs. 5(a) and 5(b)] but the CBM and VBM are kept at the Γ (gamma) point. When the SOC effect of Ca3AsI3 production is taken into consideration, the band configurations under compressive strain are as shown in Fig. 6(a), where Fig. 6(b) shows a magnified view. When compressive strain is applied, the orbital overlap causes a reduction in the length of the Ca, As, and I bonds. Covalent bonding is produced by the orbital overlap because Ca, As, and I are strongly hybridized when incorporating the SOC effect and without it. The bandgap is shown to decrease with increasing compressive strain (i.e., incorporating the SOC effect and without it). As shown in Fig. 5(c) for Ca3AsI3, the electronic band structure similarly varies when subjected to tensile strain (0%–6%). The bandgap may widen under tensile tension because the VBM and CBM have been moved from the Fermi level. To enable a lowered force between the atoms of Ca, As, and I, the bond length and atomic distance between the atoms are enhanced. The bandgap ultimately widens for the Ca3AsI3 structure under tensile tension, which is shown in a magnified view in Fig. 5(d). The gamma point (Γ-point) under compressive and tensile strain, incorporating the SOC effect and without it, is shown in a magnified view in Figs. 6(b), 5(b), 6(d), and d5(d).

FIG. 5.

Electronic bands of Ca3AsI3 under (a) and (c) compressive and tensile strain; (b) and (d) magnified view of compressive and tensile strains incorporating the SOC effect.

FIG. 5.

Electronic bands of Ca3AsI3 under (a) and (c) compressive and tensile strain; (b) and (d) magnified view of compressive and tensile strains incorporating the SOC effect.

Close modal
FIG. 6.

Electronic bands of Ca3AsI3 under (a) and (c) compressive and tensile strain and (b) and (d) magnified view of compressive and tensile strains incorporating the SOC effect.

FIG. 6.

Electronic bands of Ca3AsI3 under (a) and (c) compressive and tensile strain and (b) and (d) magnified view of compressive and tensile strains incorporating the SOC effect.

Close modal

The changes in the bandgap for Ca3AsI3 structures when subjected to compressive and tensile strains, incorporating the SOC effect and without it, are presented in Table I and Fig. 7 for reference. According to these findings, the bandgap reduces virtually linearly as compressive strain increases (incorporating the SOC effect and without it), but it increases linearly when tensile strain increases (Fig. 7). Surprisingly, Ca3AsI3 continues to be a direct-bandgap semiconductor throughout the whole strain region. Ca3AsI3 has a bandgap between 1.3 and 1.4 eV for both compressive strain (without SOC) and tensile strain (with SOC). Based on the Shockley–Queisser hypothesis, it is proposed that this structure will perform the best in terms of photovoltaics under these circumstances.59 

TABLE I.

The anticipated electronic bandgap values of Ca3AsI3 under different levels of compressive and tensile strains, incorporating the SOC effect and without it.

Compressive strain’s bandgap value (eV)Tensile strain’s bandgap value (eV)
Applied strain (%)Absence of SOCPresence of SOCAbsence of SOCPresence of SOC
1.58 1.27 1.58 1.27 
1.53 1.21 1.647 1.31 
1.46 1.12 1.66 1.34 
1.41 1.085 1.716 1.39 
1.38 1.024 1.77 1.40 
1.247 0.93 1.796 1.448 
1.14 0.84 1.84 1.48 
Compressive strain’s bandgap value (eV)Tensile strain’s bandgap value (eV)
Applied strain (%)Absence of SOCPresence of SOCAbsence of SOCPresence of SOC
1.58 1.27 1.58 1.27 
1.53 1.21 1.647 1.31 
1.46 1.12 1.66 1.34 
1.41 1.085 1.716 1.39 
1.38 1.024 1.77 1.40 
1.247 0.93 1.796 1.448 
1.14 0.84 1.84 1.48 
FIG. 7.

Changes in the energy bandgap of the Ca3AsI3 structure in response to applied strain and incorporating the SOC effect.

FIG. 7.

Changes in the energy bandgap of the Ca3AsI3 structure in response to applied strain and incorporating the SOC effect.

Close modal

Furthermore, the orbital distributions of the particles (Ca, As, and I) change as strain is introduced. Figures 8(a)8(c) show the PDOS of Ca3AsI3 under compressive strain, while Figs. 8(d)8(f) demonstrate the PDOS under tensile strain. Figures 8(a)8(c) show the PDOS of Ca3AsI3 under compressive strain, while Figs. 8(d)8(f) demonstrate the PDOS under tensile strain, Fig. 8 shows the PDOS incorporated with the SOC effect. As the strain transitions from −6% to +6%, the partial density of states (PDOS) related to the 5p orbital of the I atoms shows a nearness to the valence band side, which is positioned below the Fermi level. The density of states (DOS) within the conduction band region is greatly influenced by the presence of the 4p orbital of the As atoms. While the static position and structure of the occupied density of states (DOS) for the As-4p orbital remain unchanged on the conduction band side, the total contribution to the DOS increases as the applied strain is varied between −6% and +6%. The figure makes it evident that a higher tensile strain results in an increased total density of states (DOS) for the Ca3AsI3 structure. Specifically, there is a shift from around 18.5 electrons/eV at −6% strain to about 30 electrons/eV at +6% strain. Except for the bandgap region, hybridized orbitals of Ca–As and Ca–I can be seen across the entire energy range.

FIG. 8.

Partial density of states (PDOS) of Ca3AsI3 under various applied compressive and tensile strains, namely, (a) −6%, (b) −4%, (c) −2%, (d) +2%, (e) +4%, and (f) +6%.

FIG. 8.

Partial density of states (PDOS) of Ca3AsI3 under various applied compressive and tensile strains, namely, (a) −6%, (b) −4%, (c) −2%, (d) +2%, (e) +4%, and (f) +6%.

Close modal

To obtain a precise estimation of the band structure, the computation takes into account the SOC effect. Figures 9(a)9(f) indicate that the SOC effect is significant in both the conduction and valence band regions, particularly when there are fluctuations in the levels of the conduction band minimum (CBM) and valence band maximum (VBM). The 5p(j = 0.5) and 5p(j = 1.5) orbitals of the I atom exhibit activity near the valence band side, positioned slightly near the Fermi level. Similarly, closer to the Fermi level, the 4p(j = 0.5) and 4p(j = 1.5) orbitals of the As atom exhibit activity toward the conduction band side. Thus, the As atom’s 4p orbitals dominate the total DOS in the conduction band section.

FIG. 9.

Partial density of states (PDOS) of Ca3AsI3 under various pressures and stretching forces with SOC, namely, (a) −6%, (b) −4%, (c) −2%, (d) +2%, (e) +4%, and (f) +6%.

FIG. 9.

Partial density of states (PDOS) of Ca3AsI3 under various pressures and stretching forces with SOC, namely, (a) −6%, (b) −4%, (c) −2%, (d) +2%, (e) +4%, and (f) +6%.

Close modal

Taking the SOC effect into account, the performance of the total density of states (TDOS) increases as the applied strain is altered from −6% to +6%. Figure 9 shows that when taking the SOC effect into account, the total density of states (TDOS) for the Ca3AsI3 configuration is 14 electrons/eV at a strain of −6% and 19 electrons/eV at a strain of +6%. Figures 10(a) and 10(b) illustrate the TDOS of Ca3AsI3 for different biaxial compressive and tensile stresses, taking into account or disregarding the influence of spin–orbit coupling (SOC). The TDOS can assist us in comprehending the electronic band structure behavior of Ca3AsI3. When not subjected to strain, the contribution of the I atom's orbitals to the total density of states (TDOS) of the valence band below the Fermi level (EF) is considerable, while the impact of the Ca-4s, Ca-3p, and As-3d orbitals is minimal, regardless of the consideration of the SOC effect.

FIG. 10.

TDOS of Ca3AsI3 under different compressive and tensile strains (a) without and (b) with the SOC effect.

FIG. 10.

TDOS of Ca3AsI3 under different compressive and tensile strains (a) without and (b) with the SOC effect.

Close modal

Conversely, the contribution of the As atom's orbitals to the TDOS of the conduction band above EF is significant, while the impact of the Ca-4s, Ca-3p, and I-5p orbitals is negligible, regardless of the consideration of the SOC effect. The presence of a gap in the TDOS around the Fermi level signifies the semiconducting nature of the material and draws attention to its bandgap properties.31 When subjected to compressive strains ranging from −6% to 0%, the TDOS line of Ca3AsI3 shifts closer to the Fermi level. This occurs regardless of whether the SOC effect is taken into account or not, and it causes an enhancement in the material’s conductivity. Conversely, under tensile strains (0% to +6%), regardless of whether the SOC effect is taken into account or not, the TDOS line moves away from the Fermi level, causing a decrease in the conductivity of Ca3AsI3 materials. Based on a detailed examination of the band structure and TDOS of Ca3AsI3, there is a prediction of a transition occurring in the bandgap.

It is essential to research a material’s optical properties in order to comprehend how light interacts with it.37 In order to make optimal use of the Ca3AsI3 material in optoelectronic devices, it is essential to possess a comprehensive understanding of its various optical properties. These include the real and imaginary portions of its dielectric functions, as well as its optical absorption and loss. To gain a more thorough comprehension of how Ca3AsI3 performs under conditions of biaxial compression (ranging from 0% to −6%) and tension (ranging from 0% to +6%), additional calculations were conducted to determine the material’s dielectric functions, optical absorption, and optical loss in the visible light spectrum.36 The dielectric function is written in the form ε(ω), which has two parts, real [represented by ɛ1 (ω)] and imaginary [represented by ɛ2 (ω)],
(3)

The Kramers–Kronig transformation is employed to derive the real part of the dielectric function,60 while the components of the momentum matrix are used to calculate the imaginary part of the dielectric function.61 

By analyzing the real part of the dielectric constant, one can acquire knowledge about polarization and dispersion effects.62,63 A value of 7.98 is obtained for ɛ1 (0) of cubic Ca3AsI3 through calculation. In Fig. 10(a), it can be observed that the real part ɛ1 (0) of the dielectric function displays pronounced absorption peaks close to photon energies, particularly at ∼2.1 and 3.2 eV. In the Ca3AsI3 structure, the real part values of ε1(ω) follow a trend of rising from ε1(0), peaking, and then declining to form multiple peaks. The magnitude and location of these peaks are affected by variations in the biaxial strain. Typically, materials characterized by high bandgap values demonstrate lower peak values of a dielectric constant than their low bandgap counterparts.64  Figure 10(b) shows that compressed under force, the peak of the dielectric constant in the Ca3AsI3 structure was shifted to a lower photon energy (redshift) due to a reduction in the bandgap. An increase in tensile strain resulted in a lower peak in the dielectric constant of the Ca3AsI3 structure, which shifted toward a higher photon energy (blueshift).

The peaks detected in the imaginary part of the dielectric function are a result of carrier transitions occurring between the valence and conduction bands.65 The imaginary part ɛ2 (ω) of a system is closely linked to its band structure. The principal peak in ɛ2 (ω) occurs at a photon energy of 3.6 eV, as shown in Fig. 10(c), which corresponds to the highest optical point of 5.93 for cubic Ca3AsI3. The data obtained from the Ca3AsI3 compound imply that its maximum peaks fall within the visible and infrared range, likely due to its narrowness. Figure 11(d) demonstrates that the application of strain results in the enlargement and displacement of the dielectric function’s imaginary portion near the long wavelength region. Roughly speaking, the imaginary portion of the Ca3AsI3’s spectrum can be categorized into two distinct regions: a lower energy range spanning from 2.1 to 4.7 eV and a higher energy range spanning from 5.3 to 7.8 eV. The absorption peaks present in the imaginary portion of the spectrum have a significant impact on the valence-to-conduction band shift of charge carriers. The displacement of these peaks can be attributed to variations in the lattice constant and bandgap.

FIG. 11.

Real portion of the dielectric function of Ca3AsI3 under various tensile and compressive strains (a) without and (b) with strain and the imaginary portion of the dielectric function under various tensile and compressive strains (c) without and (d) with strain.

FIG. 11.

Real portion of the dielectric function of Ca3AsI3 under various tensile and compressive strains (a) without and (b) with strain and the imaginary portion of the dielectric function under various tensile and compressive strains (c) without and (d) with strain.

Close modal

A redshift and blueshift in the imaginary peaks are displayed under compressive and tensile strains, respectively, as demonstrated in Fig. 11(d). Based on our findings, the Ca3AsI3 perovskite material being studied could potentially exhibit an adjustable absorption spectral region under biaxial compressive and tensile strains. Furthermore, we observed that the imaginary dielectric component becomes zero beyond 8 eV photon energy, indicating the material’s higher optical transparency and lower optical absorption. The potential of a material to attain high solar PCE is significantly influenced by its ability to absorb light energy,34 and the optical absorption coefficient is a key parameter that supports this.36 Across all configurations, analogous features can be observed between the profile of the optical absorption coefficient and the imaginary portion of the dielectric constant.54 In Ca3AsI3, the optical absorption spectra exhibit the maximum peak in the visible zone.66 

Figures 12(a) and 12(b) display the absorption coefficient of Ca3AsI3 in terms of photon energy for both unstrained and various biaxially strained conditions. The absorption edge undergoes a considerable displacement toward the blue region of the spectrum in the presence of tensile strain and a notable shift toward the red end of the spectrum when there is compressive strain. In structures, the absorption coefficient in the visible light territory is increased by compressive strain, in comparison to the unstrained system. For systems that are subjected to tensile strain, the opposite is observed. The strain-induced alteration in the absorption of the Ca3AsI3 is in line with the anticipated bandgap. If the compressive strain is augmented, the absorption coefficient in the visible range of the Ca3AsI3 structure increases, and this property is beneficial for the development of solar cells. The Ca3AsI3 compound experiences a reduction in the absorption coefficient in the visible light territory when subjected to increased tensile strain.

FIG. 12.

Absorption coefficient of Ca3AsI3 in terms of photon energy (a) without and (b) with strain and the loss function (c) without and (d) with strain.

FIG. 12.

Absorption coefficient of Ca3AsI3 in terms of photon energy (a) without and (b) with strain and the loss function (c) without and (d) with strain.

Close modal

The light-responsive characteristics of a substance can be examined by studying the energy dissipation of an electron,34 denoted as L(ω). It determines the extent of energy loss that occurs within a medium or substance during propagation. In case the energy level of a released photon exceeds the material’s bandgap, then a peak in the plot of L(ω) for Ca3AsI3 indicates a loss of energy. Figure 12(d) illustrates that the cubic structure of Ca3AsI3 exhibits peaks in the range of 7–10 eV in the plot of L(ω), and the unstrained state is indicated in Fig. 12(c). If the energy is below the bandgap, then there is no scattering that can be detected. The loss function for photon energy in the strained Ca3AsI3 material is observed up to 10 eV. As shown in Fig. 12(d), the optical loss of Ca3AsI3 has been estimated under different biaxial compressive and tensile strain conditions. The findings suggest a redshift, which indicates a considerable shift toward the direction of lower photon energies. The application of compressive strain results in an optical loss variation in all structures. Within the visible light spectrum, the optical losses of the Ca3AsI3 structure are greater under tensile strain rather than under compressive strain.

Compressive strain 0–6 is clearly discussed in this article. Now we discuss how much strain should be applied on Ca3AsI3 to change its phase and how it is changes the behavior from semiconductor to metallic-phase. As the compressive strain increases (incorporating the SOC effect and without it), both the valence band maximum (VBM) and conduction band minimum (CBM) progressively move closer to the Fermi level, resulting in a reduction in the energy bandgap.43,67–69 When the compressive strain increases (incorporating the SOC effect and without it) from 0% to 15%, the band structure analysis reveals a notable decrease in the direct bandgap of the Ca3AsI3 perovskite material, decreasing from 1.58 to 0 eV (without SOC) and 1.27–0 eV (with SOC).

The band structure displays a metallic feature in Figs. 13(d)13(f) when the compression reaches the critical value of 15% (without SOC) and 12% (with SOC). Upon reaching the critical value of 15% and 12% of compression, the band structure exhibits a distinct metallic characteristic, illustrated in Figs. 13(d) and 13(e), providing clear evidence of a robust transition from a semiconductor to a metal phase. Similar trends have been occurred in CsSnBr3, CsSnCl3 and other perovskites.40,67,68

FIG. 13.

Electronic bands of Ca3AsI3 under compressive strains incorporating the SOC effect [(b), (d), and (f)] and without SOC [(a), (c), and (e)].

FIG. 13.

Electronic bands of Ca3AsI3 under compressive strains incorporating the SOC effect [(b), (d), and (f)] and without SOC [(a), (c), and (e)].

Close modal

Through the analysis of the density of states (DOS), researchers have gained a more comprehensive understanding of the metallic behavior of Ca3AsI3 under compressive strains incorporating the SOC effect and without SOC.40  Figure 14 illustrates the partial density of states (PDOS) of the Ca3AsI3 perovskite at various applied compressive strains incorporating the SOC effect and without SOC. At 12% (with SOC) and 15% (without SOC) strain, there is an increase in the contribution of the I-5p orbital, which becomes more prominent near both the valence and conduction bands. A material is considered metallic when there is a non-zero value of density of states (DOS) at the Fermi level.70,71 At 12% compressive strain (with SOC) and 15% compressive strain (without SOC), the TDOS exhibits a noteworthy non-zero value, indicating a semiconductor to metallic transition in the Ca3AsI3 metal halide at the specific pressure points shown in Figs. 14(d) and 14(e). The observation of a significant DOS at the Fermi level, particularly evident at 12% (with SOC) and 15% (without SOC) and showing an increasing trend with increasing compressive strain, strongly suggests a semiconductor to metallic transition in the Ca3AsI3 metal halide under high-compressive strain conditions. The comprehensive examination of the band structure and density of states in this study supports the prediction of the semiconductor to metallic transition in Ca3AsI3 under compressive strain. Therefore, it is essential to conduct future experimental investigations to precisely determine the transition pressure for the semiconductor to metal transition in Ca3AsI3.

FIG. 14.

Partial density of states (PDOS) of Ca3AsI3 under various applied compressive strains incorporating the SOC effect [(b), (d), and (f)] and without SOC [(a), (c), and (e)].

FIG. 14.

Partial density of states (PDOS) of Ca3AsI3 under various applied compressive strains incorporating the SOC effect [(b), (d), and (f)] and without SOC [(a), (c), and (e)].

Close modal

This study utilized first-principles DFT calculations to examine the electrical, optical, and structural properties of the inorganic material Ca3AsI3. The computed lattice parameter of 6.27 Å for Ca3AsI3 indicates the most favorable structure. The structure of Ca3AsI3 was found to have a direct bandgap value of 1.58 eV. Taking the SOC effect into account causes a decrease in the electronic bandgap of Ca3AsI3 to 1.27 eV. The bandgap exhibits a decrease with an increase in the level of compressive strain applied, regardless of whether the SOC effect is considered or not. Conversely, the bandgap increases as the induced tensile strain increases. According to the PDOS analysis, the electronic states responsible for the formation the valence band and conduction band in Ca3AsI3 primarily originate from the orbitals of As and I atoms rather than from those of Ca. Under a range of strain conditions spanning from −6% to +6%, the characteristics concerning light absorption and loss, along with the dielectric function of Ca3AsI3, were thoroughly examined. According to our research, the strain-driven optical characteristics of Ca3AsI3 exhibit absorption peaks (blue-shift and red-shift) close to the ultraviolet to visible spectrum at various induced stresses. In addition, applying compressive strain causes the peak of the Ca3AsI3 dielectric constant to shift to a lower photon energy (redshift). However, applying higher levels of tensile strain causes it to shift to a higher photon energy (blueshift). On the other hand, in the visible light spectrum, the optical losses of the Ca3AsI3 structure are greater under tensile strains rather than under compressive strains. The outcome of this research has the potential to enable further investigations into the application of Ca3AsI3 in photovoltaic and optoelectronic devices.

The authors are grateful to the Department of Electrical and Electronic Engineering, Begum Rokeya University, Rangpur 5404, Bangladesh, for permitting them to use the advanced energy materials and the solar cell research laboratory.

The authors have no conflicts to disclose.

Md. Ferdous Rahman: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Md. Azizur Rahman: Data curation (equal); Formal analysis (equal); Methodology (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Md. Rasidul Islam: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Avijit Ghosh: Data curation (equal); Formal analysis (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Md. Abul Bashar Shanto: Data curation (equal); Formal analysis (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Mithun Chowdhury: Data curation (equal); Formal analysis (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Md. Al Ijajul Islam: Data curation (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Md. Hafizur Rahman Data curation (equal); Formal analysis (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). M. Khalid Hossain: Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). M. A. Islam: Data curation (equal); Formal analysis (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
S. B.
Darling
and
S.
Fengqi
,
Energy Environ. Sci.
8
,
1953
(
2015
).
2.
Y.
Rong
,
Y.
Hu
,
A.
Mei
,
H.
Tan
,
M. I.
Saidaminov
,
S. I.
Seok
,
M. D.
McGehee
,
E. H.
Sargent
, and
H.
Han
,
Science
361
,
aat8235
(
2018
).
3.
M. F.
Rahman
,
M. J. A.
Habib
,
M. H.
Ali
,
M. H. K.
Rubel
,
M. R.
Islam
,
A. B.
Abu
, and
M. K.
Hossain
,
AIP Adv.
12
,
105317
(
2022
).
4.
M. F.
Rahman
,
J.
Hossain
,
A.
Kuddus
,
S.
Tabassum
,
M. H. K.
Rubel
,
M. M.
Rahman
,
Y.
Moriya
,
H.
Shirai
, and
A. B. M.
Ismail
,
J. Mater. Sci.
55
,
7715
(
2020
).
5.
M. F.
Rahman
,
N.
Mahmud
,
I.
Alam
,
M. H.
Ali
,
M. M. A.
Moon
,
A.
Kuddus
,
G. F. I.
Toki
,
M. H. K.
Rubel
,
M. A.
Al Asad
, and
M. K.
Hossain
,
AIP Adv.
13
,
045309
(
2023
).
6.
G.
Jarosz
,
R.
Marczyński
, and
R.
Signerski
,
Mater. Sci. Semicond. Process.
107
,
104812
(
2020
).
7.
H.
Shirai
,
T.
Ariyoshi
,
J.-i. Hanna
, and
I.
Shimizu
,
J. Non. Cryst. Solids
137–138
,
693
696
(
1991
).
8.
M.
Cagnoni
,
A.
Tibaldi
,
J. S.
Dolado
, and
F.
Cappelluti
,
iScience
25
,
105320
(
2022
).
9.
J. P.
Wilcoxon
,
G. A.
Samara
, and
P. N.
Provencio
,
Phys. Rev. B
60
,
2704
(
1999
).
10.
A.
Kojima
,
K.
Teshima
,
Y.
Shirai
, and
T.
Miyasaka
,
J. Am. Chem. Soc.
131
,
6050
(
2009
).
11.
T. W.
Kelley
,
P. F.
Baude
,
C.
Gerlach
,
D. E.
Ender
,
D.
Muyres
,
M. A.
Haase
,
D. E.
Vogel
, and
S. D.
Theiss
,
Chem. Mater.
16
,
4413
(
2004
).
12.
B. A.
Rosales
,
M. P.
Hanrahan
,
B. W.
Boote
,
A. J.
Rossini
,
E. A.
Smith
, and
J.
Vela
,
ACS Energy Lett.
2
,
906
(
2017
).
13.
Y.-Y.
Tang
,
P.-F.
Li
,
W.-Q.
Liao
,
P.-P.
Shi
,
Y.-M.
You
, and
R.-G.
Xiong
,
J. Am. Chem. Soc.
140
,
8051
(
2018
).
14.
G.
Xing
,
N.
Mathews
,
S.
Sun
,
S. S.
Lim
,
Y. M.
Lam
,
M.
Grätzel
,
S.
Mhaisalkar
, and
T. C.
Sum
,
Science
342
(
6156
),
344
347
(
2013
).
15.
V.
Chellappan
and
S.
Ramakrishna
,
Sol. Energy
122
,
678
(
2015
).
16.
T. R.
Cook
,
D. K.
Dogutan
,
S. Y.
Reece
,
Y.
Surendranath
,
T. S.
Teets
, and
D. G.
Nocera
,
Chem. Rev.
110
,
6474
(
2010
).
17.
A.
Richter
,
R.
Müller
,
J.
Benick
,
F.
Feldmann
,
B.
Steinhauser
,
C.
Reichel
,
A.
Fell
,
M.
Bivour
,
M.
Hermle
, and
S. W.
Glunz
,
Nat. Energy
6
,
429
(
2021
).
18.
See https://www.nrel.gov/pv/cell-efficiency.html for “Best research-cell efficiency chart”.
19.
M. K.
Hossain
,
M. H. K.
Rubel
,
G. F. I.
Toki
,
I.
Alam
,
M. F.
Rahman
, and
H.
Bencherif
,
ACS Omega
7
,
43210
(
2022
).
20.
M. K.
Hossain
,
A. A.
Arnab
,
R. C.
Das
,
K. M.
Hossain
,
M. H. K.
Rubel
,
M. F.
Rahman
,
H.
Bencherif
,
M. E.
Emetere
,
M. K. A.
Mohammed
, and
R.
Pandey
,
RSC Adv.
12
,
34850
(
2022
).
21.
J.-P.
Correa-Baena
,
M.
Saliba
,
T.
Buonassisi
,
M.
Grätzel
,
A.
Abate
,
W.
Tress
, and
A.
Hagfeldt
,
Science
358
(
6364
),
739
744
(
2017
).
22.
M. K.
Hossain
,
D. P.
Samajdar
,
R. C.
Das
,
A. A.
Arnab
,
M. F.
Rahman
,
M. H. K.
Rubel
,
M. R.
Islam
,
H.
Bencherif
,
R.
Pandey
,
J.
Madan
, and
M. K. A.
Mohammed
,
Energy Fuels
37
,
3957
(
2023
).
23.
M. K.
Hossain
,
G. F. I.
Toki
,
I.
Alam
,
R.
Pandey
,
D. P.
Samajdar
,
M. F.
Rahman
,
M. R.
Islam
,
M. H. K.
Rubel
,
H.
Bencherif
,
J.
Madan
, and
M. K. A.
Mohammed
,
New J. Chem.
47
,
4801
(
2023
).
24.
M. K.
Islam
,
G. F.
Ishraque Toki
,
D. P.
Samajdar
,
M. H. K.
Rubel
,
M.
Mushtaq
,
Md. R.
Islam
,
Md. F.
Rahman
,
S.
Bhattarai
,
H.
Bencherif
,
M. K. A.
Mohammed
,
R.
Pandey
, and
J.
Madan
,
Energy Fuels
37
(
10
),
7380
7400
(
2023
).
25.
M. K.
Hossain
,
S.
Bhattarai
,
A. A.
Arnab
,
M. K. A.
Mohammed
,
R.
Pandey
,
H.
Ali
,
F.
Rahman
,
R.
Islam
,
D. P.
Samajdar
,
J.
Madan
,
H.
Bencherif
,
D. K.
Dwivedi
, and
M.
Amami
,
RSC Adv.
13
,
21044
(
2023
).
26.
S.
Bonomi
,
I.
Tredici
,
B.
Albini
,
P.
Galinetto
,
A.
Rizzo
,
A.
Listorti
,
U. A.
Tamburini
, and
L.
Malavasi
,
Chem. Commun.
54
,
13212
(
2018
).
27.
G.
Liu
,
L.
Kong
,
P.
Guo
,
C. C.
Stoumpos
,
Q.
Hu
,
Z.
Liu
,
Z.
Cai
,
D. J.
Gosztola
,
H. K.
Mao
,
M. G.
Kanatzidis
, and
R. D.
Schaller
,
ACS Energy Lett.
2
,
2518
(
2017
).
28.
P.
Cheng
,
P.
Wang
,
Z.
Xu
,
X.
Jia
,
Q.
Wei
,
N.
Yuan
,
J.
Ding
,
R.
Li
,
G.
Zhao
,
Y.
Cheng
,
K.
Zhao
, and
S. F.
Liu
,
ACS Energy Lett.
4
,
1830
(
2019
).
29.
S.
Jiang
,
Y.
Fang
,
R.
Li
,
H.
Xiao
,
J.
Crowley
,
C.
Wang
,
T. J.
White
,
W. A.
Goddard
,
Z.
Wang
,
T.
Baikie
, and
J.
Fang
,
Angew. Chem.
128
,
6650
(
2016
).
30.
A.
Francisco-López
,
B.
Charles
,
O. J.
Weber
,
M. I.
Alonso
,
M.
Garriga
,
M.
Campoy-Quiles
,
M. T.
Weller
, and
A. R.
Goñi
,
J. Phys. Chem. C
122
,
22073
(
2018
).
31.
M. R.
Islam
,
M. R. H.
Mojumder
,
R.
Moshwan
,
A. S. M. J.
Islam
,
M. A.
Islam
,
M. S.
Rahman
, and
M. H.
Kabir
,
ECS J. Solid State Sci. Technol.
11
,
033001
(
2022
).
32.
D.
Liu
,
Q.
Li
,
H.
Jing
, and
K.
Wu
,
RSC Adv.
9
,
3279
(
2019
).
33.
L.
Wang
,
K.
Wang
, and
B.
Zou
,
J. Phys. Chem. Lett.
7
,
2556
(
2016
).
34.
T.
Ou
,
J.
Yan
,
C.
Xiao
,
W.
Shen
,
C.
Liu
,
X.
Liu
,
Y.
Han
,
Y.
Ma
, and
C.
Gao
,
Nanoscale
8
,
11426
(
2016
).
35.
L.
Wang
,
K.
Wang
,
G.
Xiao
,
Q.
Zeng
, and
B.
Zou
,
J. Phys. Chem. Lett.
7
,
5273
(
2016
).
36.
R.
Islam
,
K.
Liu
,
Z.
Wang
,
S.
Hasan
,
Y.
Wu
,
S.
Qu
, and
Z.
Wang
,
Mater. Today Commun.
31
,
103305
(
2022
).
37.
M. R.
Islam
,
A. J.
Islam
,
K.
Liu
,
Z.
Wang
,
S.
Qu
,
C.
Zhao
,
X.
Wang
, and
Z.
Wang
,
Physica B
638
,
413960
(
2022
).
38.
D. K.
Seo
,
N.
Gupta
,
M. H.
Whangbo
,
H.
Hillebrecht
, and
G.
Thiele
,
Inorg. Chem.
37
,
407
(
1998
).
39.
U.
Schwarz
,
F.
Wagner
,
K.
Syassen
, and
H.
Hillebrecht
,
Phys. Rev. B
53
,
12545
(
1996
).
40.
J.
Islam
and
A. K. M. A.
Hossain
,
Sci. Rep.
10
,
14391
(
2020
).
41.
A.
Jaffe
,
Y.
Lin
,
W. L.
Mao
, and
H. I.
Karunadasa
,
J. Am. Chem. Soc.
139
,
4330
(
2017
).
42.
D.
Liu
,
Q.
Li
,
H.
Jing
, and
K.
Wu
,
RSC Adv.
9
(
6
),
3279
3284
(
2019
).
43.
R.
Islam
,
S.
Islam
,
N.
Ferdous
,
K. N.
Anindya
, and
A.
Hashimoto
,
J. Comput. Electron.
18
,
407
(
2019
).
44.
M.
Pandey
,
K. W.
Jacobsen
, and
K. S.
Thygesen
,
J. Phys. Chem. Lett.
7
,
4346
(
2016
).
45.
R. J.
Elliott
,
Phys. Rev.
96
(
2
),
266
279
(
1954
).
46.
P. C.
Chow
and
L.
Liu
,
Phys. Rev.
140
(
5
A),
A1817
(
1965
).
47.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A.
Dal Corso
,
S.
de Gironcoli
,
S.
Fabris
,
G.
Fratesi
,
R.
Gebauer
,
U.
Gerstmann
,
C.
Gougoussis
,
A.
Kokalj
,
M.
Lazzeri
,
L.
Martin-Samos
,
N.
Marzari
,
F.
Mauri
,
R.
Mazzarello
,
S.
Paolini
,
A.
Pasquarello
,
L.
Paulatto
,
C.
Sbraccia
,
S.
Scandolo
,
G.
Sclauzero
,
A. P.
Seitsonen
,
A.
Smogunov
,
P.
Umari
, and
R. M.
Wentzcovitch
,
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
48.
J. P.
Perdew
and
A.
Zunger
,
Phys. Rev. B
23
,
5048
(
1981
).
49.
J. M.
Smith
,
S. P.
Jones
, and
L. D.
White
,
Gastroenterology
72
,
193
(
1977
).
50.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
51.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M.
Buongiorno Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
,
N.
Colonna
,
I.
Carnimeo
,
A.
Dal Corso
,
S.
de Gironcoli
,
P.
Delugas
,
R. A.
DiStasio
,
A.
Ferretti
,
A.
Floris
,
G.
Fratesi
,
G.
Fugallo
,
R.
Gebauer
,
U.
Gerstmann
,
F.
Giustino
,
T.
Gorni
,
J.
Jia
,
M.
Kawamura
,
H.-Y.
Ko
,
A.
Kokalj
,
E.
Küçükbenli
,
M.
Lazzeri
,
M.
Marsili
,
N.
Marzari
,
F.
Mauri
,
N. L.
Nguyen
,
H.-V.
Nguyen
,
A.
Otero-de-la-Roza
,
L.
Paulatto
,
S.
Poncé
,
D.
Rocca
,
R.
Sabatini
,
B.
Santra
,
M.
Schlipf
,
A. P.
Seitsonen
,
A.
Smogunov
,
I.
Timrov
,
T.
Thonhauser
,
P.
Umari
,
N.
Vast
,
X.
Wu
, and
S.
Baroni
,
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
52.
H. J.
Feng
and
Q.
Zhang
,
Appl. Phys. Lett.
118
,
111902
(
2021
).
53.
M. K.
Hossain
,
G. F. I.
Toki
,
A.
Kuddus
,
M. H. K.
Rubel
,
M. M.
Hossain
,
H.
Bencherif
,
M. F.
Rahman
,
M. R.
Islam
, and
M.
Mushtaq
,
Sci. Rep.
13
,
2521
(
2023
).
54.
M. H. K.
Rubel
,
M. A.
Hossain
,
M. K.
Hossain
,
K. M.
Hossain
,
A. A.
Khatun
,
M. M.
Rahaman
,
M.
Ferdous Rahman
,
M. M.
Hossain
, and
J.
Hossain
,
Results Phys.
42
,
105977
(
2022
).
55.
Y.
Saeed
,
B.
Amin
,
H.
Khalil
,
F.
Rehman
,
H.
Ali
,
M. I.
Khan
,
A.
Mahmood
, and
M.
Shafiq
,
RSC Adv.
10
(
34
),
20196
(
2020
).
56.
N. A.
Noor
,
M. W.
Iqbal
,
T.
Zelai
,
A.
Mahmood
,
H. M.
Shaikh
,
S. M.
Ramay
, and
W.
Al-Masry
,
J. Mater. Res. Technol.
13
,
2491
(
2021
).
57.
G.
Volonakis
,
A. A.
Haghighirad
,
R. L.
Milot
,
W. H.
Sio
,
M. R.
Filip
,
B.
Wenger
,
M. B.
Johnston
,
L. M.
Herz
,
H. J.
Snaith
, and
F.
Giustino
,
J. Phys. Chem. Lett.
8
,
772
(
2016
).
58.
S.
Nair
,
M.
Deshpande
,
V.
Shah
,
S.
Ghaisas
, and
S.
Jadkar
,
J. Phys.: Condens. Matter
31
,
445902
(
2019
).
59.
C. C.
Liu
,
H.
Jiang
, and
Y.
Yao
,
Phys. Rev. B
84
,
195430
(
2011
).
60.
B.
Ehrler
,
E.
Alarcón-Lladó
,
S. W.
Tabernig
,
T.
Veeken
,
E. C.
Garnett
, and
A.
Polman
,
ACS Energy Lett.
5
,
3029
(
2020
).
61.
K. E.
Peiponen
,
V.
Lucarini
,
E. M.
Vartiainen
, and
J. J.
Saarinen
,
Eur. Phys. J. B
41
,
61
(
2004
).
62.
Z.
Xu
,
Solid State Commun.
76
(
9
),
1143
1147
(
1990
). 00381098
63.
K.
Shportko
,
T.
Barlas
,
E.
Venger
,
H.
El-Nasser
, and
V.
Ponomarenko
,
Curr. Appl. Phys.
16
,
8
(
2016
).
64.
S. F.
Maehrlein
,
P. P.
Joshi
,
L.
Huber
,
F.
Wang
,
M.
Cherasse
,
Y.
Liu
,
D. M.
Juraschek
,
E.
Mosconi
,
D.
Meggiolaro
,
F.
De Angelis
, and
X. Y.
Zhu
,
Proc. Natl. Acad. Sci. U. S. A.
118
,
e2022268118
(
2021
).
65.
M. A.
Iqbal
,
A.
Ahmad
,
M.
Malik
,
J. R.
Choi
, and
P. V.
Pham
,
Materials
15
(
7
),
2617
(
2002
).
66.
L.
Yue
,
B.
Yan
,
M.
Attridge
, and
Z.
Wang
,
Sol. Energy
124
,
143
(
2016
).
67.
S.
Hossain
,
M. H.
Babu
,
T.
Saha
,
S.
Hossain
,
J.
Podder
,
S.
Rana
,
A.
Barik
, and
P.
Rani
,
AIP Adv.
11
,
055024
(
2021
).
68.
M.
Ti
,
A.
Si
,
H.
Zhong
,
W.
Xiong
,
P.
Lv
,
J.
Yu
, and
S.
Yuan
,
Phys. Rev. B
103
,
085124
(
2021
).
69.
Md. N.
Islam
and
J.
Podder
,
Heliyon
8
(
8
),
e10032
(
2022
).
70.
E.
Haque
and
M. A.
Hossain
,
Mater. Sci. Semicond. Process.
83
,
192
(
2018
).
71.
Y.
Zhao
,
L.
Qi
,
Y.
Jin
,
K.
Wang
,
J.
Tian
, and
P.
Han
,
J. Alloys Compd.
647
,
1104
(
2015
).