Open Access Article
Imtiaz Ahamed Apon
a,
Karim KRIAAb,
Md. Azizur Rahman
*c,
Md. Alamgir Hossaind,
Chemseddine Maatkib,
Amine Aymen Assadib and
Noureddine Elboughdiri
e
aDepartment of Electrical and Electronic Engineering, Bangladesh Army University of Science and Technology (BAUST), Saidpur-5311, Bangladesh
bImam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11432, Saudi Arabia
cDepartment of Electrical and Electronic Engineering, Begum Rokeya University, Rangpur, 5400, Bangladesh. E-mail: azizurrahmanatik49@gmail.com
dDepartment of Physics, Khulna University of Engineering & Technology (KUET), Khulna-9203, Bangladesh
eChemical Engineering Department, College of Engineering, University of Ha'il, P.O. Box 2440, Ha'il 81441, Saudi Arabia
First published on 22nd October 2025
Developing stable and efficient perovskite-inspired materials has become a key focus in the pursuit of next-generation solar energy technologies, with recent advances in material design highlighting the potential of novel halide structures as sustainable alternatives to conventional silicon-based solar absorbers. This study presents a comprehensive first-principles investigation of novel Rb-based double perovskites, Rb2BX6 (B = Sn/Pb; X = Cl/Br), highlighting their potential for photovoltaic applications. All compounds exhibit negative formation enthalpies, indicating thermodynamic stability, and tolerance factors around 0.80 confirm structural feasibility. Mechanical stability is validated through Born criteria and elastic constants. Band structure calculations reveal direct band gaps of 2.646 eV (Rb2SnCl6), 1.451 eV (Rb2SnBr6), 1.379 eV (Rb2PbCl6), and 0.357 eV (Rb2PbBr6), suggesting semiconducting behavior, with Rb2SnBr6 and Rb2PbCl6 falling within the optimal range for visible-light absorption. Partial density of states (PDOS) analyses show that the valence bands are mainly composed of halide p-orbitals, while conduction bands are dominated by B-site s- and p-orbitals. ELATE tensor analysis reveals moderate elastic anisotropy, with anisotropy indices of 0.22 for Rb2SnBr6 and 0.18 for Rb2PbCl6. Optical studies indicate absorption coefficients exceeding 105 cm−1 in the visible region for both materials. Mulliken population analysis confirms strong ionic bonding and moderate charge transfer between atoms. Moreover, Rb2SnCl6 and Rb2PbCl6 are dynamically stable, whereas Rb2SnBr6 and Rb2PbBr6 exhibit dynamic instability. SCAPS-1D device simulations yield a power conversion efficiency (PCE) of 20.44% for Rb2SnBr6, accompanied by a short-circuit current density (JSC) of 22.3 mA cm−2, an open-circuit voltage (VOC) of 1.01 V, and a fill factor (FF) of 89.7%. For Rb2PbCl6, a slightly higher PCE of 21.84% is achieved (JSC of 23.1 mA cm−2, VOC of 1.05 V, FF of 90.1%). Although Rb2PbCl6 demonstrates superior efficiency, its toxicity due to lead content poses environmental concerns. In contrast, Rb2SnBr6 offers a highly efficient, non-toxic alternative, positioning it as a viable candidate for eco-friendly and sustainable photovoltaic devices.
The renewed interest in A2BX6 perovskites arose after the success of lead halide perovskite solar cells post-2009.4,5 Among them, rubidium-based double perovskites (Rb2BX6) exhibit structural robustness, optimal band gaps, and superior electronic properties, making them promising candidates for optoelectronic and photovoltaic applications.6–8 Tin (Sn)-based and lead (Pb)-based double halide perovskites, particularly Rb2SnX6 (X = Cl, Br), have garnered increasing attention for their tunable band gaps, strong visible-light absorption, and enhanced stability.9,10 Pb-halide perovskites have emerged as one of the most promising classes of semiconductors, primarily because of their band gap tunability, efficient charge transport, and outstanding photoluminescence quantum yield (PLQY), which together underpin their success in photovoltaic and light-emitting applications.11–13 However, their environmental concerns have shifted research toward Sn-based alternatives.14 Density Functional Theory (DFT) calculations and experimental studies have validated the thermodynamic stability and suitable band gaps (0.9 to 0.6 eV) of Sn-based double perovskites, confirming their potential for high-performance solar cells.15–17
Recent studies have further highlighted the promise of Sn-based perovskites. Faizan et al. investigated A2BX6 (A = Rb, Cs; B = Sn, Pd, Pt = Cl, Br, I) using DFT, revealing strong dielectric constants and superior light absorption, critical for efficient energy conversion.18 Karim et al. synthesized and characterized Cs2SnX6 (X = Cl, Br, I), reporting optical band gaps ranging from 4.89 eV (Cs2SnCl6) to 1.35 eV (Cs2SnI6), with mixed halide systems exhibiting non-linear optical behavior due to symmetry distortions.19 Vamsi Krishna et al. conducted DFT studies on A2BX6 (A = Cs; B = Sn; X = Cl, Br, I) perovskites, confirming their strong potential for single- and multi-junction solar cells.20 The synergy between theoretical and experimental approaches is crucial for advancing the optoelectronic, thermoelectric, and photocatalytic applications of these materials. Double perovskites are commonly synthesized using solution-based methods or solid-state reactions, with experimental studies confirming wide band gaps, such as 2.97 eV for Rb2SnBr6.21 There are many other studies on A2BX6 compounds that have previously demonstrated promising efficiencies for photovoltaic applications. Such as according to K. Bouferrache et al., the lattice constants of Cs2MCl6 (M = Se, Sn, Te, Ti) agree with experiments within 1.3 to 3%. Cs2SnCl6 is isotropic, while the others are anisotropic, especially Cs2TiCl6. Their low elastic moduli indicate they are easily deformable.22 Rifat et al. reported that K2CeCl6-based solar cells achieved 17.22% PCE, 1.02 VOC, 22.5 mA cm−2 JSC, and ∼81% FF.23 Md. S. H. Saikot et al. reported that Na2PdCl6-based solar cells achieved 25.55% PCE (JSC 42.55 mA cm−2, VOC 0.758 V, FF 79.16%), outperforming Li2PdCl6 with 23.06% PCE (JSC 38.12 mA cm−2, VOC 0.786 V, FF 76.97%).24
This study employs first-principles calculations to systematically investigate the structural, electronic, optical, mechanical, phonon, and charge population properties of Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6. The effects of absorber layer thickness, defect density, and other key parameters on overall solar cell performance are also evaluated through device-level simulations. By examining the influence of cation substitution (Sn vs. Pb) and halide variation (Cl vs. Br), detailed insights are provided into their suitability for optoelectronic and energy-harvesting applications. This work provides a comprehensive theoretical framework linking fundamental material properties to device-level performance, paving the way for the design of efficient and environmentally sustainable perovskite solar absorbers.
m (#225).39,40 Fig. 1a and b illustrates the primitive crystalline unit cell of Rb2SnX6 and Rb2PbX6 double perovskite materials.
![]() | ||
| Fig. 1 Primitive crystalline unit cell of (a) Rb2SnX6 and (b) Rb2PbX6 and schematic design of the solar cell architectures for (c) Rb2SnBr6 and (d) Rb2PbCl6 double perovskite compounds. | ||
In these structures, the Rb atom occupies the 8c Wyckoff position with fractional coordinates (0.25, 0.25, 0.25), while the Sn/Pb atom is positioned at the 4a Wyckoff site with coordinates (0, 0, 0). The X atom is located at the 24e Wyckoff site with fractional coordinates (0.244, 0, 0).41,42 Additionally, Fig. 1c and d presents the schematic designs of solar cell architectures based on Rb2SnBr6 and Rb2PbCl6, highlighting their potential use as absorber layers in photovoltaic applications. This structural analysis provides a foundation for further exploration of their electronic, mechanical, and optical properties.
The Birch–Murnaghan equation of state (see eqn (1))43 was applied to obtain the optimized volume and lattice parameters of the studied perovskite compounds.
![]() | (1) |
To enhance the reliability of our predictions, four exchange–correlation functionals—GGA-PBE, GGA-PBEsol, m-GGA, and HSE06—were employed. GGA-PBE served as the baseline due to its wide adoption in structural and electronic studies, while PBEsol was included for its improved accuracy in describing equilibrium lattice parameters.44 The meta-GGA functional was considered for its refined density-gradient corrections, and HSE06 was applied selectively to validate key electronic and structural properties.45 The calculated lattice constants (Å) for Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6 are 7.581, 7.933, 7.697, and 8.130 with GGA-PBE; 7.320, 7.605, 7.697, and 7.437 with GGA-PBEsol; 7.581, 7.932, 7.697, and 8.129 with HSE06; and 7.581, 7.932, 7.697, and 7.678 with m-GGA. The corresponding unit cell volumes (Å3) are 308.174, 353.036, 322.443, and 354.245 with GGA-PBE; 275.548, 308.369, 293.009, and 326.250 with GGA-PBEsol; 274.806, 351.806, 291.031, and 381.810 with m-GGA; and 308.145, 352.497, 322.462, and 354.285 with HSE06. The observed variation in lattice constants and volumes reflects the intrinsic characteristics of the different functionals. GGA-PBE tends to overestimate lattice dimensions due to its underbinding nature, whereas PBEsol systematically underestimates them. The m-GGA functional generally yields intermediate values but can show larger fluctuations depending on bonding environments.46 HSE06, which incorporates a fraction of exact exchange, provides improved electronic accuracy while maintaining structural parameters close to PBE. These systematic differences are consistent with previous benchmark studies and confirm that, despite numerical variations, the overall structural trends remain robust. These values are in good agreement with previous theoretical and experimental reports. For instance, the lattice constant of Rb2SnBr6 was reported as 7.932 Å using the HSE06 hybrid functional, which is almost identical to our value, while a slightly larger lattice constant of 10.123 Å was observed experimentally.47 Similarly, the theoretical lattice volume for Rb2SnBr6 was associated with a lattice constant consistent with our findings.48 For comparison, other related halide perovskites such as Na2SnCl6 and Na2SnBr6 exhibit significantly larger lattice constants of 10.67 Å and 10.88 Å, respectively.49 Likewise, K2SnCl6, K2SnBr6, and K2SnI6 display lattice constants of 9.99 Å, 10.48 Å, and 1.90 Å, respectively.50 Furthermore, theoretical studies reported lattice constants of Cs2SnCl6, Cs2SnBr6, and Cs2SnI6 as 3.90 Å, 2.70 Å, and 1.26 Å, respectively.51 Overall, our calculated lattice parameters for Rb-based compounds are consistent with previous works, while small variations arise due to different computational methods and experimental conditions. The thermodynamic stability of Rb2BX6 (B = Sn, Pb; X = Cl, Br) double perovskites was evaluated using formation enthalpy (ΔHf) calculations. Using the relation eqn (2),52
| ΔHf = Etotal (Rb2BX6) − (2ERb + EB + 6Ex) | (2) |
All four compounds were found to be stable with negative formation energies. Specifically, Rb2SnCl6 exhibits ΔHf = −3.41123 eV per atom, Rb2SnBr6 ΔHf = −3.01976 eV per atom, Rb2PbCl6 ΔHf = −3.30123 eV per atom, and Rb2PbBr6 ΔHf = −2.93514 eV per atom. These results indicate that all compounds are thermodynamically favorable,53 with Cl-based perovskites slightly more stable than their Br analogues and Pb-based compounds showing comparable stability to Sn-based ones Table 1. This stability suggests that these Rb2BX6 materials are promising candidates for further exploration in lead-free photovoltaic applications.
| References | Compounds | Band gap, eV | Lattice constants (Å) | Density, g cm−3 | Unit cell volume, V (Å3) | Formation enthalpy, ΔHf (eV per atom) | Function/method |
|---|---|---|---|---|---|---|---|
| This work | Rb2SnCl6 | 2.646 | 7.581 | 2.707 | 308.174 | −3.411 | GGA-PBE |
| Rb2SnBr6 | 1.451 | 7.933 | 3.617 | 353.036 | −3.019 | ||
| Rb2PbCl6 | 1.379 | 7.697 | 3.042 | 322.443 | −3.301 | ||
| Rb2PbBr6 | 0.357 | 8.130 | 4.019 | 354.245 | −2.935 | ||
| Rb2SnCl6 | 4.231 | 7.581 | 2.727 | 308.145 | — | Hybride-HSE06 | |
| Rb2SnBr6 | 2.828 | 7.932 | 3.724 | 352.497 | — | ||
| Rb2PbCl6 | 1.946 | 7.697 | 3.134 | 322.462 | — | ||
| Rb2PbBr6 | 1.742 | 8.129 | 4.136 | 354.285 | — | ||
| Rb2SnCl6 | 2.572 | 7.320 | — | 275.548 | — | GGA-PBESOL | |
| Rb2SnBr6 | 1.178 | 7.605 | — | 308.369 | — | ||
| Rb2PbCl6 | 1.396 | 7.697 | — | 293.009 | — | ||
| Rb2PbBr6 | 0.277 | 7.437 | — | 326.250 | — | ||
| Rb2SnCl6 | 2.982 | 7.581 | — | 274.806 | — | m-GGA | |
| Rb2SnBr6 | 1.961 | 7.932 | — | 351.806 | — | ||
| Rb2PbCl6 | 1.631 | 7.697 | — | 291.031 | — | ||
| Rb2PbBr6 | 0.946 | 7.678 | — | 381.810 | — | ||
| 49 | Na2SnCl6 | 2.77 | 10.67 | — | — | −3.203 | GGA-PBE |
| Na2SnBr6 | 1.12 | 10.88 | — | — | −3.119 | ||
| Na2SnCl6 | 3.98 | — | — | — | — | Hybride-HSE06 | |
| Na2SnBr6 | 3.28 | — | — | — | — | ||
| 50 | K2SnCl6 | 4.54 | 9.10 | — | — | — | Experimental |
| K2SnBr6 | 3.27 | 10.48 | — | — | — | ||
| K2SnI6 | 1.90 | — | — | — | — | ||
| 48 | Rb2SnBr6 | 1.28 | — | — | 333.270 | — | Theoretical |
| 47 | Rb2SnBr6 | 4.790 | 10.123 | 3.22 | — | — | Experimental |
| 51 | Cs2SnCl6 | 3.90 | — | — | — | — | Theoretical |
| Cs2SnBr6 | 2.70 | — | — | — | — | ||
| Cs2SnI6 | 1.26 | — | — | — | — |
From the present calculations, the obtained unit cell volumes for Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6 are 308.174 Å3, 353.036 Å3, 322.443 Å3, and 354.245 Å3, respectively, with corresponding densities of 2.707 g cm−3, 3.617 g cm−3, 3.042 g cm−3, and 4.019 g cm−3. The substitution of Cl with the larger Br anion increases the cell volume, as observed in both Sn- and Pb-based compounds (Rb2SnCl6 → Rb2SnBr6 and Rb2PbCl6 → Rb2PbBr6) Table 2. Conversely, the density shows a systematic increase with the incorporation of the heavier Br atom, reflecting the strong dependence of density on atomic mass. Furthermore, comparing Sn- and Pb-based systems reveals that Pb substitution results in slightly larger volumes and higher densities than their Sn counterparts, consistent with the heavier atomic weight of Pb. These observations confirm that both the type of halogen (Cl vs. Br) and the choice of B-site cation (Sn vs. Pb) significantly influence the structural compactness and mass density of the perovskites.47–51 The tolerance factor (t) is crucial in perovskite research because it helps predict the structural stability and symmetry of the crystal. The relation of eqn (3) has computed the tolerance factor (t).54
![]() | (3) |
| Materials | rRb | rB | rX | Tolerance factor (t) |
|---|---|---|---|---|
| Rb2SnCl6 | 1.61 | 0.69 | 1.81 | 0.9672 |
| Rb2SnBr6 | 1.61 | 0.69 | 1.96 | 0.9526 |
| Rb2PbCl6 | 1.61 | 1.19 | 1.81 | 0.8060 |
| Rb2PbBr6 | 1.61 | 1.19 | 1.96 | 0.8014 |
For Rb2BX6 perovskites, the tolerance factor ranges from 0.8014 to 0.9672, indicating that all the studied materials are structurally stable. Among them, Rb2SnCl6 shows the highest t value (0.9672), representing the most well-balanced and stable structure, while Rb2SnBr6 (0.9526) exhibits a slight decrease due to the larger Br− ion. Rb2PbCl6 (0.806) and Rb2PbBr6 (0.8014) have lower tolerance factors, suggesting minor lattice distortions caused by the larger Pb2+ ion. Overall, while all compounds are stable, the B-site cation and halide size influence subtle variations in structural stability and lattice ordering, consistent with trends observed in rubidium-based perovskites.
| C11 > 0, 4C44 > 0,C11 − C12 > 0 and C11 + 2C12 > 0 | (4) |
| Mechanical values | This work | |||
|---|---|---|---|---|
| Rb2SnCl6 | Rb2SnBr6 | Rb2PbCl6 | Rb2PbBr6 | |
| C11 (GPa) | 12.999 | 32.004 | 10.901 | 5.425 |
| C12 (GPa) | 3.841 | 23.902 | 3.843 | 4.530 |
| C44 (GPa) | 3.873 | 8.084 | 6.039 | 4.364 |
| Kleinman parameter, ζ | 0.443 | 0.821 | 0.704 | 0.885 |
| Cauchy pressure, CP | −0.032 | 15.818 | −2.196 | 0.166 |
| Bulk modulus (B) | 6.894 | 26.603 | 6.196 | 4.828 |
| Shear modulus (G) | 4.141 | 6.126 | 4.868 | 1.883 |
| Young's modulus (Y) | 10.351 | 17.069 | 11.574 | 5.001 |
| Poisson's ratio (ν) | 0.249 | 0.393 | 0.188 | 0.327 |
| Pugh's modulus ratio (B/G) | 1.664 | 4.342 | 1.272 | 2.563 |
| Machinability index, μM | 1.780 | 3.290 | 1.025 | 1.106 |
Table 3 presents a detailed analysis of the Kleinman parameter (ζ) (see eqn (5)), which indicates the relative preference of a material for bond bending over bond stretching, thereby influencing its mechanical flexibility.59
![]() | (5) |
Among the studied compounds, Rb2PbBr6 exhibits the highest ζ value (0.885); however, the toxicity associated with lead limits its practical relevance. Notably, Rb2SnBr6 shows the second-highest ζ value (0.821), making it a more favorable candidate. The greater bond-bending adaptability of Rb2SnBr6 suggests enhanced mechanical resilience, which is advantageous for potential optoelectronic and photovoltaic applications.
The Cauchy pressure (CP) is a key parameter for assessing mechanical stability (see eqn (6)),60 distinguishing between ductile and brittle behavior.61
| Cp = C12 − C44 | (6) |
A positive CP indicates ductility (greater plastic deformation), while a negative CP suggests brittleness (higher fracture tendency).62 The CP values for Rb2BX6 materials reveal distinct mechanical characteristics. Rb2SnCl6 (−0.032 GPa) and Rb2PbCl6 (−2.196 GPa) exhibit negative CP, indicating a brittle nature due to strong covalent bonding. In contrast, Rb2SnBr6 (15.818 GPa) shows high ductility, making it suitable for flexible optoelectronic applications. Rb2PbBr6 (0.166 GPa), with a slightly positive CP, falls between brittle and ductile, offering moderate mechanical adaptability. The bulk modulus (B) quantifies a material's resistance to uniform compression (see eqn (7)).60
![]() | (7) |
Rb2SnBr6 exhibits the highest B (26.603 GPa), making it the least compressible, while Rb2PbBr6 (4.828 GPa) is the most compressible. The shear modulus (G) measures resistance to shape deformation under shear stress (see eqn (8)),60 with Rb2SnBr6 (6.126 GPa) showing the highest value, indicating superior structural rigidity.
![]() | (8) |
Similarly, Young's modulus (Y) quantifies stiffness in tension, where a higher Y signifies greater rigidity (see eqn (9)).60
![]() | (9) |
Rb2SnBr6, with the highest Y (17.069 GPa), confirms its superior mechanical strength compared to the other perovskites. The Poisson's ratio (ν) and Pugh's ratio (B/G) distinguish ductility variations (see eqn (10)).63
![]() | (10) |
Materials are classified as brittle (B/G < 1.75, ν < 0.26) or ductile (B/G > 1.75, ν > 0.26).64 A higher ν indicates greater ductility, as seen in Rb2SnBr6 (ν = 0.393), while lower values, like in Rb2PbCl6 (ν = 0.188), suggest brittleness. Similarly, Pugh's ratio reflects mechanical behavior; Rb2PbBr6 (B/G = 2.563) is the most ductile, whereas Rb2SnBr6 (B/G = 0.708) is the most brittle. Thus, Rb2SnBr6 exhibits superior ductility and mechanical stability. The machinability index (μm), defined in eqn (11), indicates the ease of processing of a material. Higher μm values correspond to better machinability, reflecting lower resistance to cutting, shaping, or fabrication.65
![]() | (11) |
Among the Rb2BX6 compounds, Rb2SnBr6 has the highest μm (3.290), indicating good ductility, while Rb2SnCl6 is moderately machinable (1.780). The Pb-based compounds, Rb2PbCl6 (1.025) and Rb2PbBr6 (1.106), are less machinable due to brittleness. Overall, Sn-based Rb2BX6 compounds are easier to process than Pb-based analogs, showing that B-site substitution strongly affects mechanical workability.
The electronic band gaps of Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6 were calculated using four exchange–correlation functionals: GGA-PBE, GGA-PBEsol, m-GGA, and hybrid HSE06. Using GGA-PBE, the band gaps are 2.646, 1.451, 1.379, and 0.357 eV, respectively (Fig. 2a–d). The GGA-PBEsol functional yields slightly smaller gaps (2.572, 1.391, 1.178, and 0.277 eV, Fig. 3a), reflecting its systematic tendency to slightly compress lattice parameters, which reduces the band gap. m-GGA (meta-GGA) predicts intermediate values (2.980, 1.961, 1.631, and 0.946 eV, Fig. 3b), capturing enhanced gradient corrections that moderately increase the gap compared to PBEsol.
![]() | ||
| Fig. 2 Energy band gap values of (a) Rb2SnCl6, (b) Rb2SnBr6, (c) Rb2PbCl6, and (d) Rb2PbBr6 compounds calculated using the PBE-GGA exchange–correlation functional. | ||
![]() | ||
| Fig. 3 Energy band gap values of Rb2BX6 compounds using (a) GGA-PBEsol and (b) m-GGA exchange–correlation functional. | ||
The HSE06 hybrid functional, which incorporates a fraction of exact exchange, significantly increases the band gaps to 4.231, 2.828, 1.946, and 1.742 eV, respectively (Fig. 4a–d), providing the most accurate estimate consistent with experimental trends in similar perovskites. A comparison with earlier reports shows good agreement. For instance, Na2SnCl6 (2.77 eV) and Na2SnBr6 (1.12 eV) using GGA-PBE follow the same halogen-dependent narrowing trend as in our Rb-based systems.49 Similarly, experimental values for K2SnCl6 (4.54 eV), K2SnBr6 (3.27 eV), and K2SnI6 (1.90 eV) also confirm the systematic reduction of the band gap with increasing halogen size.50 Theoretical calculations reported Cs2SnCl6 (3.90 eV), Cs2SnBr6 (2.70 eV), and Cs2SnI6 (1.26 eV),51 again consistent with our observations. Additionally, reported theoretical and experimental results for Rb2SnBr6 show band gaps of 1.28 eV and 4.79 eV,47,48 which align closely with our PBE and HSE06 values, respectively.
![]() | ||
| Fig. 4 Energy band gap values of (a) Rb2SnCl6, (b) Rb2SnBr6, (c) Rb2PbCl6, and (d) Rb2PbBr6 compounds calculated using the hybrid HSE06 exchange–correlation functional. | ||
The variation in band gaps across the compounds arises from both cation and halide effects. Substituting Sn with Pb reduces the gap due to the larger spin–orbit coupling and more diffuse Pb 6s/6p orbitals, which lower the conduction–valence band separation.
Similarly, replacing Cl with Br decreases the gap, as Br 4p orbitals are higher in energy than Cl 3p orbitals, elevating the valence band maximum. Consequently, Rb2SnCl6 exhibits the widest gap, while Rb2PbBr6 has the smallest, a trend consistent across all functionals. The choice of functional influences the absolute values but preserves the relative trends: PBE tends to slightly underestimate gaps due to self-interaction errors, PBEsol reduces gaps further due to lattice compression, m-GGA improves description with gradient corrections, and HSE06 provides the most reliable quantitative values. These observations highlight the importance of using multiple functionals to balance computational efficiency and predictive accuracy in perovskite band structure calculations.
Furthermore, the analysis of the PDOS and TDOS diagrams in Fig. 5a–d corroborates the band gap values obtained from the band structure calculations using the GGA-PBE functional, which provides a computationally efficient and reliable estimate. Notably, these calculations require relatively short computational time, making them suitable for rapid screening of multiple compounds. The PDOS graphs in Fig. 5 highlight the distinct electronic properties of Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6.
![]() | ||
| Fig. 5 The Partial Density of States (PDOS) of (a) Rb2SnCl6, (b) Rb2SnBr6, (c) Rb2PbCl6, and (d) Rb2PbBr6 double perovskite materials using the PBE-GGA exchange–correlation functional. | ||
In Rb2SnCl6 (Fig. 5a), the valence band is primarily composed of Cl-3p states with contributions from Sn-5s, while the conduction band mainly consists of Sn-5p and Rb-4p states, indicating a wide band gap and semiconducting behavior. Similarly, Rb2SnBr6 in Fig. 5b follows this trend, but the presence of Br-4p states shifts the conduction band downward, reducing the band gap and enhancing visible-light absorption. In Rb2PbCl6 Fig. 5c, the valence band originates from Cl-3p and Pb-6s states, while the conduction band is primarily composed of Pb-6p states, resulting in a moderate band gap. Among the four materials, Rb2PbBr6 Fig. 5d exhibits the smallest band gap, with Br-4p and Pb-6s states dominating the valence band, while Pb-6p states form the conduction band, enhancing its optical absorption properties.
The dielectric function, ε(ω), is a complex optical parameter that describes how a material interacts with electromagnetic radiation across different photon energies. It consists of the real part, εr(ω), which indicates how light is slowed down and polarized within the material, and the imaginary part, εimag(ω), which represents the amount of energy absorbed from the electromagnetic wave.70
| ε(ω) = εr(ω) + iεimag(ω) | (12) |
![]() | (13) |
![]() | (14) |
![]() | ||
| Fig. 6 Presents Rb2BX6 (B = Sn/Pb, X = Cl/Br) (a) dielectric function, (b) refractive index, (c) conductivity, (d) reflectivity, (e) absorption coefficient, and (f) electron loss function. | ||
Beyond these points, the refractive index η(ω) for both compounds decline as ω rises. The refractive index η(ω) consists of real η1 and imaginary η2 parts, which are represented as.76 The figure Fig. 6b, presents the refractive index η(ω) (both real η1 and imaginary η2 components) as a function of photon energy for Rb2BX6 (B = Sn, Pb; X = Cl, Br) compounds, where Rb2SnBr6 and Rb2PbCl6 exhibit prominent optical responses. Rb2SnBr6 starts at ∼2.3 at 0 eV, with peaks at 1.8 (∼5 eV) and 1.5 (∼10 eV), while Rb2PbCl6 begins at ∼2.1, peaking at 1.7 (∼6 eV) and 1.3 (∼10 eV); their imaginary parts show strong absorption around 4–10 eV. All compounds tend to converge to zero beyond 20 eV, highlighting Rb2SnBr6 and Rb2PbCl6 as highest promising for optoelectronic applications. The Fig. 6c illustrates the optical conductivity σ(ω) of Rb2BX6 (B = Sn, Pb; X = Cl, Br) as a function of photon energy. Rb2SnBr6 and Rb2PbCl6 show strong optical responses, with real conductivity peaking around 4.5 at ∼8 eV for Rb2SnBr6 and 4.2 at ∼10 eV for Rb2PbCl6, indicating efficient photon absorption. All compounds exhibit fluctuations, with conductivity declining beyond 20 eV, suggesting diminishing optical transitions at higher energies. The Fig. 6d depicts the reflectivity R(ω) of Rb2BX6 (B = Sn, Pb; X = Cl, Br) as a function of photon energy, eV. Rb2SnBr6 and Rb2PbCl6 show moderate reflectivity, peaking around 0.22 at ∼10 eV for Rb2SnBr6 and 0.3 at ∼20 eV for Rb2PbCl6, indicating their optical response. All compounds exhibit fluctuations, with reflectivity generally remaining below 0.5, suggesting strong light absorption in the studied energy range. The optical absorption coefficient indicates how effectively a material absorbs light, which is vital for solar cells, especially within the 1.5–4.0 eV range.77,78 Extending the analysis up to 30 eV (ref. 79–81) also reveals deeper electronic transitions, relevant for UV photodetectors, PL behavior, and radiation shielding applications.82,83 Fig. 6e first highlights the visible absorption range (1.5–4.0 eV), confirming that Rb2BX6 materials exhibit strong absorption in the visible spectrum. Additionally, the absorption spectra of Rb2SnBr6 and Rb2PbCl6 show distinct peaks: Rb2SnBr6 peaks at ∼8, 13, and 18 eV with intensities up to 25 × 104 cm−1, while Rb2PbCl6 peaks at ∼7, 12, 16, and 20 eV, exceeding 27 × 104 cm−1 near 17–20 eV. These high-energy peaks indicate strong ultraviolet absorption, making them suitable for UV detection, PL, and related optoelectronic applications. The sharper features in Rb2PbCl6 suggest stronger absorption transitions, whereas Rb2SnBr6 shows a smoother absorption profile. Finally, the loss function L(ω) in Fig. 6f for Rb2SnBr6 and Rb2PbCl6 shows prominent peaks in the 18–24 eV range, with Rb2PbCl6 peaking around 21 eV and Rb2SnBr6 slightly lower. Rb2SnCl6 exhibits the highest peak (∼4.5) at ∼21 eV, indicating strong plasmonic resonance, while Rb2PbBr6 and Rb2PbCl6 display similar peak positions with varying intensities. All compounds show minimal loss below 15 eV, with a sharp increase beyond 18 eV, highlighting the role of Sn vs. Pb and Cl vs. Br in plasmonic excitations and energy dissipation.
![]() | (15) |
![]() | (16) |
![]() | (17) |
![]() | (18) |
Finally, the anisotropy factors (A, AU, Aeq, AG) assess how elastic properties vary in different crystallographic directions. The Zener anisotropic factor (A) is particularly significant, where A = 1 denotes a perfectly isotropic material. Larger deviations from 1 indicate anisotropy, meaning the material's properties depend heavily on direction.
Rb2PbBr6, with an extremely high A value of 9.751, is highly anisotropic, meaning its mechanical behavior changes significantly depending on the direction of applied force. In contrast, Rb2SnCl6, with an A value of 0.845, is the most isotropic, meaning it behaves uniformly in all directions. The provided Fig. 7 and 8 represent a comprehensive visualization of the elastic anisotropy of different halide perovskite compounds, likely Rb2SnCl6, Rb2SnBr6, Rb2PbCl6, and Rb2PbBr6. The figures include 2D polar plots and corresponding 3D surfaces, illustrating how elastic properties vary with crystallographic direction. Fig. 7 and 8 illustrate the Anisotropy index for Rb2BX6 materials, representing each row as depicting a different mechanical parameter, such as Young's modulus, shear modulus, or Poisson's ratio. The leftmost column contains 2D polar plots, showing the directional dependence of a given property, while the adjacent 3D plots provide a spatial representation of the same data. The rightmost columns contain additional 2D and 3D visualizations, often incorporating multiple data sets (denoted by green and blue curves), possibly comparing theoretical and experimental values or different anisotropy measures. The diversity in shape and symmetry among the figures reflects the varying degrees of anisotropy in these materials. More spherical shapes, as seen in the middle row, suggest nearly isotropic behavior, indicating uniform mechanical responses in all directions. In contrast, highly distorted or lobed shapes, particularly in the top and bottom rows, reveal strong anisotropic behavior, meaning the material exhibits direction-dependent stiffness or flexibility.42,43 This anisotropy is critical in determining mechanical performance for applications in flexible electronics, optoelectronics, and thermoelectric materials. The contrasting line colors and the interplay of smooth versus wavy contours likely indicate different computational models or comparative analyses of elastic responses.86
![]() | ||
| Fig. 7 Anisotropic 2D and 3D Spatial dependence representation of (a) Youngs moduli, (b) linear compressibility, (c) shear moduli, (d) Poisson's ratio of the Rb2BX6 materials. | ||
![]() | ||
| Fig. 8 Anisotropic 2D and 3D Spatial dependence representation of (a) Youngs moduli, (b) linear compressibility, (c) shear moduli, (d) Poisson's ratio of the Rb2BX6 materials. | ||
Therefore, phonon calculations are essential for predicting material performance, guiding chemical substitutions, and designing compounds with robust structural, thermal, and optoelectronic properties. Fig. 9 illustrates the phonon dispersion curves of the cubic double perovskites along the high-symmetry directions W–L–Γ–X–W–R, with (a) Rb2SnCl6, (b) Rb2SnBr6, (c) Rb2PbCl6, and (d) Rb2PbBr6. Phonon dispersion analysis is a critical tool for assessing the dynamical stability of crystalline materials, as negative frequencies (imaginary modes) indicate potential lattice instabilities that could lead to phase transitions or structural distortions. In the present case, the Cl-based compounds, Rb2SnCl6 and Rb2PbCl6, display entirely positive phonon frequencies across the full Brillouin zone. This observation confirms the absence of any unstable vibrational modes and suggests that these materials are dynamically stable. The high-frequency optical phonon branches, primarily associated with the vibrations of the lighter Cl atoms, reflect strong bonding interactions within the lattice, contributing to the rigidity and robustness of the crystal structure. Conversely, the Br-based analogues, Rb2SnBr6 and Rb2PbBr6, exhibit phonon branches that dip below zero near certain high-symmetry points, signaling the presence of imaginary frequencies. These negative modes indicate lattice instabilities, likely arising from the larger atomic radius and higher mass of Br compared to Cl, which weaken the restoring forces in the lattice. Consequently, the Br-substituted compounds are prone to structural distortions or phase transitions under ambient conditions, reflecting softer bonding characteristics and reduced lattice stiffness.
![]() | ||
| Fig. 9 Phonon dispersion curves of (a) Rb2SnCl6, (b) Rb2SnBr6, (c) Rb2PbCl6 and (d) Rb2PbBr6 cubic double perovskites along the high-symmetry directions. | ||
| Compound | Charge spilling | Species | Ion | Mulliken atomic populations | Mulliken charge | Hirshfeld charge | ||||
|---|---|---|---|---|---|---|---|---|---|---|
| s | p | d | f | Total | ||||||
| Rb2PbCl6 | 0.15% | Rb | 2 | 2.06 | 6.10 | 0.13 | 0.0 | 8.28 | 0.72 | 0.22 |
| Pb | 1 | 3.60 | 7.56 | 10.0 | 0.0 | 21.16 | 0.84 | 0.44 | ||
| Cl | 6 | 1.96 | 5.42 | 0.00 | 0.0 | 7.38 | −0.38 | −0.17 | ||
| Rb2PbBr6 | 0.12% | Rb | 2 | 2.09 | 6.16 | 0.25 | 0.0 | 8.50 | 0.50 | 0.21 |
| Pb | 1 | 3.64 | 7.75 | 10.0 | 0.0 | 21.39 | 0.61 | 0.36 | ||
| Br | 6 | 1.93 | 5.34 | 0.00 | 0.0 | 7.27 | −0.27 | −0.14 | ||
| Rb2SnCl6 | 0.20% | Rb | 2 | 2.06 | 6.08 | 0.13 | 0.0 | 8.27 | 0.73 | 0.23 |
| Sn | 1 | 1.39 | 1.66 | 10.0 | 0.0 | 13.05 | 0.95 | 0.42 | ||
| Cl | 6 | 1.96 | 5.45 | 0.00 | 0.0 | 7.40 | −0.40 | −0.16 | ||
| Rb2SnBr6 | 0.16% | Rb | 2 | 2.08 | 6.18 | 0.24 | 0.0 | 8.50 | 0.50 | 0.21 |
| Sn | 1 | 1.67 | 1.85 | 10.0 | 0.0 | 13.52 | 0.48 | 0.34 | ||
| Br | 6 | 1.88 | 5.36 | 0.00 | 0.0 | 7.25 | −0.25 | −0.14 | ||
Mulliken charge analysis confirms that Rb consistently has a positive charge (0.50 to 0.73), Pb and Sn exhibit moderate positive values (0.48 to 0.95), and halides carry negative charges (−0.25 to −0.40), highlighting the ionic nature of these materials. Among them, Sn in Rb2SnCl6 has the highest Mulliken charge (0.95), suggesting stronger ionic bonding compared to Pb. Hirshfeld charge analysis, which generally yields smaller charge values, similarly shows Rb with a slightly positive charge (∼0.21 to 0.23), Pb and Sn with moderate positive values (∼0.34 to 0.44), and halogens with negative charges (−0.14 to −0.17), reinforcing the observed charge transfer trends. Comparing Pb-based and Sn-based compounds, Sn perovskites show higher positive Sn charges and lower total electron populations, suggesting a more ionic nature, while Pb compounds exhibit stronger d-orbital contributions, indicating enhanced covalent character. Similarly, chlorides (Rb2SnCl6, Rb2PbCl6) show stronger ionic bonding due to higher halide Mulliken charges (−0.38 to −0.40) compared to bromides (−0.25 to −0.27), where charge delocalization is more prominent. These findings highlight the greater ionic nature of Rb2SnCl6 and the more covalent behavior of Rb2PbBr6, influencing their electronic structure, stability, and potential optoelectronic applications.
Table 5 presents key material parameters for the FTO, CdS, Rb2SnBr6, and Rb2PbCl6 layers used in heterostructures.
| Parameters | FTO89–91 | CdS92 | Rb2SnBr6 | Rb2PbCl6 |
|---|---|---|---|---|
| Thickness (nm) | 50 | 50 | 900 | 900 |
| Band gap, Eg (eV) | 3.6 | 2.42 | 1.451 | 1.379 |
| Dielectric permitivity, εr | 10 | 9.35 | 3.71 | 3.46 |
| Electron affinity, χ (eV) | 4.5 | 4.30 | 3.860 | 4.910 |
| CB effective density of states, NC (1/cm3) | 2 × 1018 | 2.2 × 1018 | 9.424 × 1018 | 9.633 × 1018 |
| VB effective density of states, NV (1/cm3) | 1.8 × 1019 | 1.8 × 1019 | 1.415 × 1019 | 1.741 × 1019 |
| Shallow uniform acceptor density, NA (1/cm3) | 0 | 0 | 1 × 1019 | 1 × 1019 |
| Shallow uniform donor density, ND (1/cm3) | 1 × 1018 | 1 × 1017 | 0 | 0 |
| Electron thermal velocity (cm s−1) | 1 × 107 | 1 × 107 | 1 × 107 | 1 × 107 |
| Hole thermal velocity (cm s−1) | 1 × 107 | 1 × 107 | 1 × 107 | 1 × 107 |
| Electron mobility, μn (cm2 V−1 s−1) | 50 | 100 | 100 | 100 |
| Hole mobility, μh (cm2 V−1 s−1) | 20 | 25 | 20 | 50 |
| Total defect density (cm−3) | 1 × 1014 | 1 × 1012 | 1 × 1013 | 1 × 1013 |
Electron affinity (χ): calculated using the equation:
| χ = EVac − ECBM | (19) |
Dielectric function or relative permittivity (ε):
The frequency-dependent dielectric function
| ε(0) = εr(0) + iεimag(0) | (20) |
Eqn (20) represents eqn (12) evaluated at ω = 0 (static limit). At zero frequency, the imaginary part becomes negligible (εimag(0) ≈ 0), and the static dielectric constants are εr(0) = 3.71 for Rb2SnBr6 and εr(0) = 3.46 for Rb2PbCl6 and εr(0) and εimag(0) are connected through the Kramers–Kronig relations.94
Shallow donor and acceptor densities (ND, NA):
These values were adopted from literature reports on similar perovskite materials, as their direct calculation requires detailed defect energetics. The selected values are representative of solar cell device simulations.95
Electron and hole mobilities (μe, μh):
Mobilities were estimated using:
![]() | (21) |
The standard formulas below were employed to determine the effective density of states in the valence band (NV) and conduction band (NC) for double perovskite materials.98
![]() | (22) |
![]() | (23) |
The thickness of each layer is 50 nm for FTO and CdS, and 900 nm for both Rb2SnBr6 and Rb2PbCl6. The band gaps (Eg) are 3.6 eV for FTO, 2.42 eV for CdS, 1.451 eV for Rb2SnBr6, and 1.379 eV for Rb2PbCl6. The relative permittivity (εr) values are 10, 9.35, 3.71, and 3.46, respectively. The electron affinity (χ) ranges from 3.860 eV (Rb2SnBr6) to 4.910 eV (Rb2PbCl6). The conduction band effective density of states (NC) is highest for Rb2PbCl6 (9.633 × 1018 cm−3) and lowest for FTO (2 × 1018 cm−3). The valence band effective density of states (NV) follows a similar pattern, with 1.8 × 1019 cm−3 for FTO and CdS, 1.415 × 1019 cm−3 for Rb2SnBr6, and 1.741 × 1019 cm−3 for Rb2PbCl6. Regarding carrier concentrations, the shallow uniform acceptor density (NA) is 1 × 1019 cm−3 for both Rb-based compounds, while it is zero for FTO and CdS. The shallow uniform donor density (ND) is 1 × 1018 cm−3 for FTO, 1 × 1017 cm−3 for CdS, and zero for Rb2SnBr6 and Rb2PbCl6. Electron and hole thermal velocities are consistently 1 × 107 cm s−1 across all materials. The electron mobility (μn) is 50 cm2 V−1 s−1 for FTO, 100 cm2 V−1 s−1 for CdS and Rb-based compounds. Hole mobility (μh) varies from 20 cm2 V−1 s−1 (FTO and Rb2SnBr6) to 50 cm2 V−1 s−1 (Rb2PbCl6). Lastly, the total defect density is 1 × 1014 cm−3 for FTO, 1 × 1012 cm−3 for CdS, and 1 × 1013 cm−3 for both Rb2SnBr6 and Rb2PbCl6.
Table 6 summarizes the defect characteristics at the CdS/Rb2SnBr6 and CdS/Rb2PbCl6 interfaces. Both interfaces exhibit a total interface defect density of 1 × 1011 cm−2 and a neutral defect type. This value is not directly measured from experiments but is a commonly assumed moderate defect density for interface states in double perovskite solar cells, as used in previous simulation studies.99,100 The capture cross-section for both electrons and holes (1 × 10−19 cm2) at each interface was adopted from the standard SCAPS-1D default values,99,101,102 which are widely used in modeling studies of perovskite-based devices. These assumptions provide a realistic representation of interfacial recombination processes in the absence of experimental data. The consistent values further indicate similar defect behavior at the two heterostructure interfaces.
| Interfaces | Total defect density (cm−2) | Defect type | Capture cross section: electrons/holes (cm2) |
|---|---|---|---|
| CdS/Rb2SnBr6 | 1 × 1011 | Neutral | 1 × 10−19 |
| CdS/Rb2PbCl6 | 1 × 1011 | Neutral | 1 × 10−19 |
![]() | ||
| Fig. 11 Effect of variation in the thickness of absorber layer on Rb2SnBr6 and Rb2PbCl6 materials on PV parameters of VOC, JSC, FF, and PCE. | ||
| Thickness of absorber layer | Rb2SnBr6 | Rb2PbCl6 | ||||||
|---|---|---|---|---|---|---|---|---|
| VOC (V) | JSC (mA cm−2) | FF (%) | PCE (%) | VOC (V) | JSC (mA cm−2) | FF (%) | PCE (%) | |
| 0.3 | 0.988 | 16.0361 | 84.596 | 13.4043 | 0.9437 | 18.324 | 84.995 | 14.698 |
| 0.6 | 1.001 | 21.18805 | 85.352 | 18.1159 | 0.961 | 23.636 | 85.671 | 19.467 |
| 0.9 | 1.007 | 23.69723 | 85.628 | 20.4388 | 0.969 | 26.227 | 85.891 | 21.838 |
| 1.2 | 1.010 | 25.15447 | 85.759 | 21.7960 | 0.974 | 27.737 | 85.953 | 23.236 |
| 1.5 | 1.012 | 26.1256 | 85.835 | 22.6798 | 0.978 | 28.721 | 85.960 | 24.151 |
| 1.8 | 1.013 | 26.76092 | 85.884 | 23.2992 | 0.980 | 29.410 | 85.960 | 24.796 |
| 2.1 | 1.014 | 27.24803 | 85.917 | 23.7564 | 0.982 | 29.919 | 85.966 | 25.274 |
However, not all parameters improve indefinitely. The investigation demonstrates that altering the absorber layer thickness has a moderate effect on VOC and FF in Rb2SnBr6 and Rb2PbCl6 double perovskite structures. For both Rb2SnBr6 and Rb2PbCl6, the VOC and FF increase up to approximately 0.9 μm, after which they plateau due to recombination balancing the photogenerated carrier density. As the absorber layer thickness increases, enhanced optical absorption, particularly at longer wavelengths, leads to an increase in the JSC. This continues until approximately 1.5 μm, beyond which JSC and the PCE begin to saturate due to the onset of absorption saturation and increased charge carrier recombination. Meanwhile, the VOC and FF also improve initially but stabilize earlier, around 0.9 μm. These trends are consistent with previously reported studies that highlight the trade-offs between light absorption and recombination losses in thicker absorber layers.79,103,104 However, VOC and FF saturate at around 0.9 μm, while further increases in JSC and PCE beyond this point show only marginal improvements. Therefore, 0.9 μm is considered the optimal absorber thickness. Beyond this threshold, additional thickness leads to higher material usage and fabrication costs without significant efficiency gains, making further expansion economically inefficient.
These findings align with prior simulation-based studies and provide an upper threshold beyond which increasing absorber thickness yields diminishing performance returns. The optimized efficiencies are 20.44% for Rb2SnBr6 and 21.84% for Rb2PbCl6, with Rb2PbCl6 exhibiting higher efficiency. However, the presence of lead raises environmental concerns, making Rb2SnBr6 a preferable alternative as a lead-free material with promising potential for future solar cell applications.
![]() | ||
| Fig. 12 Effects of defect density in the Rb2SnBr6 and Rb2PbCl6 absorber layer on PV parameters: (a) VOC, (b) JSC, (c) FF, and (d) PCE. | ||
Notably, optimal performance is observed at a defect density of 1013 cm−3 for the CdS ETL layer, where Rb2SnBr6 achieves a VOC of 1.007 V, JSC of 23.697 mA cm−2, FF of 85.628%, and PCE of 20.439%. In comparison, Rb2PbCl6 reaches a VOC of 0.996 V, JSC of 26.227 mA cm−2, FF of 85.891%, and PCE = 21.893%. It is important to emphasize that these high FF values result from simulations under idealized conditions, assuming negligible interfacial recombination, optimal charge transport, and uniform material quality. While these values represent the theoretical performance limits, actual experimental devices may exhibit lower FF due to non-idealities. These results underscore the critical role of defect engineering in optimizing the efficiency of double perovskite solar cells.
![]() | ||
| Fig. 13 Analysis of photovoltaic performance parameters in Rb2SnBr6 and Rb2PbCl6 perovskite materials. | ||
In comparison, Rb2PbCl6 exhibits slightly higher values for PCE (21.84%) and JSC (26.227 mA cm−2) but a lower VOC (0.970 V), while maintaining a similar FF (85.89%). The left y-axis represents PCE and JSC, whereas the right y-axis corresponds to VOC and FF, allowing a clear visualization of the performance trends. Although Rb2PbCl6 demonstrates superior photovoltaic performance with higher PCE and JSC, its lead content poses significant environmental and toxicity concerns. On the other hand, Rb2SnBr6, despite its slightly lower efficiency, offers a lead-free and environmentally safer alternative. Given the increasing focus on sustainable and non-toxic materials for solar energy applications, Rb2SnBr6 emerges as a promising candidate for future solar cell development, striking a balance between performance and environmental responsibility.
Table 8 compares the performance parameters of photovoltaic (PV) cells for Rb2SnBr6 and Rb2PbCl6 with previously reported cell structures. Table 8 presents key metrics including PCE, JSC, FF, and VOC.
| Structures | PCE (%) | JSC (mA cm−2) | VOC (volt) | FF (%) | References |
|---|---|---|---|---|---|
| Al/FTO/SnS2/Ca3NCl3 | 8.54 | 7.044 | 1.378 | 88.10 | 108 |
| Al/FTO/SnS2/Sr3NCl3 | 18.11 | 16.786 | 1.248 | 86.44 | 108 |
| Al/FTO/CdS/Ba3NCl3/Au | 32.00 | 38.21 | 1.036 | 80.75 | 109 |
| Al/FTO/CdS/Rb2SnBr6/Ni | 20.44 | 26.697 | 1.0073 | 85.63 | This work |
| Al/FTO/CdS/Rb2PbCl6/Ni | 21.84 | 26.227 | 0.9695 | 85.89 | This work |
The cells reported in this work include the Al/FTO/CdS/Rb2SnBr6/Ni structure, which achieved a PCE of 20.44%, a JSC of 26.697 mA cm−2, VOC of 1.0073 V, and FF of 85.63%, and the Al/FTO/CdS/Rb2PbCl6/Ni structure, which showed a PCE of 21.84%, JSC of 26.227 mA cm−2, VOC of 0.9695 V, and FF of 85.89%. These results are competitive when compared to earlier reported structures, such as Al/FTO/SnS2/Ca3NCl3 with a PCE of 8.54% and Al/FTO/CdS/Ba3NCl3/Au with a PCE of 32.00%.
![]() | ||
| Fig. 14 (a and c) J–V characteristics, and (b and d) Q–E response for Rb2SnBr6 and Rb2PbCl6 materials. | ||
The thickness of the absorber layer significantly influences the J–V and Q–E characteristics of the solar cell. Increasing the absorber thickness enhances JSC due to improved light absorption and higher photocurrent generation. However, while thicker layers enhance absorption, they may also lead to increased recombination losses, potentially reducing VOC, FF, and overall efficiency.
| This journal is © The Royal Society of Chemistry 2025 |