Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Numerical optimization of Rb2AuScBr6 and Rb2AuScCl6-based lead-free perovskite solar cells: device engineering and performance mapping

Md. Al-Amina, A. Haquea, S. Mahmud*a, M. M. Hossainb, M. M. Uddinb and M. A. Ali*b
aDepartment of Electrical and Electronic Engineering, Jatiya Kabi Kazi Nazrul Islam University (JKKNIU), Mymensingh-2224, Bangladesh. E-mail: shuaib.eee.iu@gmail.com
bAdvanced Computational Materials Research Laboratory, Department of Physics, Chittagong University of Engineering and Technology (CUET), Chattogram-4349, Bangladesh

Received 27th September 2025 , Accepted 13th November 2025

First published on 19th November 2025


Abstract

Perovskite solar cells (PSCs) exhibit significant potential for next-generation photovoltaic technology, integrating high power conversion efficiency (PCE), cost-effectiveness, and tunable optoelectronic properties. This report presents a comprehensive numerical optimization of Rb2AuScBr6 and Rb2AuScCl6-based PSCs, with particular emphasis on the influence of electron transport layers (ETLs) and critical device parameters. The configuration ITO/TiO2/Rb2AuScBr6/CBTS/Ni achieves a PCE of 27.49%, whereas the configuration ITO/WS2/Rb2AuScCl6/CBTS/Ni attains 22.41%, thereby underscoring the high efficiency of these lead-free materials. Device performance is markedly improved through increased perovskite layer thickness and reduced defect density. Further stabilization of performance is achieved by optimizing electron affinity, series resistance, and shunt resistance. Additionally, thermal stability is enhanced through the adjustment of operational temperature. The superior PCE observed in Rb2AuScBr6 is ascribed to the selection of the ETL, an optimal band gap, absorber layer thickness, lower defect density, and appropriate contact interfaces. Overall, these Rb2AuScBr6 and Rb2AuScCl6 perovskites demonstrate exceptional promise for practical, efficient, and stable PSC applications, thereby encouraging further experimental validation and device engineering.


1. Introduction

The growing global need for energy, fueled by rapid population growth and industrial development, has spurred a concerted search for sustainable and eco-friendly energy sources.1 Among these, solar energy provides an ample, renewable resource that can be harnessed to overcome the constraints of fossil fuels in a cost-efficient manner.2 With its unique characteristics and advantages, photovoltaic (PV) technology is a rather efficient method for converting sunlight into electricity, which presents low operation cost and less environmental pollution than other traditional energy generation methods.3

Perovskite materials have garnered widespread attention in recent years owing to their outstanding light-harvesting capabilities, tunable band gaps, and exceptional power conversion efficiencies (PCEs).4 Within this family, double perovskites—typically denoted A2BB′X6 (X = halide)—offer a particularly promising avenue.5 Unlike traditional single-cation ABX3 perovskites, their expanded B-site cation combinations facilitate precise tuning of electronic structures and band gaps. This flexibility enables the creation of lead-free, environmentally benign alternatives with band gaps better matched to the solar spectrum.6,7

One notable candidate, Cs2AgBiBr6, exhibits a large but stable indirect band gap of around 2.2 eV. Treating conventional perovskites with Cs2AgBiBr6 nanoparticles has yielded enhanced stability and an impressive PCE of 19.52%, underscoring the advantage of its passivation capabilities.8 Through compositional engineering, this band gap can be further tuned: Tl-doping narrows it to 1.57 eV, while Sb-for-Bi substitution reaches 1.86 eV, improving absorption and photovoltaic potential.9

Similarly, Cs2AgInBr6 has been explored using first-principles and optical studies, revealing a direct band gap of approximately 1.427 eV and excellent absorption (>104 cm−1), positioning it as a viable PV absorber.10 Another intriguing family, A2AgIrCl6 (A = Cs, Rb, K), displays direct band gaps of about 1.43–1.55 eV (Cs2 = 1.43 eV; Rb2 = 1.50 eV; K2 = 1.55 eV), alongside low effective electron masses and strong visible-light absorption, making them highly promising for solar cell applications.11

A different strategy involves alloying: in Cs2Ag (SbxBi1−x) Br6, band gaps reduce below those of the pure end members down to about 2.08 eV when x = 0.9, thanks to band-gap bowing from Sb–Bi orbital mixing.12 Additional studies have explored Cs2AgBiBr6 doped with aluminum: the pristine material shows Eg ≈ 1.91 eV, whereas Al-doped films exhibit a slightly lowered Eg of 1.82 eV, resulting in improved device metrics (PCE from ∼3.02% to ∼3.40%).13 Another notable double perovskite, Cs2CuBiCl6, has emerged via SCAPS-1D simulation as a lead-free absorber with high predicted efficiencies; optimized devices yielded PCE up to 24.51%, Voc = 1.73 V, and Jsc = 32.82 mA cm−2.14 These theoretical models affirm the material's considerable promise for durable, high-performance photovoltaics.

Collectively, these studies underscore the diverse strategies available—doping, substitution, alloying, and novel compounds—to engineer double perovskites with direct, suitably low band gaps (spanning ∼1.4 to 2.2 eV). Realizing these tunable properties positions them as key materials for next-generation, lead-free, high-efficiency solar technologies, potentially even surpassing conventional limits.15 Several electron transport layers (ETLs) have been explored to improve the performance of perovskite solar cells (PSCs). WS2, a transition metal dichalcogenide, offers suitable conduction band alignment and decent electron mobility, enhancing power conversion efficiency (PCE).16 ZnO, with its wide band gap (∼3.3 eV), high electron mobility, and transparency, is attractive, though interfacial reactions with perovskites may cause stability issues requiring surface modification.17 PCBM, a fullerene derivative, is widely used for its strong electron-accepting ability and reduced recombination losses.18 TiO2 is the most common ETL due to its band alignment, chemical stability, and transparency, but its low electron mobility and recombination tendencies necessitate surface treatments or doping.19 IGZO, with high electron mobility and optical transparency, and C60, with strong electron transport properties, are also promising ETLs.20,21

For hole transport layers (HTLs), CBTS has demonstrated strong potential owing to its favorable energy level alignment and ability to enhance hole extraction and stability.22 Other candidates include CdTe, CuI, PTAA, P3HT, and PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS, each offering distinct advantages such as high hole mobility, thermal stability, or compatibility with perovskite layers, though limitations like poor conductivity or stability often require doping or modification.23–27

In this work, we present the simulation and optimization of lead-free double perovskite solar cells based on Rb2AuScBr6 and Rb2AuScCl2 absorbers. Using SCAPS-1D, we systematically investigate n–i–p planar heterojunction structures with various ETLs (IGZO, WS2, ZnO, PCBM, C60, and TiO2) and HTLs (CBTS, CdTe, CuI, PTAA, P3HT, and PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS). Among the HTLs, CBTS demonstrates superior performance for both absorbers. Device parameters such as absorber thickness, defect density, resistances, and operating temperature are tuned to optimize photovoltaic performance. The results indicate that these double perovskite materials hold great promise for high-efficiency, environmentally friendly PSCs, contributing to the development of next-generation renewable energy technologies.28–31

2. Methodology

The numerical studies of this work were performed using the Solar Cell Capacitance Simulator (SCAPS-1D), the widely used one-dimensional solar cell simulation tool developed at the University of Ghent, Belgium. Data analysis of a wide range of photovoltaic device parameters, such as energy band alignment, charge transport mechanisms, recombination losses, and current–voltage (JV) characteristics, can be performed through SCAPS-1D.32

The simulation process adheres to the usual operating routine of SCAPS-1D as follows:33.

image file: d5ra07344h-u1.tif

The SCAPS-1D simulator is built upon three key semiconductor equations: Poisson's equation, the electron continuity equation, and the hole continuity equation. These equations define how electric potential and charge carriers (electrons and holes) behave and distribute within a semiconductor device. SCAPS-1D numerically solves these equations to model the internal physics of solar cells and predict their performance. Poisson's equation is given as:34–36

 
image file: d5ra07344h-t1.tif(1)
Here, ψ (psi) is the electric potential, ε (epsilon) is the material's ability to store electric charge (called the dielectric constant), q is the charge of an electron, and ρ (def) is the charge from defects inside the material.

Eqn (2) and (3) are used to show how free electrons and free holes are conserved or balanced inside the solar cell device.

 
image file: d5ra07344h-t2.tif(2)
 
image file: d5ra07344h-t3.tif(3)
Here, image file: d5ra07344h-t4.tif is the rate of change of hole concentration p with time, q is the charge of an electron or hole, image file: d5ra07344h-t5.tif is the change of hole current density Jp over distance x, Gp is the hole generation rate, Rp is the hole recombination rate, image file: d5ra07344h-t6.tif is the rate of change of electron concentration, image file: d5ra07344h-t7.tif is the change of electron current density Jn over distance x, Gn is the electron generation rate and Rn is the electron recombination rate.

Eqn (4) and (5) are used to calculate the drift current and the diffusion current.

 
Jn = n + qDnδn (4)
 
JP = pqDpδp (5)
where µn is the mobility of electrons, µp is the mobility of holes and the diffusion coefficient (Dn for electrons, Dp for holes) shows how fast they spread out from high to low concentration.

3. Results & discussion

The design configuration of the Rb2AuScBr6-and Rb2AuScCl6-based perovskite solar cell (PSC) follows an n–i–p planar heterojunction structure (Fig. 1), where the ETL is in the n-region, the perovskite absorber layer in the i-region, and the HTL in the p-region. When exposed to sunlight, excitons (electron–hole pairs) are generated within the perovskite layer.37,38 The built-in electric field at the interfaces aids in exciton dissociation, directing electrons toward the n-layer and holes toward the p-layer for efficient charge extraction.39 Indium-doped tin oxide (ITO) is used as the transparent conductive front contact. Different ETL materials, including IGZO, WS2, ZnO, PCBM, C60, and TiO2, are investigated to optimize electron transport. The Rb2AuScCl6 and Rb2AuScBr6 perovskites serve as the absorber layer, enabling efficient light absorption. Among 36 different ETL–HTL combinations, CBTS is selected as the HTL due to its superior hole transport properties and compatibility with the absorber. The Ni completes the device structure and collects the extracted charges.
image file: d5ra07344h-f1.tif
Fig. 1 The design configuration of the Rb2AuScCl6 and Rb2AuScBr6-based PSC.

Simulations were performed under AM1.5 G solar spectrum conditions to analyze the impact of different ETL and HTL combinations, aiming to determine the most efficient structure. The study explores lead-free, stable perovskite materials as potential candidates for environmentally friendly and high-performance solar cells.

To determine the input parameters required for the equations mentioned above, the following calculation procedures are employed, with all relevant values detailed in Tables 1–3. Eqn (6) defines the carrier's effective mass m* in terms of the curvature of the energy–momentum relation:40

 
image file: d5ra07344h-t8.tif(6)
where ℏ is the reduced Planck constant (1.05 × 10−34 J s−1) and image file: d5ra07344h-t9.tif is the second derivative of the band energy E with respect to the wavevector k.

Table 1 Input parameters of HTLs42
Parameters CBTS CdTe P3HT PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS PTAA CuI
t (nm) 100 200 50 50 150 100
Eg (eV) 1.9 1.5 1.7 1.6 2.96 3.1
X (eV) 3.6 3.9 3.5 3.4 2.3 2.1
Er 5.4 9.4 3 3 9 6.5
NC (cm−3) 2.2 × 1018 8 × 1017 2.0 × 1021 2.2 × 1018 2.0 × 1021 2.8 × 1019
NV (cm−3) 1.8 × 1019 1.8 × 1019 2.0 × 1021 1.8 × 1019 2.0 × 1021 1 × 1019
µe (cm2 V−1 s−1) 30 3.2 × 102 1.8 × 10−3 4.5 × 10−2 1 100
µh (cm2 V−1 s−1) 10 4 × 101 1.86 × 10−2 4.5 × 10−2 40 43.9
ND (cm−3) 0 0 0 0 0 0
NA (cm−3) 1018 2.0 × 1014 1018 1018 1018 1018
Nt (cm−3) 1015 1015 1015 1015 1015 1015


Table 2 Input parameters of ETLs43
Parameters Ws2 ZnO TiO2 PCBM IGZO C60
t (nm) 100 50 30 50 30 50
Eg (eV) 1.8 3.3 3.2 2 3.05 1.7
X (eV) 3.95 4 4 3.9 4.16 3.9
Er 13.6 9.00 9 3.9 10 4.2
NC (cm−3) 1 × 1018 3.7 × 1018 2.0 × 1018 2.5 × 1021 5 × 1018 8 × 1019
NV (cm−3) 2.4 × 1019 1.8 × 1019 1.8 × 1019 2.5 × 1021 5 × 1018 8 × 1019
µe (cm2 V−1 s−1) 100 100 20 0.2 15 8 × 10−2
µh (cm2 V−1 s−1) 100 250 10 0.2 0.1 3.5 × 10−3
ND (cm−3) 1018 1018 9 × 1016 2.93 × 1017 1017 1017
NA (cm−3) 0 0 0 0 0 0
Nt (cm−3) 1015 1015 1015 1015 1015 1015


Table 3 Input parameters of Front contact (ITO) and Absorber Rb2AuScCl6 and Rb2AuScBr6 for initial optimization44 a
Parameters ITO Rb2AuScBr6 Rb2AuScCl6
a [t = thickness, X = electron affinity, Er = dielectric relative permittivity, Nc and Nv = effective density of states in the conduction band and valence band, µe and µh = electron mobility and hole mobility, and ND, NA, and Nt = donor, acceptor, and defect density in Tables (1–3)].
t (nm) 500 1200 1200
Eg (eV) 3.5 1.71 1.93
X (eV) 4 3.95 3.75
Er 9 6.34 6.01
NC (cm−3) 2.2 × 1018 1.33 × 1018 1.94 × 1018
NV (cm−3) 1.8 × 1019 4.56 × 1018 4.59 × 1018
µe (cm2 V−1 s−1) 20 83.20 64.71
µh (cm2 V−1 s−1) 10 43.14 36.40
ND (cm−3) 1021 0 0
NA (cm−3) 0 1015 1015
Nt (cm−3) 1015 1015 1015


Eqn (7) and (8) then use these effective masses to compute the effective density of states in the valence band (Nv) and conduction band (Nc), respectively, at temperature T:

 
image file: d5ra07344h-t10.tif(7)
 
image file: d5ra07344h-t11.tif(8)
where KB is Boltzmann's constant, h is Planck's constant, and image file: d5ra07344h-t12.tif and image file: d5ra07344h-t13.tif are the hole and electron effective masses. Eqn (9) and (10) relate the carrier mobilities µe (electron mobility) and µh (hole mobility) to the carrier charge q, the relaxation time τ, and the respective effective masses:41
 
image file: d5ra07344h-t14.tif(9)
 
image file: d5ra07344h-t15.tif(10)

3.1. Absorber material parameters and band alignment of Rb2AuScBr6 and Rb2AuScCl6-based absorbers with different ETLs and HTL

The electronic and optical suitability of Rb2AuScBr6 and Rb2AuScCl6 as absorber materials has been theoretically validated through density functional theory (DFT) calculations. DFT results confirm indirect band gaps of 1.71 eV and 1.93 eV,45 respectively, within the optimal photovoltaic range (1.4–2.2 eV), with high absorption coefficients (≈106 cm−1) and low reflectivity (<15%), indicating their potential as efficient light-absorbing semiconductors.

Fig. 2(a) illustrates the band alignment of the Rb2AuScBr6 absorber with various ETLs and HTLs. The energy band offsets at these interfaces are critical for efficient charge carrier separation and transport, directly influencing solar cell performance. To explore the dynamics of charge carriers, the conduction band minimum (CB), valence band maximum (VB), and quasi-Fermi levels for electrons (Fn) and holes (Fp) are analyzed. Rb2AuScBr6 has a conduction band at 1.47574 eV, a valence band at −0.23426 eV, and a bandgap (Eg) of 1.71 eV, which favors photon absorption and contributes to better overall power conversion efficiency (PCE). Among the ETLs, TiO2 and ZnO exhibit closely matched conduction band levels (−3.11 eV and −3.26 eV), allowing comparable electron injection capabilities.


image file: d5ra07344h-f2.tif
Fig. 2 Energy band alignment between the different ETL and HTL materials of (a) Rb2AuScBr6 absorber and (b) Rb2AuScCl6 absorber.

WS2 (−1.78 eV) and PCBM (−1.74 eV) also demonstrate suitable alignments, making them promising ETL alternatives. IGZO (−2.94 eV) and C60 (−1.49 eV) offer diverse work functions and conduction band offsets, introducing flexibility for device tuning.

On the HTL side, CBTS (conduction band edge ∼0.15 eV) aligns effectively with the absorber, promoting smooth hole extraction. Other HTLs such as P3HT (0.45 eV), CdTe (0.25 eV), CuI (0.45 eV), PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS (0.65 eV), and PTAA (0.39 eV) exhibit various energy level alignments, influencing hole extraction efficiency. CuI and PTAA, in particular, have high valence band energies (3.55 eV and 3.35 eV, respectively), favoring hole mobility. Nickel (Ni), with a work function of 5.5 eV, ensures efficient hole collection by aligning well with these HTLs, reducing recombination at the back contact.

Fig. 2(b) shows the band alignment of the Rb2AuScCl6 absorber with the same ETLs and HTLs. While the alignment is generally analogous to the Rb2AuScBr6 system, subtle differences are observed in the energy levels, which slightly alter charge transport. Rb2AuScCl6 has a conduction band at 1.67588 eV, a valence band at −0.25412 eV, and a larger bandgap of 1.93 eV. Due to the wider bandgap, Rb2AuScCl6 tends to absorb fewer photons in the solar spectrum, which can reduce current generation and thus lower the overall PCE compared to its Br-based counterpart.46–48

3.2. Optimization of the HTL layer and the ETL layer

The selection of optimal ETL and HTL materials for the Rb2AuScBr6-and Rb2AuScCl6-based perovskite solar cells was guided by both photovoltaic performance parameters (PCE, Voc, Jsc, and FF) and key physical considerations, including energy-band alignment, carrier mobility, defect density, material stability, and interfacial charge-transport behavior. Efficient charge extraction requires precise band matching, wherein the conduction band of the ETL lies slightly below that of the absorber to facilitate electron transfer while suppressing hole leakage, and the valence band of the HTL aligns closely with that of the absorber to enable rapid hole transport. Such optimized alignment minimizes interfacial barriers and recombination losses, thereby enhancing Voc, Jsc, and FF. Secondary selection criteria included high carrier mobility, low trap density, and strong chemical and thermal stability to ensure robust and loss-free charge transport pathways. Among the investigated materials, IGZO and WS2 emerged as the most effective ETLs due to their high electron mobility, low defect density, and superior environmental stability, while CBTS proved to be an excellent HTL owing to its favorable energy-level alignment, high hole mobility, and long-term durability. Furthermore, optimized acceptor doping in the absorber enhanced the internal electric field, promoting efficient charge separation and reducing recombination. The IGZO/CBTS combination for the Rb2AuScBr6 device and the WS2/CBTS combination for the Rb2AuScCl6 device provided the most favorable band alignment and carrier extraction pathways, leading to the highest PCEs.

The variations in efficiency among different ETL/HTL configurations can be primarily attributed to differences in interfacial recombination, which were modeled through interface defect densities (Nt = 1015 cm−3). These optimized configurations exhibited lower recombination losses owing to improved interfacial quality and energetically well-matched interfaces.

3.2.1. Selection of ETL and HTL for Rb2AuScBr6-based solar cells. Fig. 3(a) shows that using WS2 as ETL and CBTS as HTL results in a PCE of 23.68%, with an open-circuit voltage (Voc) of 1.258 V, short-circuit current density (Jsc) of 21.433 mA cm−2, and fill factor (FF) of 87.85%. WS2, a two-dimensional transition metal dichalcogenide, exhibits high electron mobility and favorable energy level alignment with the perovskite absorber, enabling efficient electron extraction. This combination maintains stable charge transport and hinders recombination, leading to high efficiency. However, the lower Voc and FF than that of IGZO translate to slightly lower PCE. Fig. 3(b) depicts that using ZnO as the ETL and CBTS as the HTL lead to PCE 23.65% (Voc = 1.257 V, Jsc = 21.414 mA cm−2, FF = 87.83%). ZnO is a common ETL with high electron transport properties, but its electron mobility being lower than IGZO and WS2, makes its efficiency take a hit. Even though ZnO has good performance, its interface properties and energy level alignment are not as good as IGZO, resulting in a relatively low PCE. In Fig. 3(c) achieves 23.57% PCE is achieved with Voc = 1.256 V, Jsc = 21.378 mA cm−2, and FF = 87.80% with a combination of PCBM ETL and CBTS HTL. A well-known electron acceptor, PCBM, has much lower electron mobility than IGZO, WS2, and ZnO, resulting in lower efficiency. Although the fill factor is indeed high, the reduced electron mobility results in a mild drop of the charge collection efficiency, which ends up decreasing the PCE. Fig. 3(d) demonstrates TiO2 as ETL and CBTS as HTL yields a PCE of 23.64% (Voc = 1.257 V, Jsc = 21.415 mA cm−2, and FF = 87.83%). TiO2 is known to have great electron transport properties and high stability, which makes it a widespread choice for ETL. Nevertheless, due to its low conductivity and moderate energy level alignment, its efficiency is slightly lower than that obtained with WS2 and IGZO. TiO2 is not considered the most optimal in terms of charge transport properties, even though it is widely used. Fig. 3(e) shows that the best performance of 23.90% PCE achieved with IGZO ETL and CBTS HTL, Voc = 1.262 V, Jsc = 21.414 mA cm−2, and FF = 88.48%. IGZO possesses good electron transport properties, thanks to its high electron mobility, which allows for rapid extraction of charge. Also, its low trap density minimizes recombination losses, which leads to better performance. As an HTL, the CBTS offers excellent stability and improved hole transport. The synergy between IGZO ETL and CBTS HTL results in a low recombination with high charge transport, providing the best photovoltaic efficiency with the total highest PCE in all tested combinations. Fig. 3(f) demonstrates that C60 as ETL with CBTS as HTL provide the lowest PCE of 21.57% (Voc = 1.240 V, Jsc = 19.957 mA cm−2, and FF = 87.11%). C60 has reasonable electron transport, but its higher electron mobility than other ETLs leads to the lowest efficiency. While the CBTS is still effective as an HTL, the overall device performance is limited by the reduced electron extraction efficiency of C60. IGZO ETL + CBTS HTL has the highest PCE among all combinations, exceeding 23.90% in the most optimum conditions, thus qualifying as the best performing pair. Thus, the high electron mobility, excellent energy-level alignment and chemical stability of IGZO lead to superior electron extraction, while CBTS provides efficient hole collection with minimal recombination losses. This combination optimizes both charge separation as well as charge transport, yielding the highest overall efficiency.49,50
image file: d5ra07344h-f3.tif
Fig. 3 HTL & ETL selection for Rb2AuScBr6 of (a) WS2 (b) ZnO (c) PCBM, (d) TiO2, (e) IGZO and (f) C60 with HTL of CBTS, CdTe, CuI, PTAA, P3HT and PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS.
3.2.2. Selection of ETL and HTL for Rb2AuScCl6-based solar cells. Fig. 4(a) depicts the optimal performance of 20.99% PCE of WS2 ETL/CBTS HTL is obtained with Voc, Jsc, FF of 1.447 V, 16.289 mA cm−2, and 89.04% respectively. The high electron mobility of WS2, together with good alignment of its energy levels with the perovskite material, gives better electron extraction than TiO2 and ZnO. These also match well with CBTS, which provides ideal hole transport; thus, the pair yields the highest PCE attributed to ideal transport properties of both electrons and holes, resulting in minimized recombination and maximized efficiency during transport. Fig. 4(b) illustrates that when use ZnO as ETL with CBTS as HTL (PCE = 20.21% Voc = 1.451 V, 15.659 mA cm−2, 88.92% FF). Though ZnO has excellent electron transport layer properties, its performance is slightly less favorable than WS2 due to slight differences in the energy level alignment and a lower electron mobility than WS2, yielding a slightly lower PCE. Fig. 4(c) shows PCE reaches 19.68% with this structure obtained by ETL of PCBM and HTL of CBTS, while showing Voc of 1.401 V, Jsc of 15.634 mA cm−2 and FF of 89.84%. While PCBM is known to be a good electron acceptor, the relatively lower mobility of electrons compared to WS2 and ZnO results in a slight decrease in PCE value. Assure that the fill factor is high, the net efficiency is reduced by a factor of lower electron mobility. While the configuration of TiO2 ETL and CBTS HTL enables a PCE of 20.33% (Voc = 1.452 V, Jsc = 15.659 mA cm−2, and FF = 89.41%) (Fig. 4(d)). This arrangement strikes a good trade-off between the extraction of electrons and collection of holes, and thus their resulting performance is moderate, whereas the material characteristics of TiO2 do not allow for the most optimal conversion efficiency. In Fig. 4(e) achieves PCE of 19.04% with Voc, Jsc, FF of 1.416 V, 15.657 mA cm−2, 85.85% respectively using IGZO as ETL and CBTS as HTL combination. IGZO possesses high electron mobility yet is somewhat less efficient than WS2 and ZnO, which is likely a result of interface effects and energy level mismatches with the perovskite layer. This means that despite IGZO's electron mobility, the performance is not as good. Fig. 4(f) depicts the composition of C60 ETL and CBTS HTL gives the lowest PCE of 16.51%, with Voc of 1.266 V, Jsc of 14.888 mA cm−2, and FF of 87.61%. While C60 offers decent electron transport, its lower electron mobility compared to the other ETLs results in the overall lowest device efficiency. CBTS is efficient as the HTL, but its reduced electron mobility hampers performance.51
image file: d5ra07344h-f4.tif
Fig. 4 HTL & ETL Selection for Rb2AuScCl6 of (a) Ws2 (b) ZnO (c) PCBM, (d) TiO2, (e) IGZO and (f) C60 with HTL of CBTS, CdTe, CuI, PTAA, P3HT and PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS.

3.3. Role of absorber thickness and ETL thickness on solar cell performance

Fig. 5 demonstrates a series of contour plots showing the PCE of perovskite solar cells. We are testing various materials as the ETL, while using Rb2AuScBr6 as the perovskite absorber material. Each sub-figure (a–f) represents a different ETL configuration.
image file: d5ra07344h-f5.tif
Fig. 5 Contour mapping of PCE, when the absorber is Rb2AuScBr6 and ETLs as (a) C60. (b) IGZO, (c) PCBM, (d) WS2, (e) TiO2 and (f) ZnO.

Fig. 5(a) shows the PCE achieved is around 21.80% when C60 is used as the ETL. The contour lines indicate how the PCE changes with varying thicknesses of both the absorber and the C60 ETL. Fig. 5(b) demonstrates that using Indium Gallium Zinc Oxide (IGZO) as the ETL results in the highest PCE of 24.07%. The contour lines are more spread out compared to C60, suggesting a wider range of thicknesses for both the absorber and IGZO that yield high efficiency. This indicates that the performance is less sensitive to small variations in the thicknesses. Fig. 5(c) illustrates the use of PCBM as the ETL, achieving a peak PCE of approximately 23.80%. The optimal performance region appears to be in the lower left corner, implying that thinner layers of both the absorber and PCBM are generally preferred for maximizing efficiency in this configuration. Fig. 5(d) shows that using WS2 is the ETL reaches PCE is approximately 23.91%. Similar to PCBM, the optimal performance seems to be achieved with thinner layers, as indicated by the concentration of contour lines in the lower-left region. From Fig. 5(a–d) the highest PCE of absorber thicknesses is ≥1200 nm and ETL thicknesses are ≤50 nm. Fig. 5(e) employs titanium dioxide (TiO2) as the ETL and achieves a comparable peak PCE of approximately 23.87%, where the optimal condition is observed at absorber thicknesses ≥1200 nm and an ETL thickness precisely at 200 nm. The contour lines are closely packed, suggesting a narrow region of high efficiency. This implies that the performance is quite sensitive to the thicknesses of both the absorber and TiO2, and precise control is crucial. Fig. 5(f) demonstrates that Zinc Oxide (ZnO) is the ETL, and peak PCE is about 23.88%. Where absorber thicknesses are ≥1200 nm and ETL thicknesses are 450 nm. Like TiO2, the contour plot indicates a narrow region of high efficiency and thus sensitivity to thickness variations.

The superior PCE of IGZO is mainly attributed to its favorable energy band alignment with the Rb2AuScBr6 absorber, which enables electron extraction and prevents the leakage of holes into the ETL. Moreover, the high electron mobility of IGZO decreases recombination losses by quickly transporting charge carriers to the contact. The wide thickness tolerance (shown as contours that are further apart) means that even with small variations in the thickness of the absorber or ETL layer high efficiency can be obtained. This upon favorable band alignment, fast charge transport, and thickness-independence, which altogether result in the superior device performances of IGZO. Fig. 6 shows how the PCE of perovskite solar cells changes when using different electron transport layers (ETLs) with the perovskite material Rb2AuScCl6. Each of the six sub-figures (a–f) represents a different ETL. Fig. 6(a) (C60 ETL) shows a maximum PCE of 16.65%. Where absorber thicknesses are ≥1000 nm and ETL thicknesses are ≤50 nm. The contour lines are relatively spread out, suggesting a less critical dependence on precise thickness control compared to some other ETLs. Fig. 6(b) (IGZO ETL) exhibits a significantly higher maximum PCE, reaching approximately 18.96%. Where absorber thicknesses are ≥1500 nm and ETL thicknesses are ≤450 nm. The high PCE region is broad, indicating good tolerance to variations in both absorber and ETL thicknesses. Fig. 6(c) (PCBM ETL) achieves a maximum PCE around 19.80%. The optimal performance seems to be associated with a specific range of thicknesses, as indicated by the concentrated contour lines. Fig. 6(d) (WS2 ETL) shows a maximum PCE of about 22.76%. This is the highest PCE among the tested ETLs in this figure. The contour plot suggests a well-defined region of optimal performance. From the Fig. 6(c) and (d) reveal that the highest PCEs are achieved when the absorber thicknesses are ≥1200 nm and ETL thicknesses are ≤50 nm. Fig. 6(e) (TiO2 ETL) reaches a maximum PCE of approximately 20.41% when absorber thicknesses are equal to 1200 nm and ETL thicknesses are less than or equal to 50 nm. Fig. 6(f) (ZnO ETL) exhibits a maximum PCE of 20.30% where the thicknesses are equal to 1400 nm and ETL thicknesses are equal to 450 nm. Based on Fig. 6, WS2 is the best-performing ETL layer, achieving the highest maximum PCE of approximately 22.76%. This suggests that WS2, in combination with the Rb2ScCl6 perovskite absorber, facilitates the most efficient charge transfer and transport among the materials tested. WS2 performs best because its energy band alignment with the Rb2AuScCl6 absorber is optimal for efficient electron extraction. Its conduction band level is ideally positioned to collect electrons while preventing hole injection, thereby reducing recombination losses. Additionally, WS2's high electron mobility and layered structure facilitate rapid charge transport, ensuring that the generated carriers are efficiently transferred through the device, which in turn maximizes the power conversion efficiency.52


image file: d5ra07344h-f6.tif
Fig. 6 Contour mapping of PCE when absorber is Rb2ScCl6 and ETLs as (a) C60, (b) IGZO, (c) PCBM, (d) WS2, (e) TiO2 and (f) ZnO.

3.4. Impact of absorber thickness with best ETL

Absorber thickness had a large impact on the performance of ITO/ETL/Rb2AuScBr6/CBTS/Ni and ITO/ETL/Rb2AuScCl6/CBTS/Ni structures. The absorber thickness was varied from 400 nm to 2000 nm to improve the efficiency of devices and the effect has been studied on performance parameters. All the results obtained in the PSC showed that the Voc tended to decrease as the absorber thickness increased. This drop is accompanied by the increase of reverse saturation current, resulting in charge recombination losses.53,54 The Rb2AuScBr6-based device exhibited a decrease in Voc from 1.299 V at 400 nm to 1.25 V at 2000 nm; however, Rb2AuScCl6 remained at a more stable Voc of 1.44–1.46 V. As illustrated in Fig. 7(a) Jsc increased as a result of improved light absorption with increasing absorber thickness. This trend is explained by the better spectral response at longer wavelengths for thicker absorbers. The Jsc increased from 18.46 mA cm−2 (400 nm) to 21.87 mA cm−2 (2000 nm) for the Rb2AuScBr6-based structure and from 15.65 mA cm−2 (600 nm) to 16.46 mA cm−2 (2000 nm for the Rb2AuScCl6).
image file: d5ra07344h-f7.tif
Fig. 7 Effect of absorber thickness on performance parameters (a) Voc, and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

In Fig. 7(b), the Rb2AuScBr6-based device exhibited FF of up to 88.4–88.5%, while that of Rb2AuScCl6 stayed constant at 88.4–89.1%. With respect to power conversion efficiency (PCE), both materials exhibit an increase in PCE with increasing absorber thickness, peaking and subsequently declining slightly due to recombination losses.

For Rb2AuScBr6, the highest was 24.08% at 2000 nm, and Rb2AuScCl6 reached 21.07% at 1600 nm. Although an optimized absorber thickness of 1600 nm for both Rb2AuScBr6 and R2AuScCl6 results in the highest device performance, a thinner absorber layer is generally more desirable for perovskite solar cells (PSCs).55 Therefore, an absorber thickness of 1200 nm was selected as the optimal design choice for this study.

3.5. Influence of metal back contact on solar cell performance

Fig. 8 illustrates the influence of the metal back contact's work function (WF) on key performance parameters Voc, (Jsc), PCE, and FF for two device configurations: ITO/IGZO/Rb2AuScBr6/CBTS/Metal and ITO/WS2/Rb2AuScCl6/CBTS/Metal. In the case of the Rb2AuScBr6-based device, the highest performance is achieved with a work function of 5.5 eV, where the device exhibits a Voc of 1.36 V, Jsc of 21.42 mA cm−2, FF of 87.14%, and a maximum PCE of 25.3%. Notably, enhancing the work function beyond 5.5 eV does not significantly affect the PCE, indicating stability in device performance. Due to this optimized performance and the favorable alignment with the Ni back contact (WF = 5.5 eV), nickel is selected as the preferred back electrode for this device. For the Rb2AuScCl6-based device, the highest PCE of 22.48% is observed at a work function of 5.8 eV, with corresponding values of Voc = 1.53 V, Jsc = 16.15 mA cm−2, and FF = 90.78%.
image file: d5ra07344h-f8.tif
Fig. 8 Effect of Metal Back Contact on performance parameters (a) Voc, and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

However, metals with a work function of 5.8 eV, such as platinum (Pt), are highly expensive and thus less viable for commercial use. Therefore, nickel (WF = 5.5 eV) is adopted as a cost-effective alternative. With Ni as the back contact, the device still demonstrates respectable performance, achieving a PCE of 20.45%, Voc of 1.53 V, Jsc of 14.83 mA cm−2, and FF of 90.43%.

In summary, the feasibility and cost-effectiveness of using nickel as the back contact support its selection for both device architectures.56

3.6. Effect of acceptor density on the solar cell PCE

The impact of acceptor density on PCE has been analyzed for two device architectures: ITO/IGZO/Rb2AuScBr6/CBTS/Ni and ITO/WS2/Rb2AuScCl6/CBTS/Ni. Fig. 9 illustrates variations in Voc, Jsc, FF, and PCE as functions of acceptor density, highlighting efficiency trends in both configurations. From Fig. 9(b), PCE follows a similar trend with the increase of acceptor density, reaches a maximum value, and then decreases steadily. The maximum value of PCE (25.81%) occurs for the Rb2AuScBr6-based device with an acceptor density of 1018 cm−3, at Voc = 1.37 V, Jsc = 21.26 mA cm−2, and FF = 88.94%. Likewise, the Rb2AuScCl6-based device exhibits a maximum PCE of 22.41% at the acceptor density of 1017 cm−3, at Voc = 1.53 V, Jsc = 16.16 mA cm−2, and FF = 90.65%, after which the efficiency drops. The increase in Voc may be attributed to the increased charge carrier separation due to increased internal electric field as a result of increased acceptor density.
image file: d5ra07344h-f9.tif
Fig. 9 Variation of cell parameters with changing acceptor density (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) and (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

Increase in doping concentration contributes to a reduction in series resistance (as such, improving FF) and a reduction in surface recombination losses. In contrast with Voc and FF, Jsc is almost unchanged, suggesting that the doping variations have little impact on light absorption and carrier generation. The results highlight the critical role of optimizing acceptor density in enhancing solar cell efficiency. Among the studied devices, Rb2AuScBr6-based solar cells demonstrated superior performance compared to Rb2AuScCl6 counterparts, which can be attributed to their improved charge transport characteristics and more favorable optical properties.

3.7. Influence of absorber defect density on solar cell behavior

The value of defect density (Nt) in the absorber layer has a great impact on the overall efficiency of solar cell devices.57 Fig. 10(a) and (b) of devices based on Rb2AuScBr6 and Rb2AuScCl6, respectively. The characteristic of increasing Nt leads to the reduction of Voc, Jsc, FF, and PCE in both devices. Under these conditions, the best PCE of 27.11% can be obtained for the Rb2AuScBr6-based device with the parameters of Voc = 1.44 V, Jsc = 21.42 mA cm−2, and FF = 87.88%. This high efficiency is maintained when the defect density in the absorber layer is at 1010 cm−3. In a similar, the Rb2AuScCl6-based device reaches its highest PCE = 22.96% with Voc = 1.54 V, Jsc = 16.29 mA cm−2, and FF = 91.53%.
image file: d5ra07344h-f10.tif
Fig. 10 Effect of absorber defect density (Nt) on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

Although the highest PCEs of 27.11% for the Rb2AuScBr6-based absorber and 22.96% for the Rb2AuScCl6-based absorber were obtained at a defect density of 1010 cm−3, such a low defect density is not considered realistic or acceptable. In practical conditions, the defect density cannot be reduced below 1014 cm−3 and must at least be maintained at this level. At a more acceptable defect density of 1015 cm−3,58 the Rb2AuScBr6-based absorber achieves a PCE of 23.90%, while the Rb2AuScCl6-based absorber reaches a PCE of 20.99%.

3.8. Influences of electron affinity on solar cell behavior

Solar cell efficiency is influenced by the electron affinity of the absorber layer. Fig. 11(a) and (b) depict the variations of Voc, Jsc, FF, and PCE with electron affinity for the Rb2AuScBr6 and Rb2AuScCl6-based devices, respectively.
image file: d5ra07344h-f11.tif
Fig. 11 Effect of absorber electron affinity on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

For the Rb2AuScBr6-based device, the optimal electron affinity (3.95) eV, producing the maximum PCE of 23.12% with Voc = 1.26 V, Jsc = 21.41 mA cm−2, and FF = 88.65%. When the electron affinity becomes larger than this optimal value, the efficiency declines due to a band misalignment, with a corresponding loss of Voc and FF. However, with further increasing EA at 4.50 eV, PCE decreases sharply to 15.93%, highlighting the negative influences of excessive electron affinity on carrier extraction.

For the device based on Rb2AuScCl6, the peak PCE of 21.06% is achieved at 3.70 eV, with Voc = 1.45 V, Jsc = 16.29 mA cm−2, and FF = 88.94%. Above this region, the efficiency begins to decrease, with a serious drop at 4.30 eV and 4.50 eV, with the value of PCE being as low as 8.54% and 6.33%, respectively. The diminished performance at increased electron affinity can be attributed to larger recombination losses and diminished built-in potential, which compromise charge separation and collection efficiency. Thus, tuning the electron affinity for the absorber is crucial for achieving higher cell efficiency. These findings confirm that Rb2AuScBr6 is indeed a better candidate for the development of efficient devices than Rb2AuScCl6, because its superior electrical properties allow for boosting charge transport and decreasing recombination losses.

This improvement can be attributed to the better light intensity of Rb2AuScBr6 caused by the optimal band gap, which leads to better light absorption and efficient excitation of charge carriers.59,60 It achieves a better conduction-band alignment to maintain a matched band gap, improving the tandems Voc, Jsc, and FF balance, and therefore higher efficiencies are obtained than for Rb2AuScCl6. To put it plainly, Rb2AuScBr6 is more efficient at absorbing light, produces a lot of more beneficial charge carriers, and minimizes energy losses, which all contribute to it being better suited for high-performance solar cells.

3.9. Influences of electron mobility on solar cell performance

In contrast, the power conversion efficiency (PCE) exhibits a slight upward trend, particularly at lower µe values. For Rb2AuScCl6, the PCE increases from 20.73% to 21.00% as µe reaches 90 cm2 V−1 s−1, while for Rb2AuScBr6, it rises from 23.00% to 23.92% within the same range. This improvement is attributed to enhanced charge-transport capability, reduced series resistance, and suppression of electron accumulation at interfaces, which collectively lower recombination losses and improve carrier extraction.

Fig. 12(a) and (b) shows the effect of electron mobility (µe) in the range of 10–100 cm2 V−1 s−1 on the photovoltaic performance of Rb2AuScCl6-and Rb2AuScBr6-based perovskite solar cells. The short-circuit current density (Jsc) and open-circuit voltage (Voc) remain largely unchanged with increasing µe, indicating stable charge generation and extraction across the studied mobility range.


image file: d5ra07344h-f12.tif
Fig. 12 Effect of absorber electron mobility on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

As µe approaches 90 cm2 V−1 s−1, the PCE reaches a near-saturation region, indicating that carrier transport becomes sufficiently efficient and further mobility enhancement yields minimal benefit. Beyond this point, the device performance is primarily limited by interfacial recombination, optical absorption constraints, and hole-transport efficiency rather than electron mobility alone.

Overall, electron mobility has a positive yet bounded influence on device performance.61 An optimal mobility window of 80–90 cm2 V−1 s−1 ensures maximum PCE, beyond which additional increases in µe do not result in notable efficiency gains.

3.10. Effect of series resistance (Rs) on solar cell performance

The effect of Rs (series resistance) is a significant factor in solar cell efficiency. Rs was varied from 1 Ω to 8 Ω cm2 in order to investigate its effect on output parameters while maintaining shunt resistance (Rsh) constant at a value of 105 Ω cm2, as shown in Fig. 13(a) and (b). Jsc, on the other hand, is relatively insensitive to series resistance; Jsc remains almost unchanged, but Voc and FF gradually decrease as Rs increases due to more significant resistive losses.62 For devices based on Rb2AuScBr6, the maximum PCE of 23.90% is achieved with Rs = 1 Ω, Voc = 1.26 V, Jsc = 21.41 mA cm−2, and FF = 88.48%. In a similar trend for Rb2AuScCl6-based devices, a maximum PCE of 20.99% occurs at Rs = 1 Ω (Voc = 1.45 V, Jsc = 16.29 mA cm−2, and FF = 89.03%). With higher Rs, PCE decreases due to a fall in FF and a small decrease in Voc. The increased resistive losses and suppressed charge transport caused by a higher Rs result in larger resistive losses and lower power output. Nevertheless, Rb2AuScBr6-based devices show higher PCE than their Rb2AuScCl6 counterparts, revealing better charge carrier mobility and reduced recombination losses.63,64 We found that reducing Rs is critical for improving solar cell performance, as lower resistance in series leads to better efficiency and improved charge extraction.65 Additionally, a higher Rs increases the voltage drop across the device under forward bias, which limits the charge extraction capability and accelerates recombination within the absorber and interfaces. This results in diminished diode quality and reduced operational stability, particularly under high-intensity illumination or prolonged operation. Therefore, maintaining a low Rs is essential not only for maximizing PCE but also for ensuring long-term device reliability and efficient carrier transport pathways.
image file: d5ra07344h-f13.tif
Fig. 13 Effect of series resistance (Rs) on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

3.11. Effect of shunt resistance (Rsh) on solar cell performance

Shunt resistance (Rsh) is an important parameter for the overall performance of perovskite solar cells (PSCs). Low Rsh causes power losses mainly resulting from leakage current and non-geminate recombination losses. These losses happen due to partial junction shorts that arise from pinholes and metal filling reaching the junctions. In this case, low Rsh (parallel resistance) allows for a bypass of the solar cell itself and hence reduces voltage (Voc) and calculated current flow through the junction. Rsh plays a more important role at low light due to the contribution of current loss being dominant when the light-generated current is already at low levels.66,67 To reduce these losses, a variety of techniques have been employed to increase Rsh, such as doping the ETL with lead (Pb) to increase resistance further. The Rsh value can be, therefore, conveniently amended for making good device efficiency when the Pb content is carefully adjusted. Voc, Jsc, FF, and PCE are analyzed for Rb2AuScBr6 and Rb2AuScCl6-based solar cells—keeping Rsh in the range of 10 to 108 Ω, while RS was fixed throughout at 1 Ω cm2.

The data are summarized in Fig. 14(a) and (b). The maximum PCE of Rb2AuScBr6-based devices reached 23.90% at Rsh = 106 Ω, with Voc = 1.26 V, Jsc = 21.41 mA cm−2, and FF = 88.48%. In the same manner, for Rb2AuScCl6-based devices, 20.99% of the maximum PCE occurs at Rsh = 107 Ω, with Voc = 1.45 V, Jsc = 16.29 mA cm−2 and FF = 89.03%. With the increase of Rsh, Voc and FF is effectively increased, significantly improving the PCE. But it plateaus above 106 Ω (Rb2AuScBr6) and 107 Ω (Rb2AuScCl6), demonstrating that additional increases in Rsh have an insignificant influence on performance.


image file: d5ra07344h-f14.tif
Fig. 14 Effect of shunt resistance (Rsh) on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

These data underscore that Rsh is important for optimizing PSC efficiency. High Rsh (≥104 Ω) is crucial for reducing leakage currents and recombination losses to guarantee better charge collection and enhanced device performance.68,69 The improved performance seen in Rb2AuScBr6-based devices demonstrates that with higher charge transport ability and minimized recombination losses, Rb2AuScBr6-based devices can outperform their Rb2AuScCl6 analogs. For further insights to better understand the behavior of Rsh and Rs, an ideal single-diode model has been described using eqn (11).

 
image file: d5ra07344h-t16.tif(11)
where A stands for ideality factor, kB for Boltzmann constant, T for temperature, q for electron charge, and M = V + (I × Rs). The opposite saturation current of the diode is I0, while the light-induced current is IL. Eqn (12) and (13) are the Rs and Rsh equations, respectively, derived from the JV curves, in which ΔV stands for voltage change and ΔI for current change.
 
image file: d5ra07344h-t17.tif(12)
 
image file: d5ra07344h-t18.tif(13)

As described in eqn (11), the term image file: d5ra07344h-t19.tif represents the parasitic leakage current path, meaning that a lower Rsh increases the portion of current that bypasses the junction, thereby reducing the net photocurrent and weakening the diode characteristics. When Rsh is sufficiently high, this leakage component becomes negligible, ensuring efficient charge extraction and sustaining higher Voc and FF. Therefore, maintaining a large Rsh is critical for suppressing shunt-induced recombination losses and achieving ideal diode behavior and superior device performance.

3.12. Role of temperature on solar cell performance

Temperature causes a large change in Voc, Jsc, FF, and PCE of PSCs, influencing the efficiency of PSCs. A high temperature can cause Voc and FF to drop, leading to a corresponding drop in PCE. This decrease is due to increased recombination of charge carriers, thermally activated effects, and changes in material properties that limit device performance. In PSCs, Jsc shows almost no temperature dependence, revealing that the absorption properties of the absorber layers are not affected by the temperature.70 The temperature (300–440 K) effects on Voc, Jsc, FF, and PCE of both Rb2AuScBr6 and Rb2AuScCl6-based solar cells are presented in Fig. 14(a) and (b). The maximum PCE of Rb2AuScBr6-based devices is up to 23.901% at 300 K with Voc = 1.262 V, Jsc = 21.41 mA cm−2, and FF = 88.475%. Likewise, the PCE of Rb2AuScCl6-based devices is 21.013% at 300 K, with Voc = 1.449 V, Jsc = 16.289 mA cm−2, and FF = 89.036%. This ensures a negative impact of higher temperatures on the solar cell performance, as well as a positive impact of lower operating temperature on efficiency.71

3.13. Impact of capacitance, Mott–Schottky, generation and recombination rate

The C–V behavior of Rb2AuScBr6 and Rb2AuScCl6-based PSCs is shown in Fig. 15(a) An increased capacitance with voltage indicates charge accumulation at the interface. Capacitance for Rb2AuScBr6 is much higher than Rb2AuScCl6, suggesting a larger charge storage capability of Rb2AuScCl6 along with enhanced charge carrier density. This implies that Rb2AuScCl6 possesses enhanced dielectric behavior and more efficient charge transport, factors that may be responsible for its corresponding higher power conversion efficiency (PCE). The junction capacitance per area was determined using a known relation:72
 
image file: d5ra07344h-t20.tif(14)
where ε0 is vacuum permittivity, εr is the dielectric constant, q is electronic charge, and A is the area of the device. The value of Nd is derived from the gradient of the linear region in the M–S plot, while Vbi is the extrapolated segment from the linear segment to the voltage axis.

image file: d5ra07344h-f15.tif
Fig. 15 Effect of temperature on performance parameters (a) Voc and Jsc, (b) PCE and FF of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) and (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

As seen in Fig. 16(b), the Mott–Schottky plot (1/C2 vs. Voltage) shows that the equation gives the information on the built-in potential and the carrier density. The flatter slope in the Rb2AuScCl6 device corresponds to higher Nd, which improves the internal electric field distribution, leading to better carrier transport and reduced recombination losses. Notably, the built-in potential V is higher in the Rb2AuScBr6 cell, supporting stronger band bending and more efficient charge separation at the junction. Fig. 16(c) and (d) demonstrate the spatial profiles of the carrier generation and recombination rates within both device structures. The generation rate profile reveals that Rb2AuScBr6-based cells exhibit a peak carrier generation rate of ∼1022 cm−3 s−1, slightly exceeding that of the Rb2AuScCl6-based cells, which peak around ∼1021 cm−3 s−1. This behavior is described quantitatively by the following relation:

 
G(λ,x) = a(λ,x)Nphot(λ,x) (15)
Here, G(λ, x) represents the generation rate of electron–hole pairs at a given wavelength λ and depth x within the absorber. The term α(λ, x) is the absorption coefficient, which reflects how strongly the material absorbs light at different wavelengths and depths, while Nphot(λ, x) denotes the photon flux—the number of incident photons available for absorption at that wavelength and position. This equation shows that carrier generation is directly influenced by both the material's optical absorption properties and the incident light intensity. In this context, although Rb2AuScBr6 shows a slightly higher peak generation rate, the favorable absorption behavior of Rb2AuScCl6 enables it to harness a wider portion of the solar spectrum effectively, resulting in substantial exciton generation and contributing to an elevated Jsc. Fig. 16(d) shows that the Rb2AuScCl6-based device exhibits significantly lower recombination rates throughout both the absorber layer and the interfaces, compared to the Rb2AuScBr6-based device. This lower recombination indicates that fewer charge carriers are lost before contributing to current, suggesting longer carrier lifetimes and less defect-induced trapping within the material. These advantages are likely due to better crystal quality, optimized interface alignment, and reduced trap densities in Rb2AuScCl6. As a result, although both devices are capable of generating charge carriers efficiently under illumination, the reduced recombination in Rb2AuScCl6 allows more carriers to reach the electrodes and contribute to the output power. Combined with its strong dielectric properties and higher concentration of mobile carriers, these factors lead to a noticeable improvement in key photovoltaic parameters such as the Voc, FF, and overall PCE. Therefore, the Rb2AuScCl6-based perovskite solar cell demonstrates greater promise for high-performance and stable lead-free solar technologies.73


image file: d5ra07344h-f16.tif
Fig. 16 (a) Capacitance, (b) Mott–Scotty plot (1/C2), (c) generation, and (d) recombination of (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) and (ITO/WS2/Rb2AuScCl6/CBTS/Ni) devices.

3.14. Current density vs. voltage (JV) and quantum efficiency (QE) curve

Fig. 17(a) and (b) shows the JV characteristics of Rb2AuScBr6 and Rb2AuScCl6 based Absorber investigated perovskite solar cell structures are presented with voltage varying from 0 to 1.50 V; from which the Rb2AuScBr6-based PSC initially shows a current density of nearly 21.41 mA cm−2 and finally shows a current density of nearly 21.20 mA cm−2 the Rb2AuScCl6-based device presents initial Jsc of ∼16.26 mA cm−2 and final Jsc of 16.13 mA cm−2. Within this analysis, we observe that compared to the initial value slightly decrease final value of Jsc. The initial Voc is 1.28 V for Rb2AuScBr6 and 1.46 V for Rb2AuScCl6, and final Voc 1.44 for Rb2AuScBr6 and 1.56 for Rb2AuScCl6 as shown in Fig. 17(a) and (b).
image file: d5ra07344h-f17.tif
Fig. 17 (a and b) JV characteristics and (c and d) QE curve of Rb2AuScBr6 and Rb2AuScCl6-baesd absorber initial and final devices.

The red curve exhibits an improved and delayed current drop, indicating enhanced charge separation and reduced carrier recombination. Although the maximum current density is slightly lower than the initial case, the extended voltage range suggests an overall improvement in device efficiency. The final curve exhibits an improved and delayed current drop, indicating enhanced charge separation and reduced carrier recombination. Although the maximum current density is slightly lower than the initial case, the extended voltage range suggests an overall improvement in device efficiency.74 Their maximum QE initial value is 99.91% and final value 99.81% measured at 360 nm for Rb2AuScBr6 and QE initial value 99.99% and final value 99.84% at 410 nm for Rb2AuScCl6 across both curves. In the visible spectrum (especially in the range of 360–450 nm), optimal QE of 100% is achieved for both materials. It is important to highlight that (as shown in Fig. 17(c) and (d)), QE reaches near 100% for both initial and final curves in most of the visible spectrum region, although after the wavelength of 700 nm for (Rb2AuScBr6) and 650 nm for (Rb2AuScCl6), QE drops dramatically.

However, there is a decrease in the QE for both solar cells at longer wavelengths due to recombination, which occurs when charge carriers do not pass into an external circuit. The same processes that influence the collection probability also affect QE. For example, changes to the front surface can affect the carriers that are produced near the surface. High doping on the front surface layers can lead to free carrier absorption, which also has an increase in absorption, leading to a decrease in the QE in longer wavelengths.75

3.15. Final devices after optimization

Initially, the devices were configured as (ITO/IGZO/Rb2AuScBr6/CBTS/Ni) and (ITO/WS2/Rb2AuScCl6/CBTS/Ni). For the Br-based absorber, IGZO exhibited superior electron transport performance with a PCE of 23.90%. Upon optimizing absorber layer parameters—including acceptor density, electron affinity, defect density, series resistance, shunt resistance, and temperature—the highest PCE of 25.81% was achieved at an acceptor density of 1018 cm−3 using IGZO as the ETL.

Further improvement was observed by replacing IGZO with alternative ETLs, when the acceptor density was maintained at 1018 cm−3, TiO2, ZnO, and WS2 offered enhanced PCEs of 27.49%, 27.37%, and 27.30%, respectively. Among them, TiO2 demonstrated the best performance, leading to the final optimized device structure for the Br-based absorber:

Glass/ITO (0.500 µm)/TiO2 (0.030 µm)/Rb2AuScBr6 (1.200 µm)/CBTS (0.100 µm)/Ni (5.5 eV), with a PCE of 27.49% at a defect density of 1015 cm−3.

On the other hand (Fig. 18), the Cl-based absorber initially showed the best performance with WS2 as the ETL, yielding a PCE of 20.98%. After acceptor density optimization (1017 cm−3), the PCE improved to 22.41% at a defect density of 1015 cm−3. This confirmed the final optimized device configuration as: Glass/ITO (0.500 µm)/WS2 (0.030 µm)/Rb2AuScCl6 (1.200 µm)/CBTS (0.100 µm)/Ni (5.5 eV).


image file: d5ra07344h-f18.tif
Fig. 18 The final device configuration of the absorber (a) Rb2AuScBr6 and (b) Rb2AuScCl6-based PSC.

3.16. Comparison with other works

In comparison with previously reported theoretical studies on lead-free perovskite solar cells, the present work demonstrates outstanding photovoltaic (PV) performance, particularly in devices based on the Rb2AuScBr6 and Rb2AuScCl6 perovskite absorbers. The optimized structure ITO/TiO2/Rb2AuScBr6/CBTS/Ni achieves a maximum PCE of 27.49%, with a Voc of 1.428 V, (Jsc) of 21.280 mA cm−2, and FF of 89.951%. Similarly, structures incorporating ZnO and WS2 as ETLs, ITO/ZnO/Rb2AuScBr6/CBTS/Ni and ITO/WS2/Rb2AuScBr6/CBTS/Ni—achieved high PCEs of 27.37% and 27.30%, respectively. These performances significantly surpass those of other reported lead-free configurations. For instance, devices such as ITO/PCBM/KgeCl3/CBTS/Ni76 and FTO/C60/KgeCl3/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/Ni77 exhibited lower efficiencies of 21.79% and 20.46%, respectively. Furthermore, structures like FTO/ZnO/KsnI3/CuI/Au78 PCE of (20.99%), ITO/IGZO/CsSnCl3/CBTS/Au PCE of (21.07%), and ITO/WS2/CsSnCl3/CBTS/Au79 PCE of (21.32%) also reported considerably lower PCE values. Even the FTO/PCBM/CsSnCl2/PTAA/Au80 configuration yielded a much lower efficiency of 17.93%. Notably, the Rb2AuScCl6-based device.

ITO/WS2/Rb2AuScCl6/CBTS/Ni only achieved a PCE of 22.41%, which is substantially lower than that of the Rb2AuScBr6-based devices. These results clearly indicate that the combination of IGZO, TiO2, or ZnO as ETLs, CBTS as the HTL, and Rb2AuScBr6 as the absorber forms a highly promising structure for next-generation, high-efficiency, lead-free perovskite solar cells (Table 4).

Table 4 The comparison of PV parameters of Rb2AuScBr6-and R2AuScCl6-based solar cells
Optimized devices Types Voc (V) Jsc (mA cm−2) FF (%) PCE (%) Ref.
ITO/TiO2/Rb2AuScBr6/CBTS/Ni Theo 1.428 21.280 89.951 27.49 This work
ITO/ZnO/Rb2AuScBr6/CBTS/Ni Theo 1.428 21.209 90.342 27.37 This work
ITO/WS2/Rb2AuScBr6/CBTS/Ni Theo 1.426 21.280 89.951 27.30 This work
ITO/WS2/Rb2AuScCl6/CBTS/Ni Theo 1.210 22.517 83.055 22.41 This work
ITO/PCBM/KgeCl3/CBTS/Ni Theo 0.6745 41.405 78.02 21.79 72
FTO/C60/KgeCl3/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/Ni Theo 0.632 41.13 78.71 20.46 73
FTO/ZnO/KSnI3/CuI/Au Theo 1.44 17.06 85.24 20.99 74
ITO/IGZO/CsSnCl3/CBTS/Au Theo 1.02 26.14 78.69 21.07 30
ITO/WS2/CsSnCl3/CBTS/Au Theo 1.01 25.92 81.33 21.32 30
FTO/PCBM/CsSnCl3/PTAA/Au Theo 1.30 15.34 89.90 17.93 75


4. Conclusion

This study systematically optimized lead-free PSCs employing Rb2AuScBr6 and Rb2AuScCl6 absorbers. Initial screening identified ITO/IGZO/Rb2AuScBr6/CBTS/Ni and ITO/WS2/Rb2AuScCl6/CBTS/Ni as the best-performing baseline architectures. Device performance was then enhanced through absorber thickness optimization, back-contact work function tuning, doping concentration adjustment, and defect density control.

For Rb2AuScBr6-based PSCs, an optimal absorber thickness of ∼1200 nm balanced light absorption with charge transport, achieving a peak PCE of 25.81% at an acceptor density of 1018 cm−3 and a defect density of 1015 cm−3. Increasing the back-contact work function to 5.5 eV further improved efficiency. Substituting the IGZO ETL with TiO2, ZnO, or WS2 under these optimized conditions yielded maximum PCEs of 27.49%, 27.37%, and 27.30%, respectively, with TiO2 performing best.

For Rb2AuScCl6-based PSCs, WS2 consistently provided superior performance across all optimization stages. The highest recorded efficiency was 22.41% at an acceptor density of 1017 cm−3 and defect density of 1015 cm−3. Both material systems demonstrated minimal performance degradation under variations in series resistance, shunt resistance, and operating temperature, confirming strong electrical and thermal stability.

The comparative analysis reveals that Rb2AuScBr6 offers higher ultimate efficiency potential, while Rb2AuScCl6 provides competitive performance with robust ETL compatibility. The environmentally friendly composition of both absorbers, combined with high efficiency and operational stability, underscores their suitability as sustainable alternatives to toxic lead-based perovskites.

Given their promising efficiency, thermal stability, and non-toxic composition, these devices are particularly well-suited for building-integrated photovoltaics (BIPV), where safety, stability, and aesthetic flexibility are crucial, as well as for next-generation lightweight solar modules for portable and off-grid applications. The findings emphasize that careful device engineering, especially ETL selection, doping control, and absorber layer optimization, can enable lead-free PSCs to achieve performance levels competitive with their lead-based counterparts, paving the way for their integration into large-scale, eco-friendly solar energy systems.

Author contributions

Md. Al-Amin: writing – original draft, data calculations, methodology, validation. A. Haque: writing – review & editing, investigation, validation. S. Mahmud: writing – review & editing, methodology, conceptualization, formal analysis, validation. M. M. Hossain: writing – review & editing, validation. M. M. Uddin: writing – review & editing, validation. M. A. Ali: writing – review & editing, supervision, project administration, validation.

Conflicts of interest

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

Data availability

The datasets generated and/or analyzed during the current study are available within the article, and the raw data are available from the corresponding author upon reasonable request.

References

  1. J. S. Riti and Y. Shu, Renewable energy, energy efficiency, and eco-friendly environment (R-E5) in Nigeria, Energy Sustain. Soc., 2016, 6(1), 13 CrossRef.
  2. S. Rabhi and et al., Rethinking Solar Energy: Innovations in Eco-friendly Materials, in Breaking Boundaries: Pioneering Sustainable Solutions through Materials and Technology, Singapore, Springer Nature Singapore, 2025, pp. 3–15 Search PubMed.
  3. A. O. Maka and J. M. Alabid, Solar energy technology and its roles in sustainable development, Clean Energy, 2022, 6(3), 476–483 CrossRef.
  4. I.-A. Vlad and A. Drăgulinescu, High Efficiency Perovskite Solar Cells Optimization, in 2023 IEEE 29th International Symposium for Design and Technology in Electronic Packaging (SIITME), IEEE, 2025, pp. 102–105 Search PubMed.
  5. M. K. Hossain, et al., Design and Simulation of Cs2BiAgI6 Double Perovskite Solar Cells with Different Electron Transport Layers for Efficiency Enhancement, Energy Fuels, 2023, 37(5), 3957–3979 CrossRef CAS.
  6. F. Ji, G. Boschloo, F. Wang and F. Gao, Challenges and Progress in Lead-Free Halide Double Perovskite Solar Cells, Sol. RRL, 2023, 7(6), 2201112 CrossRef CAS.
  7. M. H. Miah, et al., Lead-free alternatives and toxicity mitigation strategies for sustainable perovskite solar cells: a critical review, Mater. Adv., 2025, 6, 2718–2752 RSC.
  8. R. Ahmad, et al., Colloidal lead-free Cs2AgBiBr6 double perovskite nanocrystals: Synthesis, uniform thin-film fabrication, and application in solution-processed solar cells, Nano Res., 2021, 14(4), 1126–1134 CrossRef CAS.
  9. M. I. Khan, et al., Bandgap reduction and efficiency enhancement in Cs2AgBiBr6 double perovskite solar cells through gallium substitution, RSC Adv., 2024, 14(8), 5440–5448 RSC.
  10. G. Volonakis, et al., Lead-Free Halide Double Perovskites via Heterovalent Substitution of Noble Metals, J. Phys. Chem. Lett., 2016, 7(7), 1254–1259 CrossRef CAS PubMed.
  11. M. A. Rayhan, M. M. Hossain, M. M. Uddin, S. H. Naqib, and M. A. Ali, DFT Exploration of Novel Direct Band Gap Semiconducting Halide Double Perovskites, A2AgIrCl6 (A = Cs, Rb, K), for Solar Cells Application, 2024 Search PubMed.
  12. J. F. R. V. Silveira and J. L. F. Da Silva, Mixed Halide Lead-free Double Perovskite Alloys for Band Gap Engineering, ACS Appl. Energy Mater., 2020, 3(8), 7364–7371 CrossRef CAS.
  13. M. T. Iqbal, et al., Next-Generation Materials Discovery Using DFT: Functional Innovation, Sol. Energy Catal. Eco Toxic. Model. Sch J Eng Tech, 2025, 7, 454–486 Search PubMed.
  14. S. A. U. Wasti and et al., Optimization and Performance Evaluation of Cs2CuBiCl6 Double Perovskite Solar Cell for Lead-Free Photovoltaic Applications, 2025 Search PubMed.
  15. H. Chen, et al., Lead Sequestration in Perovskite Photovoltaic Device Encapsulated with Water-Proof and Adhesive Poly (ionic liquid), ACS Appl. Mater. Interfaces, 2023, 15(10), 13637–13643 CrossRef CAS PubMed.
  16. M. K. S. Bin Rafiq, et al., WS2: a new window layer material for solar cell application, Sci. Rep., 2020, 10(1), 771 CrossRef CAS PubMed.
  17. M. E. Islam, M. R. Islam, S. Ahmmed, M. K. Hossain and M. F. Rahman, Highly efficient SnS-based inverted planar heterojunction solar cell with ZnO ETL, Phys. Scr., 2023, 98(6), 065501 CrossRef CAS.
  18. P. Patil, D. S. Mann, U. T. Nakate, Y.-B. Hahn, S.-N. Kwon and S.-I. Na, Hybrid interfacial ETL engineering using PCBM-SnS2 for High-Performance pin structured planar perovskite solar cells, Chem. Eng. J., 2020, 397, 125504 CrossRef CAS.
  19. X. Zhang, et al., Amorphous TiO2 film with fiber-like structure: An ideal candidate for ETL of perovskite solar cells, Mater. Lett., 2022, 324, 132684 CrossRef CAS.
  20. J. Li, S. Kundu, L. Souriau, P. Bezard, R. Izmailov and F. Lazzarino, Plasma Etch of IGZO Thin Film and IGZO/SiO2 Interface Diffusion in Inductively Coupled CH4/Ar Plasmas, Plasma Process. Polym., 2025, 22(2), 2400186 CrossRef CAS.
  21. A. Mohandes and M. Moradi, Improved performance of inorganic CsPbI3 perovskite solar cells with WO3/C60 UTL bilayer as an ETL structure: A computational study, Phys. Scr., 2024, 99(5), 055951 CrossRef CAS.
  22. G. Pindolia, S. M. Shinde and P. K. Jha, Optimization of an inorganic lead free RbGeI3 based perovskite solar cell by SCAPS-1D simulation, Sol. Energy, 2022, 236, 802–821 CrossRef CAS.
  23. I. Montoya De Los Santos, et al., The study of inorganic absorber layers in perovskite solar cells: the influence of CdTe and CIGS incorporation, Sci. Rep., 2025, 15(1), 10353 CrossRef CAS PubMed.
  24. A. Ghosh, et al., Solar power conversion: CuI hole transport layer and Ba3NCl3 absorber enable advanced solar cell technology boosting efficiency over 30%, RSC Adv., 2024, 14(33), 24066–24081 RSC.
  25. Q. Zhao, et al., Achieving efficient inverted planar perovskite solar cells with nondoped PTAA as a hole transport layer, Org. Electron., 2019, 71, 106–112 CrossRef CAS.
  26. C. D. Ramabadran and K. S. Sudheer, A Study on the influence of different HTL and ETL layers on the performance of a P3HT, C60 bulk heterojunction organic solar cell, Indian J. Phys., 2025, 99(1), 273–282 CrossRef CAS.
  27. C.-C. Shih and C.-G. Wu, Synergistic Engineering of the Conductivity and Surface Properties of PEDOT: PSS-Based HTLs for Inverted Tin Perovskite Solar Cells to Achieve Efficiency over 10%, ACS Appl. Mater. Interfaces, 2022, 14(14), 16125–16135 CrossRef CAS PubMed.
  28. M. Al-Hattab, L. Moudou, M. Khenfouch, O. Bajjou, Y. Chrafih and K. Rahmani, Numerical simulation of a new heterostructure CIGS/GaSe solar cell system using SCAPS-1D software, Sol. Energy, 2021, 227, 13–22 CrossRef CAS.
  29. M. Bošnjaković, Advances of sustainable energy materials: technology trends for silicon-based photovoltaic cells, Sustainability, 2024, 16(18), 7962 CrossRef.
  30. S. Jouttijärvi, G. Lobaccaro, A. Kamppinen and K. Miettunen, Benefits of bifacial solar cells combined with low voltage power grids at high latitudes, Renew. Sustain. Energy Rev., 2022, 161, 112354 CrossRef.
  31. Z. Song, C. Li, L. Chen and Y. Yan, Perovskite Solar Cells Go Bifacial—Mutual Benefits for Efficiency and Durability, Adv. Mater., 2022, 34(4), 2106805 CrossRef CAS PubMed.
  32. M. Vishnuwaran, K. Ramachandran and P. Roy, SCAPS simulated FASnI3 and MASnI3 based PSC solar cells: A comparison of device performance, IOP Conf. Ser.: Mater. Sci. Eng., 2025, 012048 Search PubMed.
  33. S. N. Ali, M. B. Zahran and A. A. Z. Diab, Comprehensive comparison between the experimental and simulated characteristics of the mono-crystalline silicon solar cell using SCAPS-1D, Int. J. Sci. Eng. Appl., 2023, 4(2), 165–172 Search PubMed.
  34. M. K. Hossain, et al., An extensive study on multiple ETL and HTL layers to design and simulate high-performance lead-free CsSnCl3-based perovskite solar cells, Sci. Rep., 2023, 13(1), 2521 CrossRef CAS PubMed.
  35. A. T. Ngoupo, S. Ouédraogo, F. Zougmoré and J. M. B. Ndjaka, Numerical analysis of ultrathin Sb2Se3-based solar cells by SCAPS-1D numerical simulator device, Chin. J. Phys., 2021, 70, 1–13 CrossRef CAS.
  36. S. Karthick, S. Velumani and J. Bouclé, Chalcogenide BaZrS3 perovskite solar cells: A numerical simulation and analysis using SCAPS-1D, Opt. Mater., 2022, 126, 112250 CrossRef CAS.
  37. M. G. Hudedmani, V. Soppimath and C. Jambotkar, A study of materials for solar PV technology and challenges, Eur. J. Appl. Eng. Sci. Res., 2017, 5(1), 1–13 CAS.
  38. M. Dada and P. Popoola, Recent advances in solar photovoltaic materials and systems for energy storage applications: a review, Beni-Suef Univ. J. Basic Appl. Sci., 2023, 12(1), 66 CrossRef.
  39. M. V. Dambhare, B. Butey, and S. V. Moharil, Solar photovoltaic technology: A review of different types of solar cells and its future trends, in Journal of Physics: Conference Series, IOP Publishing, 2025, p. 012053 Search PubMed.
  40. S. Mahmud, M. M. Hossain, M. M. Uddin and M. A. Ali, Prediction of X2AuYZ6 (X= Cs, Rb; Z= Cl, Br, I) double halide perovskites for photovoltaic and wasted heat management device applications, J. Phys. Chem. Solids, 2025, 196, 112298 CrossRef CAS.
  41. N. Vukmirović, Calculations of electron mobility in II-VI semiconductors, Phys. Rev. B, 2021, 104(8), 085203 CrossRef.
  42. Y. Zhou and G. Long, Low Density of Conduction and Valence Band States Contribute to the High Open-Circuit Voltage in Perovskite Solar Cells, J. Phys. Chem. C, 2017, 121(3), 1455–1462 CrossRef CAS.
  43. S. S. Bagade, S. B. Barik, M. M. Malik and P. K. Patel, Impact of band alignment at interfaces in perovskite-based solar cell devices, Mater. Today Proc., 2023 DOI:10.1016/j.matpr.2023.02.117.
  44. P. Gupta, Band alignment studies of Zn1-xNixO/ZnO: as bilayer electron transport layer in perovskite solar cells, Opt. Mater., 2024, 150, 115305 CrossRef CAS.
  45. S. Mahmud, M. A. Ali, M. M. Hossain and M. M. Uddin, DFT aided prediction of phase stability, optoelectronic and thermoelectric properties of A2AuScX6 (A= Cs, Rb; X= Cl, Br, I) double perovskites for energy harvesting technology, Vacuum, 2024, 221, 112926 CrossRef CAS.
  46. X. Yu, et al., Numerical simulation analysis of the effect of energy band alignment and functional layer thickness on the performance for perovskite solar cells with Cd1-xZnxS electron transport layer, Mater. Res. Express, 2020, 7(10), 105906 CrossRef CAS.
  47. A. Kumar, D. Punetha and S. Chakrabarti, Performance optimization of phenylethylammonium-formamidinium tin iodide perovskite solar cell by contrasting various ETL and HTL materials, J. Nanoparticle Res., 2023, 25(3), 52 CrossRef CAS.
  48. Md. S. Reza, et al., Optimizing Charge Transport Layers to Enhance the Performance of Lead-Free RbGeI3 Perovskite Solar Cells: A Comprehensive Analysis of ETL and HTL Engineering, Langmuir, 2025, 41(11), 7865–7885 CrossRef CAS PubMed.
  49. Md. S. Reza, et al., Optimizing Charge Transport Layers to Enhance the Performance of Lead-Free RbGeI3 Perovskite Solar Cells: A Comprehensive Analysis of ETL and HTL Engineering, Langmuir, 2025, 41(11), 7865–7885 CrossRef CAS PubMed.
  50. S. Bichave, J. Mundupuzhakal, P. N. Gajjar and S. K. Gupta, Analysis of varying ETL/HTL material for an effective perovskite solar cell by numerical simulation, Mater. Today Proc., 2023, 2025 Search PubMed.
  51. U. C. Obi, D. M. Sanni and A. Bello, Effect of Absorber Layer Thickness on the Performance of Bismuth-Based Perovskite Solar Cells, Semiconductors, 2021, 55(12), 922–927 CrossRef CAS.
  52. N. A. A. Malek, et al., Enhanced Charge Transfer in Atom-Thick 2H–WS2 Nanosheets' Electron Transport Layers of Perovskite Solar Cells, Sol. RRL, 2020, 4(10), 2000260 CrossRef.
  53. Z. Chen and G. Chen, The effect of absorber thickness on the planar Sb2S3 thin film solar cell: Trade-off between light absorption and charge separation, Sol. Energy, 2020, 201, 323–329 CrossRef CAS.
  54. U. C. Obi, D. M. Sanni and A. Bello, Effect of Absorber Layer Thickness on the Performance of Bismuth-Based Perovskite Solar Cells, Semiconductors, 2021, 55(12), 922–927 CrossRef CAS.
  55. A. Bag, R. Radhakrishnan, R. Nekovei and R. Jeyakumar, Effect of absorber layer, hole transport layer thicknesses, and its doping density on the performance of perovskite solar cells by device simulation, Sol. Energy, 2020, 196, 177–182 CrossRef CAS.
  56. R. S. Hall, D. Lamb and S. J. C. Irvine, Back contact materials used in thin film CdTe solar cells—A review, Energy Sci. Eng., 2021, 9(5), 606–632 CrossRef CAS.
  57. M. S. Jamal, et al., Effect of defect density and energy level mismatch on the performance of perovskite solar cells by numerical simulation, Optik, 2019, 182, 1204–1210 CrossRef CAS.
  58. J. Madan, S. Garg, K. Gupta, S. Rana, A. Manocha and R. Pandey, Numerical simulation of charge transport layer free perovskite solar cell using metal work function shifted contacts, Optik, 2020, 202, 163646 CrossRef CAS.
  59. P. Ru, et al., High Electron Affinity Enables Fast Hole Extraction for Efficient Flexible Inverted Perovskite Solar Cells, Adv. Energy Mater., 2020, 10(12), 1903487 CrossRef CAS.
  60. A. V. Mumyatov and P. A. Troshin, A review on fullerene derivatives with reduced electron affinity as acceptor materials for organic solar cells, Energies, 2023, 16(4), 1924 CrossRef CAS.
  61. Q. Guo, et al., Effect of Energy Alignment, Electron Mobility, and Film Morphology of Perylene Diimide Based Polymers as Electron Transport Layer on the Performance of Perovskite Solar Cells, ACS Appl. Mater. Interfaces, 2017, 9(12), 10983–10991 CrossRef CAS PubMed.
  62. K. C. Fong, K. R. McIntosh and A. W. Blakers, Accurate series resistance measurement of solar cells, Prog. Photovolt. Res. Appl., 2013, 21(4), 490–499 CrossRef CAS.
  63. M. Wolf and H. Rauschenbach, Series resistance effects on solar cell measurements, Adv. Energy Convers., 1963, 3(2), 455–479 CrossRef.
  64. K. Rajkanan and J. Shewchun, A better approach to the evaluation of the series resistance of solar cells, Solid-State Electron., 1979, 22(2), 193–197 CrossRef.
  65. J. D. Servaites, S. Yeganeh, T. J. Marks and M. A. Ratner, Efficiency Enhancement in Organic Photovoltaic Cells: Consequences of Optimizing Series Resistance, Adv. Funct. Mater., 2010, 20(1), 97–104 CrossRef CAS.
  66. A. D. Dhass, E. Natarajan, and L. Ponnusamy, Influence of shunt resistance on the performance of solar photovoltaic cell, in 2012 International Conference on Emerging Trends in Electrical Engineering and Energy Management (ICETEEEM), IEEE, 2025, pp. 382–386 Search PubMed.
  67. M. Barbato, M. Meneghini, A. Cester, G. Mura, E. Zanoni and G. Meneghesso, Influence of shunt resistance on the performance of an illuminated string of solar cells: theory, simulation, and experimental analysis, IEEE Trans. Device Mater. Reliab., 2014, 14(4), 942–950 Search PubMed.
  68. M. Fortes, E. Comesana, J. A. Rodriguez, P. Otero and A. J. Garcia-Loureiro, Impact of series and shunt resistances in amorphous silicon thin film solar cells, Sol. Energy, 2014, 100, 114–123 CrossRef CAS.
  69. C. M. Proctor and T.-Q. Nguyen, Effect of leakage current and shunt resistance on the light intensity dependence of organic solar cells, Appl. Phys. Lett., 2015, 106(8), 2025 CrossRef.
  70. P. Singh and N. M. Ravindra, Temperature dependence of solar cell performance—an analysis, Sol. Energy Mater. Sol. Cells, 2012, 101, 36–45 CrossRef CAS.
  71. Q. Meng, et al., Effect of temperature on the performance of perovskite solar cells, J. Mater. Sci. Mater. Electron., 2021, 32(10), 12784–12792 CrossRef CAS.
  72. S. Ravishankar, Z. Liu, U. Rau and T. Kirchartz, Multilayer Capacitances: How Selective Contacts Affect Capacitance Measurements of Perovskite Solar Cells, PRX Energy, 2022, 1(1), 013003 CrossRef.
  73. J. Chen and N. Park, Causes and Solutions of Recombination in Perovskite Solar Cells, Adv. Mater., 2019, 31(47), 1803019 CrossRef CAS PubMed.
  74. S. Mudgal, S. Singh, and V. K. Komarala, Interfacial spectral response under voltage and light bias to analyse low voltage in amorphous-crystalline silicon heterojunction solar cell with S-shape characteristics, in 2018 IEEE 7th World Conference on Photovoltaic Energy Conversion (WCPEC)(a Joint Conference of 45th IEEE PVSC, 28th PVSEC & 34th EU PVSEC), IEEE, 2025, pp. 2158–2161 Search PubMed.
  75. L. Liu, et al., Quantum efficiency and temperature coefficients of GaInP/GaAs dual-junction solar cell, Sci. China, Ser. E: Technol. Sci., 2009, 52(5), 1176–1180 CrossRef CAS.
  76. M. Tarekuzzaman, et al., An in-depth investigation of lead-free KGeCl3 perovskite solar cells employing optoelectronic, thermomechanical, and photovoltaic properties: DFT and SCAPS-1D frameworks, Phys. Chem. Chem. Phys., 2024, 26(43), 27704–27734 RSC.
  77. M. E. Sarhani, T. Dahame, M. L. Belkhir, B. Bentria and A. Begagra, AB-INITIO study of electronic, mechanical, optical and thermoelectric properties of KGeCl3 for photovoltaic application, Heliyon, 2023, 9(9), e19808 CrossRef CAS PubMed.
  78. F. B. Sumona, et al., Optimization of Perovskite-KSnI3 Solar Cell by Using Different Hole and Electron Transport Layers: A Numerical SCAPS-1D Simulation, Energy Fuels, 2023, 37(23), 19207–19219 CrossRef CAS.
  79. S. Srivastava, A. K. Singh, P. Kumar and B. Pradhan, Comparative performance analysis of lead-free perovskite solar cells by numerical simulation, J. Appl. Phys., 2022, 131(17), 175001 CrossRef CAS.
  80. Y. Selmani and L. Bahmad, Computational study of mixed halide perovskites CsSnX3 (X= Br, Cl, mixed halides) for optoelectronic and thermoelectric applications: A DFT and Boltzmann transport methods, Mater. Sci. Eng., B, 2025, 319, 118373 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.