DOI:
10.1039/D5RA05885F
(Paper)
RSC Adv., 2025,
15, 38122-38133
Advancing photovoltaics with Cs2NaInI6-based perovskites: a simulation study on ETL optimization
Received
11th August 2025
, Accepted 30th September 2025
First published on 13th October 2025
Abstract
Developing reliable, energy-efficient, and eco-friendly photovoltaic materials is crucial for advancing next-generation solar technologies. Among lead-free options, double perovskites such as Cs2NaInI6 show strong potential due to their direct bandgap (∼1.6 eV), excellent light absorption, high carrier mobility, and environmental durability. The efficiency of Cs2NaInI6-based perovskite solar cells (PSCs), however, is strongly influenced by the electron transport layer (ETL). In this work, Solar Cell Capacitance Simulator in One Dimension (SCAPS-1D) simulations were employed to analyze ITO/ETL/Cs2NaInI6/Au structures using WS2, SnS2, In2S3, and IGZO as ETLs. Critical factors—absorber thickness, defect density, doping levels, interface traps, and temperature—were systematically tuned to assess their effects on power conversion efficiency (PCE), open-circuit voltage (VOC), short-circuit current density (JSC), and fill factor (FF). Among the tested ETLs, WS2 delivered the best performance with a PCE of 22.63%, VOC of 1.189 V, JSC of 21.406 mA cm−2, and FF of 88.92%, attributed to favorable band alignment, high mobility, and reduced recombination. Additionally, Cs2NaInI6 demonstrated promising thermal and defect stability, emphasizing its viability for real-world applications. Overall, this study underscores the critical role of ETL engineering and provides a simulation-guided approach for designing efficient lead-free perovskite solar cells (PSCs).
1. Introduction
Growing global demand for sustainable energy drives photovoltaic (PV) innovation. Achieving high solar conversion efficiency remains challenging due to intrinsic and extrinsic losses from material imperfections, interfacial recombination, and energy mismatches, all closely linked to the properties of photoactive materials and the design of device architectures.1,2 First-generation silicon-based photovoltaics dominate the market for their efficiency and reliability, but energy-intensive wafer fabrication raises costs, limiting broader adoption despite their commercial success. Second-generation thin-film solar cells use fewer materials and lower temperatures but face sustainability concerns from Cd toxicity and scarce indium reserves.3,4 To overcome these constraints, third-generation solar cell technologies—notably organic photovoltaics, quantum dot solar cells, and perovskite-based devices—have emerged as attractive alternatives.5 Among them, halide perovskites have stimulated widespread academic attention owing to their exceptional power conversion efficiencies (PCEs),6 adjustable bandgaps, strong optical absorption capabilities, and extended charge carrier diffusion lengths. Lead perovskites suffer from instability and toxicity, while double perovskites (A2BB'X6) provide stable, eco-friendly alternatives for sustainable photovoltaic absorber development.7–10 Double perovskites (A2BB'X6), derived from the ABX3 framework with ordered cation substitution, enable tunable structural, electronic, and optical properties for photovoltaics.11 Non-lead variants like Cs2AgBiBr6, Cs2TiBr6, and Cs2SnI6 show promise as eco-friendly alternatives12–14 but face challenges including wide/indirect bandgaps, poor charge transport, and limited defect tolerance, restricting solar cell efficiencies below 2%.15,16
Recently, Cs2NaInI6 has become recognized as a compelling lead-free double perovskite absorber material, distinguished by its desirable optoelectronic and structural properties. It reveals a direct bandgap around 1.6 eV, closely matching the optimal range for single-junction solar cells as defined by the Shockley–Queisser limit. Additionally, it boasts a strong absorption coefficient (∼106 cm−1), excellent charge carrier mobility, and exceptional thermal and moisture stability.17,18 The material's Goldschmidt tolerance factor (∼0.88) ensures a stable cubic structure, and its resistance to oxidation further enhances device longevity under operational stress.19 These attributes make Cs2NaInI6 a highly promising platform for stable, lead-free, and efficient solar energy harvesting. While the photoactive properties of Cs2NaInI6 are well documented, the comprehensive effectiveness of a solar cell strongly relies on its interfacial layers, especially the ETL. The ETL serves an essential function in facilitating charge extraction, suppressing carrier recombination, and maintaining appropriate energy band orientation with the active layer.20–22 Hence, the judicious selection and optimization of the ETL are crucial for achieving superior performance in PSCs.
This research provides a thorough simulation-driven analysis of the photovoltaic behavior of Cs2NaInI6-based PSCs employing four different ETL materials WS2, SnS2, In2S3, and IGZO within a planar heterojunction configuration of ITO/ETL/Cs2NaInI6/Au. The choice of ETL materials (WS2, SnS2, In2S3, and IGZO) was motivated by their favorable electronic and physical compatibility with Cs2NaInI6. WS2 and SnS2 possess high electron mobilities and tunable band alignments that form small positive conduction band offsets with Cs2NaInI6, which helps suppress interfacial recombination while enabling efficient charge extraction.23,24 In2S3, a widely used n-type semiconductor, has been investigated for its benign processing conditions and ability to form relatively stable interfaces with halide perovskites.25 IGZO, an amorphous oxide semiconductor with good transparency and high mobility, is compatible with large-area, low-temperature deposition methods, making it attractive for scalable photovoltaic applications.26,27 Thus, these four candidates were selected to comparatively evaluate their suitability as ETLs for Cs2NaInI6 absorbers. Simulations were performed applying SCAPS-1D, a widely validated numerical tool for assessing the optoelectronic characteristics of multilayer solar cells. The study systematically explores the role of absorber thickness, defect density (both in the bulk and at interfaces), shallow donor/acceptor densities, and transport layer properties on device performance metrics. Notably, WS2 demonstrated the best performance among the evaluated ETLs, followed by SnS2, In2S3, and IGZO, respectively. By highlighting the synergistic role of ETL material selection and device engineering, this work establishes a framework for optimizing Cs2NaInI6-based lead-free perovskite solar cells. The results offer critical insights into interface-dependent performance enhancement and provide a roadmap for future experimental realization of efficient, stable, and environmentally benign perovskite photovoltaics.
2. Methodology
2.1. Simulation utilizing SCAPS-1D
This study employs numerical modeling and simulation of ITO/ETL (WS2, SnS2, In2S3, and IGZO)/Cs2NaInI6/Au PSCs applying the SCAPS-1D. SCAPS-1D is a widely utilized simulation tool in photovoltaic research, known for its ability to accurately replicate experimental performance trends.1,2 The software enables detailed device modeling by supporting up to seven semiconductor layers, six interface regions, and two electrical contacts.28 It computes key photovoltaic output parameters through the computational approach by solving the core equations of semiconductor physics. These include Poisson's equation, the electron and hole continuity equations, and the current density equations, which collectively describe the electrostatic potential and carrier transport dynamics within the device structure.29 The governing equations and their boundary conditions form the basis for simulating the steady-state behavior of the PSC and are elaborated comprehensively in the referenced literature29,30 and mentioned below.| |
 | (1) |
where, ε represents the dielectric constant, n denotes the electron concentration, p stands for the hole concentration, q corresponds to the elementary charge, ND refers to the donor concentration, and NA designates the acceptor concentration.| |
 | (2) |
| |
 | (3) |
Here, JP and Jn represent the current densities of holes and electrons, respectively, while RP and GP denote the rates of hole recombination and generation. Furthermore, Gn indicates the electron generation rate, and Rn corresponds to the electron recombination rate.| |
 | (4) |
| |
 | (5) |
Here, DP and Dn denote the diffusion coefficients for holes and electrons, respectively, while μP and μn represent the mobilities of holes and electrons.
2.2. The Cs2NaInI6-based PSC structure
The absorber crystal structure has shown in the Fig. 1(a). On the other hand, the proposed planar heterojunction PSC architecture, ITO/ETL (WS2, SnS2, In2S3, IGZO)/Cs2NaInI6/Au, illustrated in Fig. 1(b), Cs2NaInI6 is employed as the key light-absorbing layer in the device architecture. This material plays a central role in enhancing solar cell performance because of its optimal bandgap, which is well-suited for photovoltaic applications, and its strong optical absorption coefficient, enabling efficient harvesting of incident photons. The aluminum (Al) grid shown in the schematic represents the front contact metallization, which improves charge extraction by providing low-resistance pathways and thereby minimizes series resistance losses. Indium-doped tin oxide (ITO) is utilized as the transparent conducting oxide (TCO) at the front contact, allowing both high optical transparency and good electrical conductivity. Finally, gold (Au) functions as the rear metal electrode of the perovskite solar cell, with a work function of approximately 5.1 eV, ensuring efficient hole collection and stable back contact formation.20 With the intention of analyzing the effect of ETLs on device output, four materials including WS2, SnS2, In2S3, and indium gallium zinc oxide (IGZO) were systematically investigated. The material parameters and layer-specific properties used in the simulations are detailed in Tables 1 and 2. All computational analyses were accomplished using standard AM1.5 G illumination conditions with an incident power density of 1000 mW cm−2 and an ambient operating temperature of 300 K, confirming consistency with real-world solar cell test environments.
 |
| | Fig. 1 (a) Crystal structure of the Cs2NaInI6, and (b) device architecture of the Cs2NaInI6-based PSC. | |
Table 1 The parameters used for the proposed device structure
| Parameters |
ITO20 |
IGZO20,31 |
In2S3 (ref. 32) |
WS2 (ref. 20) |
SnS2 (ref. 33) |
Cs2NaInI6 (ref. 34) |
| Thickness (nm) |
30 |
30 |
50 |
100 |
100 |
1200 |
| Band gap, Eg (eV) |
3.5 |
3.05 |
2.82 |
1.80 |
2.24 |
1.6 |
| Electron affinity, χ (eV) |
4.0 |
4.16 |
4.50 |
3.95 |
4.24 |
4.379 |
| Dielectric permittivity (relative), εr |
9.0 |
10 |
13.50 |
13.60 |
10.0 |
3.5 |
| Effective density of states in conduction band, NC (cm−3) |
2.2 × 1018 |
5.0 × 1018 |
2.2 × 1017 |
1.0 × 1018 |
2.2 × 1018 |
3.16 × 1018 |
| Effective density of states in valence band, NV (cm−3) |
1.8 × 1019 |
5.0 × 1018 |
1.8 × 1019 |
2.4 × 1019 |
1.80 × 1019 |
1.71 × 1019 |
| Hole mobility, μh (cm2 V−1 s −1) |
10 |
10 |
25 |
100 |
50 |
50 |
| Electron mobility, μn (cm2 V−1 s −1) |
20 |
15 |
100 |
100 |
50 |
50 |
| Electron thermal velocity (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
| Hole thermal velocity (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
| Shallow uniform donor density, ND (cm−3) |
1 × 1018 |
1 × 1019 |
1 × 1016 |
1 × 1019 |
1 × 1018 |
0 |
| Shallow uniform acceptor density, NA (cm−3) |
0 |
0 |
0 |
0 |
0 |
1 × 1018 |
| Defect density, Nt (cm−3) |
1 × 1016 |
1 × 1014 |
1 × 1015 |
1 × 1016 |
1 × 1015 |
1 × 1014 |
Table 2 The interface parameters applied in the Cs2NaInI6-based PSCs
| Parameters |
Cs2NaInI6/IGZO |
Cs2NaInI6/In2S3 |
Cs2NaInI6/WS2 |
Cs2NaInI6/SnS2 |
| Total defect density (cm−2) |
1013 |
1013 |
1013 |
1013 |
3. Results and discussion
3.1. Energy band diagram
Fig. 2(a)–(d) reveals the band diagrams of the four optimized Cs2NaInI6-based PSCs. Band alignment discontinuities arise at the interfaces of heterostructures formed by semiconductors with different electronic properties. These interfacial discontinuities are characterized by the conduction band offset (CBO) and valence band offset (VBO), which significantly influence charge carrier dynamics at the absorber/ETL interface.35,36 A small positive CBO, or ‘spike’ (∼0–0.3 eV), is generally favorable because it prevents hole back-injection while still enabling efficient electron extraction. In contrast, a large negative CBO, or ‘cliff,’ enhances interfacial recombination and degrades performance. In this study, WS2, SnS2, and IGZO form small positive spikes with Cs2NaInI6 due to their slightly lower electron affinities, whereas In2S3 forms a negative cliff because of its larger electron affinity relative to the absorber. These results are consistent with the values in Table 1 and the band alignments in Fig. 2, underscoring the importance of achieving a modest spike for optimal photovoltaic operation.
 |
| | Fig. 2 The energy band diagrams of Cs2NaInI6-based solar cell devices having (a) WS2, (b) SnS2, (c) In2S3, and (d) IGZO ETL. | |
As demonstrated in Fig. 2(a)–(d), the quasi-Fermi levels for electrons (Fn) and holes (FP) appropriately align near the conduction band minimum (EC) and valence band maximum (EV), respectively, indicating effective charge separation and transport under illumination. Across all ETLs, Fn and EC, while FP aligns with EV demonstrate a consistent and harmonious affiliation. The band gap of Cs2NaInI6 is identified as 1.60 eV, with varying band alignments observed across the different ETLs.
3.2. Impact of thickness changes of absorber and ETL layer
Fig. 3(a) presents the influence of varying the Cs2NaInI6 double perovskite active layer thickness, varied from 0.3 to 2.1 μm, in determining photovoltaic output of solar cells incorporating different ETL configurations each with a fixed thickness of 100 nm. In the ITO/ETL/Cs2NaInI6/Au architecture, increasing the absorber thickness leads to marked enhancements in key device parameters, primarily as a result of enhanced light absorption and increased efficiency in generating charge carriers. The enhanced photon absorption with greater thickness leads to higher JSC values, while VOC and FF improve due to reduced recombination and optimized charge extraction. However, beyond a certain thickness, the performance starts to stabilize, suggesting a suitable absorber thickness that achieves a balance between effective light absorption and optimal charge transport.37–39 For the WS2 ETL, the device demonstrates a notable improvement in performance metrics with increasing absorber thickness. The VOC reveals a gradual increase from 1.147 V to 1.204 V, whereas the JSC demonstrates a significant enhancement from 15.02 mA cm−2 to 22.95 mA cm−2. Simultaneously, the FF enhances moderately from 87.55% to 89.12%, culminating in a substantial enhancement in the PCE from 15.08% to 24.62%. An analogous behavior is seen for the SnS2 ETL, where VOC increases from 1.15 V to 1.21 V, and JSC exhibits a considerable rise from 14.36 mA cm−2 to 22.88 mA cm−2. The FF improves slightly yet consistently, ranging from 87.70% to 89.49%, while the PCE shows a remarkable gain from 14.45% to 24.77%. In the case of the In2S3 ETL, the photovoltaic parameters also display favorable progression. The VOC increases modestly from 1.15 V to 1.21 V, whereas the JSC improves significantly from 13.73 mA cm−2 to 22.80 mA cm−2, FF rises from 86.82% to 89.06%, and the PCE enhances significantly from 13.67% to 24.57%. Lastly, for the IGZO ETL, the device shows consistent performance enhancement across all metrics. The VOC gradually rises from 1.145 V to 1.209 V, while the JSC rises appreciably from 13.12 mA cm−2 to 22.73 mA cm−2, FF improves from 86.67% to 89.36%, and the PCE increases impressively from 13.02% to 24.57%. These improvements are ascribed to enhanced photon capture and more efficient generation of charge carriers with increased thickness.40 The optimized absorber thickness for all ETL materials is found to be 1.20 μm, with the WS2 ETL showing the best comprehensive performance. This demonstrates the significance of thickness optimization in Cs2NaInI6-based solar cells, where balancing absorption and charge transport properties maximizes both efficiency and stability.
 |
| | Fig. 3 Changes in solar cell characteristics with the variation of (a) Cs2NaInI6 double perovskite absorber, and (b) ETL thicknesses. | |
The Fig. 3(b) demonstrates the role of layer thickness of the ETL on the photovoltaic output of PSCs with absorber thickness of 1.20 μm. As the ETL thickness increases, the PCE rises, with WS2 showing the most notable improvement, reaching around 23.40%, while IGZO experiences the smallest gain, reaching about 22.50%. A similar pattern is seen in the FF, where WS2 achieves the highest value of 89.10%, while IGZO shows the least change. The JSC remains relatively constant for all materials, with slight increases as the ETL thickness grows, and WS2 consistently performs better than the others. The VOC also shows a modest increase across all materials, with WS2 maintaining the highest value near 1.195 V. In conclusion, WS2 stands out as the most efficient material in terms of PCE, FF, JSC, and VOC, making it the best performer for this ETL setup.
3.3. The consequences of variations in the ETL defect density
Fig. 4 demonstrates the role of changes in the ETL defect density affect the output of different PSC configurations. In the ITO/ETL (WS2, SnS2, In2S3 and IGZO)/Cs2NaInI6/Au structure (Fig. 4(a)), increasing the ETL defect density (Nt) from 1012 cm−3 to 1016 cm−3 causes no substantial changes in the VOC, which remains fixed at 1.19 V for WS2 and IGZO, and at 1.194 V for SnS2 and In2S3. This suggests that variations in the defect density do not notably affect the voltage behavior for these materials within this range. In Fig. 4(b), the JSC remains stable as far as the defect density of Nt = 1015 cm−3.
 |
| | Fig. 4 Role of changing ETL defect density on the performance parameter, (a) VOC, (b) JSC, (c) FF, and (d) PCE of PSC with different ETL. | |
However, exceeding this value, a minor reduction in JSC is noticed for all ETL materials, implying that higher defect densities reduce current density, likely due to increased recombination or scattering at the defect sites, which affects charge transport. Fig. 4(c) shows that the FF is constant for In2S3 and IGZO across all defect densities, indicating that these materials maintain efficient charge extraction despite the defect density variations.41 For WS2 and SnS2, the FF remains steady up to Nt = 1014 cm−3, but increases in defect density beyond this led to a slight reduction in FF—from 88.92% to 88.87% for WS2 and from 88.182% to 88.173% for SnS2. This reduction suggests that higher defect densities in these materials slightly impair charge collection efficiency, possibly due to increased recombination or the formation of trap states that hinder charge extraction. Finally, Fig. 4(d) shows that the PCE declines as the defect density of the ETL rises, a trend consistent with the drops in JSC and FF. This decrease in PCE highlights the negative role of increased defect density on the overall output of the PSC due to the combined effects of lower current density and less efficient charge collection.
Notably, the impact of ETL defect density on device performance is much more significant for IGZO compared to WS2, SnS2, and In2S3. This distinction arises from the amorphous nature of IGZO, which inherently exhibits a higher density of localized sub-gap states.27 Defects in IGZO act as efficient recombination centers, thereby intensifying non-radiative losses and limiting charge extraction.27 In contrast, crystalline ETLs such as WS2, SnS2, and In2S3 are structurally more ordered and display stronger defect tolerance, allowing their photovoltaic behavior to remain relatively stable under similar defect density variations. Consequently, IGZO-based devices show stronger performance degradation with increasing defect density, consistent with the trends observed in Fig. 4.
3.4. The result of changes in the ETL shallow donor density
Fig. 5(a–d) comprehensively demonstrates the role of varying shallow donor densities (ND) in the ETLs spanning from 1015 to 1021 cm−3 (ref. 42 and 43) on the photovoltaic operation of PSCs employing various ETL materials. As the shallow donor density increases, a progressive decline in VOC is consistently observed across all configurations. The observed decrease is largely caused by the enhanced recombination of charge carriers, as donor-like defect states in the ETL serve as non-radiative recombination centers, thereby reducing the quasi-Fermi level separation and lowering the achievable VOC.44 Specifically, for WS2, VOC declines markedly from 1.236 V to 1.194 V; for SnS2, a subtle lessening from 1.197 V to 1.194 V is detected; In2S3 exhibits a minor drop from 1.194 V down to 1.192 V; and for IGZO, VOC decreases from 1.199 V to 1.193 V. These reductions, while varying in magnitude, underscore the sensitivity of voltage output to defect-induced recombination processes.
 |
| | Fig. 5 The role of varying ETL shallow donor density on the output, (a) VOC, (b) JSC, (c) FF, and (d) PCE of PSCs with various ETL configurations. | |
In contrast, the JSC shows material-specific behavior in response to increased donor density. For WS2, a slight but consistent improvement is seen, with JSC rising from 21.303 mA cm−2 up to 21.414 mA cm−2. Similarly, SnS2 reveals a moderate increase from 21.044 mA cm−2 to 21.273 mA cm−2, and IGZO improves from 20.331 mA cm−2 to 20.939 mA cm−2, suggesting enhanced charge collection possibly due to increased carrier concentration. However, In2S3 deviates from this trend, exhibiting a marginal decline in JSC from 21.083 mA cm−2 to 20.830 mA cm−2, likely due to a rise in recombination losses outweighing the benefits of elevated carrier concentration. These variations reflect the influence of shallow donor defects on charge transport, with some materials benefiting from increased current density while others show a reduction due to higher defect-related recombination.45
The FF reveals substantial changes for various ETL material: for WS2, FF increases from 32.92% to 89.13%; for SnS2, FF rises from 71.79% to 89.09%; for In2S3, FF decreases from 88.71% to 87.59%; and for IGZO, FF improves from 34.67% to 88.96%. The increase in FF for most ETL materials indicates better charge extraction efficiency, particularly for WS2 and IGZO, which see substantial improvements. Regarding PCE, WS2 shows a remarkable increase from 8.87% to 22.79%; SnS2 increases from 18.08% to 22.62%; In2S3 experiences a slight decrease from 22.32% to 21.23%; and IGZO rises from 8.45% to 22.22%. The optimized values for shallow donor defect density are found to be 1019 cm−3 for WS2, 1018 cm−3 for SnS2, 1016 cm−3 for In2S3, and 1019 cm−3 for IGZO. The analysis reveals highlight the critical role of shallow donor defects in the ETL layer, demonstrating that their density significantly influences the optimizing charge transport, solar cell performance, minimizing recombination, and ultimately enhancing the PCE. Furthermore, it is observed in Fig. 5 that the variation of ETL shallow donor density primarily affects the FF, with the effect being most pronounced for WS2 and IGZO, while relatively insignificant for In2S3. This trend can be understood in terms of conductivity enhancement and recombination dynamics. For WS2 and IGZO, increased shallow donor density significantly alters carrier transport by improving electron conductivity, but at the same time it can intensify interfacial recombination processes, leading to notable FF fluctuations. In contrast, In2S3 exhibits more stable performance because its moderate electron affinity46 and band alignment with Cs2NaInI6 result in less sensitivity to donor density changes. Consequently, the FF of In2S3-based devices remains comparatively stable under varying shallow donor densities, explaining the weaker impact observed in Fig. 5.
Based on the comparative analysis in Sections 3.1–3.4, WS2 was identified as the most promising ETL due to its favorable conduction band alignment, low recombination losses, and superior device metrics relative to SnS2, In2S3, and IGZO. Therefore, the subsequent analyses of defect density, interface defect density, thermal response, J–V and QE characteristics (Sections 3.5 and 3.7) focus primarily on the WS2-based configuration, which best represents the optimized device design.
3.5. Role of simultaneous variations in defect and acceptor concentrations
The device output noticeably declines due to extensive carrier recombination effects, comprising non-radiative recombination, Shockley–Read–Hall, and Auger, produced by larger values of acceptor density (NA) and defect density (Nt).47–49 Elevated recombination of photogenerated carriers markedly reduces cell efficiency.50–52 In this study, we investigate how different NA and Nt values within specified boundaries of 1012–1016 cm−3 and 1015–1021 cm−3, respectively, affect output in Cs2NaInI6 absorber-based PSC. The aim is to evaluate device performance using practical and reliable data for realistic responses, as exhibited in Fig. 6.
 |
| | Fig. 6 Changes in defect and acceptor density of the Cs2NaInI6 active layer affect the following output metrics; (a) VOC (b) JSC (c) FF and (d) PCE. | |
Fig. 6(a) vividly demonstrates the impact of Nt and NA on the VOC of the ITO/WS2/Cs2NaInI6/Au PSC structure. In this analysis, NA is varied over an extensive span ranging from 1015 to 1021 cm−3, even though Nt spans from 1012 to 1016 cm−3. The VOC exhibits a pronounced enhancement, increasing from 0.96 V to a peak value of 1.41 V. An optimized VOC of 1.189 V is consistently accomplished once NA is maintained at 1018 cm−3 and Nt remains below 1015 cm−3. However, a noticeable degradation occurs when NA drops below 1017 cm−3, causing VOC to decline significantly to 0.96 V, underscoring the detrimental effect of inadequate doping. Fig. 6(b) presents the variation in JSC, which ranges widely from 5.650 mA cm−2 to 23.90 mA cm−2. The maximum JSC of 21.409 mA cm−2 is obtained under optimal doping conditions (NA = 1018 cm−3, Nt < 1015 cm−3), suggesting enhanced carrier generation and collection efficiency. According to Fig. 6(c), the FF presents a dramatic variation between 26.80% and 90.81%. The optimum FF of 88.92% is observed at NA = 1018 cm−3 and Nt < 1014 cm−3, reflecting a highly efficient charge transport mechanism with minimized recombination. Fig. 6(d) further highlights the corresponding variation in PCE, ranging from 6.850% to an impressive maximum of 22.70%. The highest PCE of 22.63% is achieved under the same optimized conditions, reinforcing the strong dependency of overall device performance on finely tuned doping and defect parameters.
3.6. Role of interface defect density and temperature on device performance
Photogenerated charge carrier populations are critically suppressed by interface trap states that facilitate non-radiative recombination and adversely impact their efficient extraction and transport.53 Variations in the defect density at the ETL/absorber interface specifically within the Cs2NaInI6 double perovskite layer ranging from 1011 to 1015 cm−2, significantly impact the overall photovoltaic output as demonstrated in Fig. 7(a). The interfacial characteristics between the double perovskite active layer and the WS2 ETL serves an vital and highly sensitive role in determining device performance.54–56 Once the interface defect density surpasses 1013 cm−2, a pronounced degradation in output is noticed. This performance worsening is primarily linked to the higher availability of trap-assisted recombination pathways, which accelerate charge carrier loss and undermine the built-in electric field, thereby limiting voltage output and carrier collection.57,58 As the interface defect density rises, VOC drops from 1.19 V down to 1.16 V, with an optimized value of 1.19 V attained at 1013 cm−2. Similarly, JSC declines noticeably from 21.47 mA cm−2 to 18.39 mA cm−2, with its maximum optimized value of 21.41 mA cm−2 also corresponding to the same defect level. Moreover, both FF and PCE follow a downward trajectory, where FF drops from 89.10% to 87.63%, and PCE is reduced from 22.84% to 18.44%. Once the interface defect density surpasses 1013 cm−2, a pronounced degradation in output is noticed. However, at an interface defect density of 1013 cm−2, both PCE and VOC remain close to their highest simulated values, confirming this point as the optimum condition before significant performance decline occurs. This performance worsening at higher defect levels is primarily linked to the increased availability of trap-assisted recombination pathways,59 which accelerate charge carrier loss and undermine the built-in electric field, thereby limiting voltage output and carrier collection. Thus, Fig. 7(a) clearly indicates the optimum interface defect density as 1013 cm−2.
 |
| | Fig. 7 Role of (a) interface defect density and (b) temperature variation on the output of PSCs. | |
As described in Fig. 7(b), this study systematically investigates simulated thermal response of the device as the operating temperature varied from 280 K to 370 K, on the output of the proposed PSC architectures. The simulation outcomes clearly reveal that with a progressive rise in temperature, the PCE gradually declines across all device configurations. This efficiency degradation is predominantly attributed to the thermally induced enhancement in reverse saturation current, which intensifies non-radiative carrier recombination and diminishes the overall charge extraction efficiency.35,60 Despite this universal trend, Cs2NaInI6-based PSCs demonstrate remarkable thermal response, particularly in configurations utilizing the WS2 ETL. Even under elevated thermal stress, the PCE exhibits only a modest decline, signifying the material's inherent stability and resilience against thermal fluctuations rendering it suitable for potential application in real-world surroundings where operational temperatures can vary substantially.
Furthermore, Fig. 7(b) indicates a distinct downward trend in VOC with increasing temperature, most notably in the WS2-based configuration. This reduction is fundamentally associated with the temperature-induced narrowing of the semiconductor bandgap, which lowers the quasi-Fermi level splitting and, consequently, the maximum attainable voltage output. In contrast, the JSC remains largely unaffected across the entire temperature spectrum. This relative invariance is ascribed to the stability of photogeneration mechanisms, as the incident photon flux and intrinsic absorption characteristics remain substantially constant with temperature, thereby sustaining steady carrier generation. The FF, however, exhibits a consistent decline as the temperature increases. This behavior can be correlated with elevated series resistance and diminished carrier mobility, which collectively impede charge transport and reduce the ideality of the current–voltage (J–V) behavior. The synergistic lessening in both VOC and FF largely contributes to the detected deterioration in PCE. Overall, these observations highlight the significant role of material and interface optimization to boost the thermal stability and maintain the high operational performance of Cs2NaInI6-based PSCs under diverse environmental conditions.
3.7. Analysis of quantum efficiency (QE) and current–voltage (J–V) properties
A comprehensive evaluation of the QE and J–V behavior is fundamentally critical for elucidating the optoelectronic output of devices. These metrics serve as powerful diagnostic tools, offering quantitative comprehensive grasp of the effectiveness of light absorption, carrier photogeneration, and the electrical response of the solar cell.40,61,62 The QE spectrum specifically reflects the proportion of collected charge carriers relative to the incoming photon flux, thereby directly correlating with spectral responsivity. Conversely, the J–V profile reveals the current output as a function of applied bias, enabling detailed assessment of key parameters.29,36,63
Fig. 8(a) and (b) illustrate the J–V and QE characteristics for different ETLs. The optimized Al/ITO/WS2/Cs2NaInI6/Au structure was achieved with finely tuned parameters: absorber thickness of 1.2 μm, bulk defect density of 1014 cm−3, acceptor density of 1018 cm−3, and interface defect density of 1013 cm−2. The simulated J–V curve in Fig. 8(a) shows a VOC of 1.189 V and a high JSC of 21.409 mA cm−2. However, as forward bias exceeds 1.20 V, JSC decreases sharply toward zero due to carrier saturation and recombination. Fig. 8(b) presents the QE response, where quantum efficiency rises steadily up to ∼400 nm in the UV-visible region, then gradually declines with increasing wavelength. A sharp drop occurs beyond 700 nm, reaching zero near 790 nm, corresponding to the absorber's bandgap limit. Together, these results highlight the wavelength-dependent absorption and electrical behavior of the optimized device, confirming WS2 as an effective ETL and demonstrating the strong efficiency potential of Cs2NaInI6-based PSCs under practical conditions.
 |
| | Fig. 8 (a) J–V, (b) QE curve with various ETL of Cs2NaInI6-based double PSC. | |
3.8. Solar cell performance
Table 3 provides an in-depth comparison of key output metrics for Cs2NaInI6-based PSCs employing various ETLs. Among the evaluated configurations, the device incorporating WS2 as the ETL exhibits the highest PCE of 22.63%, surpassing the others in overall performance. It also achieves a notably high JSC of 21.406 mA cm−2, a moderately strong FF of 88.92%, and a respectable VOC of 1.189 V. Following closely, the SnS2-based configuration records a PCE of 22.36%, a JSC of 21.234 mA cm−2, and an FF of 89.18%, indicating efficient charge transport and collection. The In2S3-based solar cell demonstrates a competitive performance with a PCE of 22.32%, a JSC of 21.083 mA cm−2, and an FF of 88.71%. Meanwhile, the IGZO-based device shows the lowest efficiency, with a PCE of 22.23%, a JSC of 20.943 mA cm−2, and an FF of 88.96%. Overall, the WS2 ETL configuration distinctly outperforms the others, offering the most favorable synergy of high efficiency, enhanced current output, and stable voltage, thereby positioning itself as the most promising ETL candidate for high-performance Cs2NaInI6-based PSCs.
Table 3 Comparative Performance Metrics of different double perovskite PSCs Configurationsa
| Structure |
VOC (V) |
JSC (mA cm−2) |
FF (%) |
PCE (%) |
Ref. |
| E = Experimental, T = Theoretical. |
| ITO/IGZO/Cs2NaInI6/Au |
1.1932 |
20.943 |
88.96 |
22.23 |
This work |
| ITO/In2S3/Cs2NaInI6/Au |
1.1936 |
21.083 |
88.71 |
22.32 |
This work |
| ITO/WS2/Cs2NaInI6/Au |
1.189 |
21.406 |
88.92 |
22.65 |
This work |
| ITO/SnS2/Cs2NaInI6/Au |
1.194 |
21.234 |
89.18 |
22.60 |
This work |
| ITO/CuNiO/Cs2AgBiBr6/C60/BCP/Ag |
1.01 |
3.19 |
69.2 |
2.23 |
E64 |
| FTO/SnO2/Cs2AgBiBr6/Cu2O/Au |
1.511 |
15.76 |
69.84 |
16.63 |
T65 |
| FTO/TiO2/Cs2AgBiBr6/Cu2O/Au |
1.508 |
15.71 |
71.1 |
16.85 |
T65 |
| FTO/ZnO/Cs2AgBiBr6/Cu2O/Au |
1.510 |
15.75 |
72.52 |
16.79 |
T65 |
4. Conclusions
This study presents a detailed simulation-based investigation into the performance optimization of Cs2NaInI6 double perovskite solar cells (PSCs) by carefully selecting and tuning electron transport layers (ETLs). Using SCAPS-1D, four ETL candidates-WS2, SnS2, In2S3, and IGZO-were evaluated within a planar ITO/ETL/Cs2NaInI6/Au architecture. Device performance was systematically analyzed by varying absorber thickness, defect density, doping concentration, and operating temperature to assess their impact on power conversion efficiency (PCE), open-circuit voltage (VOC), short-circuit current density (JSC), and fill factor (FF). Among the studied ETLs, WS2 delivered the most promising results, achieving a maximum PCE of 22.63%, supported by favorable VOC, JSC, and FF values. This superior performance is attributed to its advantageous band alignment, high carrier mobility, and minimized interfacial recombination. The optimized structure, with a 1.2 μm absorber and interface defect density of 1013 cm−2, also demonstrated strong electronic and thermal stability, underscoring the potential of Cs2NaInI6 for sustainable photovoltaic applications. Overall, the work highlights the critical influence of ETL engineering on lead-free PSC efficiency and provides simulation-based guidelines for designing environmentally friendly, cost-effective solar devices. However, these stability claims are theoretical, and experimental validation remains essential to confirm long-term device reliability.
Conflicts of interest
There are no conflicts of interest to declare.
Data availability
Data will be available on request.
Acknowledgements
The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through Large Research Project under grant number RGP2/250/46.
References
- A. Kuddus, M. K. A. Mohammed, A. K. Al-mousoi, A. Ghosh, S. Bhattarai, R. Pandey, J. Madan and M. K. Hossain, RSC Adv., 2023, 31330–31345 Search PubMed.
- M. K. Hossain, M. S. Uddin, G. F. I. Toki, M. K. A. Mohammed, R. Pandey, J. Madan, F. Rahman, R. Islam, S. Bhattarai, H. Bencherif, D. P. Samajdar, M. Amami and D. K. Dwivedi, RSC Adv., 2023, 13, 23514–23537 CAS.
- M. M. A. Moon, M. F. Rahman, J. Hossain and A. B. M. Ismail, Adv. Mater. Res., 2019, 1154, 102–111 Search PubMed.
- N. I. Ahmad, Y. B. Kar, C. Doroody, T. S. Kiong, K. S. Rahman, M. N. Harif and N. Amin, Heliyon, 2023, 9, e21622 CAS.
- S. Khalid, M. Sultan, E. Ahmed and W. Ahmed, in Emerging Nanotechnologies for Renewable Energy, Elsevier, 2021, pp. 3–35 Search PubMed.
- M. K. Hossain, M. S. Uddin, G. F. I. Toki, M. K. A. Mohammed, R. Pandey, J. Madan, M. F. Rahman, M. R. Islam, S. Bhattarai, H. Bencherif, D. P. Samajdar, M. Amami and D. K. Dwivedi, RSC Adv., 2023, 13, 23514–23537 RSC.
- G. Mohamed, A. Amina, H. Ali, H. Bendjilali and O. ZERROUKI, The structural, electronic, elastic, and optical properties of new double perovskite Cs 2 CdPbI 6 were investigated using a DFT and SCAPS-1D simulation, J. Inorg. Organomet. Polym. Mater., 2025, 35, 4616–4630 CrossRef.
- H. El-assib, M. Alla, S. Tourougui, M. Alla, F. Elfatouaki, S. Dar, A. Chauhan, Naima, N. Chawki, N. Shrivastav, V. Manjunath, M. Rouchdi and B. Fares, Renewable Energy, 2024, 239, 122092 CrossRef.
- A. Alfares, Y. A. Sha’aban and A. Alhumoud, Eng. Appl. Artif. Intell., 2025, 141, 109747 CrossRef.
- L. Rani Karna, R. Upadhyay and A. Ghosh, Sol. Energy, 2024, 281, 112843 CrossRef CAS.
- D. Bharath Raja, R. Vidya and K. Shanmuga Sundaram, Sol. Energy, 2022, 245, 353–364 CrossRef CAS.
- T. Chonamada, A. Dey and P. Santra, ACS Appl. Energy Mater., 2020, 3(1), 47–55 CrossRef CAS.
- Y. Li, T. Yang, Y. She, B. Xu, Y. Du and M. Zhang, Inorg. Chem., 2023, 62, 19248–19255 CrossRef CAS PubMed.
- S. Umer, M. I. Khan, A. Ullah, Ihtisham-ul-haq, M. Asad, I. kebaili, W. Mnif, Z. Algarni and M. I. Saleem, Opt. Mater., 2024, 156, 115896 CrossRef CAS.
- A. Amjad, S. Qamar, C. Zhao, K. Fatima, M. Sultan and Z. Akhter, RSC Adv., 2023, 13, 23211–23222 RSC.
- H. H. AbdElAziz, M. Taha, W. M. A. El Rouby, M. H. Khedr and L. Saad, Heliyon, 2022, 8, e10808 CrossRef CAS PubMed.
- G. M. Mustafa, M. U. Nawaz, N. A. Noor, S. Saba, A. Mahmood and M. Iqbal, Phys. Scr., 2024, 99, 45960 CrossRef CAS.
- M. Waqas Mukhtar, M. Ramzan, M. Rashid, G. Naz, M. Imran, F. Fahim, A. A. AlObaid, T. I. Al-Muhimeed and Q. Mahmood, Mater. Sci. Eng., B, 2021, 273, 115420 CrossRef CAS.
- H. Jebari, I. Hamideddine and H. Ez-Zahraouy, J. Korean Ceram. Soc., 2024, 61, 599–604 CrossRef CAS.
- M. K. Hossain, G. F. I. Toki, A. Kuddus, M. H. K. Rubel, M. M. Hossain, H. Bencherif, M. F. Rahman, M. R. Islam and M. Mushtaq, Sci. Rep., 2023, 13, 1–24 CrossRef PubMed.
- A. S. Shaikh, S. C. Yadav, A. Srivastava, A. R. Kanwade, M. K. Tiwari, S. M. Rajore, J. A. K. Satrughna, M. Dhonde and P. M. Shirage, Energy Adv., 2024, 3, 1505–1552 RSC.
- H. Pan, X. Zhao, X. Gong, H. Li, N. H. Ladi, X. L. Zhang, W. Huang, S. Ahmad, L. Ding, Y. Shen, M. Wang and Y. Fu, Mater. Horiz., 2020, 7, 2276–2291 RSC.
- M.-H. (Michelle)Wu, J. S. Konthoujam, I. Lin, T.-Y. Peng, Y.-J. Lu and M.-H. Shih, Nanoscale Adv., 2025, 7, 5944–5950 RSC.
- I. Paul, A. Rayhan and M. A. Khan, New Energy Exploit. Appl., 2025, 4, 83–101 CrossRef.
- M. T. Islam and S. K. Mukaddar, Phys. Scr., 2025, 100, 025911 CrossRef CAS.
- M. R. Kabir, M. H. Ishraq, M. Tarekuzzaman, N. I. Nahid, J. K. Modak, S. Ahmad, A. M. Arfi, M. Rasheduzzaman and M. Z. Hasan, Eng. Res. Express, 2025, 7, 015319 CrossRef.
- K. Nomura, A. Takagi, T. Kamiya, H. Ohta, M. Hirano and H. Hosono, JJAP, 2006, 45, 4303–4308 CrossRef CAS.
- S. M. Seyed-Talebi, M. Mahmoudi and C. H. Lee, Micromachines, 2023, 14, 1562 CrossRef PubMed.
- A. Ghosh, M. F. Islam Buian, M. Maniruzzaman, M. M. Hossain, A. K. Azad, A. A. Miazee, I. Ragab, A. A. Hassan, H. A. Alrafai and S. K. Alla Abdelrahim, RSC Adv., 2024, 14, 26437–26456 RSC.
- Y. He, L. Xu, C. Yang, X. Guo and S. Li, Nanomaterials, 2021, 11, 1–19 Search PubMed.
- D. Dastan, M. K. A. Mohammed, A. K. Al-Mousoi, A. Kumar, S. Q. Salih, P. S. JosephNg, D. S. Ahmed, R. Pandey, Z. M. Yaseen and M. K. Hossain, Sci. Rep., 2023, 13, 1–11 CrossRef PubMed.
- A. U. Rehman, S. Afzal, I. Naeem, T. Munir, S. G. Sarwar, M. Saleem and R. M. Obodo, Proc. Pak. Acad. Sci.: A, 2025, 62, 81–95 Search PubMed.
- M. F. Rahman, M. Rahman, M. F. Hossain, B. Islam, S. R. Al Ahmed and A. Irfan, Adv. Photonics Res., 2025, 6, 1–14 CrossRef.
- H. Jebari, H. Hamideddine and I. Ez-Zahraouy, J. Korean Ceram. Soc., 2024, 61, 599–604 CrossRef CAS.
- S. Malik and A. Dutta, J. Phys. Chem. Solids, 2024, 190, 112027 CrossRef CAS.
- A. Ghosh, M. Ferdous Rahman, A. Kuddus, M. K. A. Mohammed, M. Rasidul Islam, S. Bhattarai, A. R. Chaudhry and A. Irfan, J. Alloys Compd., 2024, 174097 CAS.
- M. Boujelbene, M. Rochi Khan, M. Azizur Rahman, A. Ghosh, M. Sadia Islam Ria, M. Asikul Islam, J. Y. Al-Humaidi, N. Elboughdiri, A. Albishri, A. Rasool Chaudhry and M. M. Rahman, Inorg. Chem. Commun., 2024, 170, 113472 CrossRef CAS.
- M. A. Rahman, M. F. Rahman, M. H. Rahman, M. R. Islam and M. Z. Bani-Fwaz, Appl. Phys. Lett., 2024, 125, 193902 CrossRef CAS.
- M. F. Rahman, T. Al Galib, M. A. Rahman, M. H. Rahman, M. Harun-Or-Rashid, M. A. I. Islam, M. M. Islam, N. Dhahri and A. Irfan, Phys. Chem. Chem. Phys., 2025, 27, 1155–1170 RSC.
- M. Sadia Islam Ria, A. Ghosh, M. Azizur Rahman, J. Y. Al-Humaidi, N. S. Awwad, H. A. Ibrahium and M. M. Rahman, Mater. Sci. Eng., B, 2024, 310, 117674 CrossRef CAS.
- M. Lamba, K. Maurya and V. N. Singh, Heliyon, 2022, 8, e10925 CrossRef PubMed.
- M. N. Nahid, M. S. Shah, H. Mamur, R. Hosen and M. R. A. Bhuiyan, Chem. Inorg. Mater., 2024, 4, 100069 Search PubMed.
- M. D. Wanda, S. Ouédraogo, F. Tchoffo, F. Zougmoré and J. M. B. Ndjaka, Int. J. Photoenergy, 2016, 2016, 2152018 Search PubMed.
- M. Dhankhar, O. Pal Singh and V. N. Singh, Renewable Sustainable Energy Rev., 2014, 40, 214–223 CrossRef.
- C. Ewels, R. Jones, S. Öberg, J. Miro and P. Deak, Phys. Rev. Lett., 1996, 77, 865–868 CrossRef CAS PubMed.
- T. A. Chowdhury, Opt. Continuum, 2024, 3, 2377–2399 Search PubMed.
- T. Liang, C. Hu, M. Lou, Z. Feng, D. Wang, X. Cai and L. Lin, Langmuir, 2023, 39, 10769–10778 CrossRef CAS PubMed.
- X. Cai, G. Chen, R. Li and Y. Jia, Langmuir, 2023, 39(41), 14791–14799 CrossRef CAS PubMed.
- S.-N. Guo, H. Wu, D. Wang and J.-X. Wang, Langmuir, 2021, 37, 11520–11525 CrossRef CAS PubMed.
- G. K. Gupta, A. Garg and A. Dixit, J. Appl. Phys., 2018, 123, 013101 CrossRef.
- F. Ferdiansjah and K. Tirtakusuma Mularso, E3S Web Conf., 2018, 43, 01006 CrossRef.
- M. Samiul Islam, K. Sobayel, A. Al-Kahtani, M. A. Islam, G. Muhammad, N. Amin, M. Shahiduzzaman and M. Akhtaruzzaman, Nanomaterials, 2021, 11, 1218 Search PubMed.
- M. Salem, M. Salah, M. Moussa, A. Shaker, A. Zekry, M. Abouelatta-Ebrahim, M. Alshammari, K. Aldhlan and C. Gontrand, IEEE Access, 2021, PP1 Search PubMed.
- L. Xiang, Y. Cao, J. Sun, D. Li, H. Liu, F. Gao, C. Chen and S. Li, Surf. Interfaces, 2024, 44, 103641 CrossRef CAS.
- A. S. Chouhan, N. P. Jasti and S. Avasthi, Mater. Lett., 2018, 221, 150–153 CrossRef CAS.
- F. Rahman, M. Chowdhury and L. Marasamy, RSC Adv., 2024, 1924–1938 RSC.
- N. Rahman, M. D. Haque, M. F. Rahman, M. M. Islam, M. A. N. Juthi, A. R. Roy, M. A. Akter and M. F. Islam, Discover Mater., 2023, 3, 25 CrossRef CAS.
- T. Ouslimane, L. Et-taya, L. Elmaimouni and A. Benami, Heliyon, 2021, 7, e06379 CrossRef CAS PubMed.
- M. Hasnain, A. Wahed, J. Das, B. Arkook, M. Harb and N. Mia, JPhys Energy, 2025, 7, 025004 Search PubMed.
- Q. Meng, Y. Chen, Y. Xiao, J. Sun, X. Zhang, C. Han, H. Gao, Y. Zhang and H. Yan, J. Mater. Sci.: Mater. Electron., 2021, 32, 12784–12792 CrossRef CAS.
- M. F. Rahman, M. A. Monnaf, M. Amami, L. Ben Farhat and M. A. Rahman, J. Phys. Chem. Solids, 2024, 112216 Search PubMed.
- M. A. Rahman, M. F. Rahman, L. Marasamy, M. Harun-Or-Rashid, A. Ghosh, A. R. Chaudhry and A. Irfan, Energy Fuels, 2024, 38, 8199–8217 Search PubMed.
- M. S. Reza, A. Ghosh, S. N. Wornob, M. S. Reza, A. K. Azad, M. M. Hossain, N. S. Awwad and H. A. Ibrahium, J. Phys. Chem. Solids, 2024, 194, 112250 Search PubMed.
- W. Gao, C. Ran, J. Xi, B. Jiao, W. Zhang, M. Wu, X. Hou and Z. Wu, ChemPhysChem, 2018, 19, 1696–1700 CrossRef CAS PubMed.
- S. Chand Yadav, J. A. K. Satrughna and P. M. Shirage, J. Phys. Chem. Solids, 2023, 181, 111515 Search PubMed.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.