Abstract
Perovskite solar cells (PSCs) are rapidly advancing due to their high power conversion efficiencies (PCEs) and low fabrication costs. However, their commercialization is hindered by lead toxicity and the use of expensive materials, such as Spiro-OMeTAD and gold electrodes. This study presents a comprehensive SCAPS-1D simulation-based analysis of 14 perovskite absorber materials, spanning both Pb-based and lead-free compounds, under a unified device architecture using low-cost, nontoxic components: ZnO as the electron transport material (ETM), PEDOT:PSS + WO3 as a dual hole transport material, and nickel as the back contact. Key device parameters, including absorber thickness, defect density, electron affinity, series resistance, and temperature, were systematically optimized. Among the materials evaluated, MASnI3 exhibited the highest PCE of 27.98%, surpassing even MAPbI3 (27.10%), and demonstrating its viability as a sustainable, nontoxic absorber. The proposed cell architecture achieves high efficiency at a total material cost below 4 euro/g, significantly reducing the economic barrier compared to conventional designs. This work establishes a scalable, eco-friendly PSC architecture and provides a simulation-driven framework to accelerate experimental development and industrial application.


1. Introduction
Photovoltaics has long played a pivotal role in global energy harvesting, with solar energy emerging as a reliable, abundant, and renewable solution to meet growing energy demand. Among emerging photovoltaic technologies, perovskite solar cells (PSCs) have garnered significant attention due to their exceptional power conversion efficiencies (PCEs), low-temperature fabrication, and tunable optoelectronic properties. Since the first report of a perovskite-based device with a 3.81% efficiency by Kojima et al., continuous research and material innovation have propelled PSCs to a certified efficiency of 25.8%,, making them one of the fastest-advancing solar technologies to date.
The fundamental way of representation for the crystal structure of perovskite is AMX3, where A denotes organic cation, whereas metal cations like Co2+, Cu2+, Cr2+, Fe2+, Ge2+, Cd2+, Mn2+, Sn2+, or Pb2+ are represented through M and X is used to illustrate halide anions like Br–, Cl–, or I–. Nowadays, HC(NH2)2 (FA), Cs+, or CH3NH3 + (MA) is being used as organic cation A, and Sn2+or Pb2+ is getting more preference as metal cation M to the researchers. Perovskite has several advantages over other absorber materials; for example, a minimal amount of excitation is required to dislodge the ions, which means that it has low excitation binding energy, and the absorption coefficient is relatively higher. Moreover, tunability is another crucial feature of perovskite materials. Additionally, the carrier (electron and hole) mobility is sublime, and the diffusion length is longer compared to other organic materials. Perovskite has very high absorption in the visible-light wavelength range, and no ellipsometric effect is available in this region. These properties have established perovskites as highly attractive absorber materials for thin-film solar applications.
Despite these promising attributes, two critical bottlenecks hinder the commercialization of PSCs: (i) the reliance on toxic lead-based compounds and (ii) the use of costly materials such as Spiro-OMeTAD and gold for charge transport and electrode layers. Studies have shown that these components can account for more than one-third (33.9%) of total device cost, making the search for low-cost, sustainable alternatives a pressing need. While some prior research has investigated nontoxic or inexpensive components individually, few have addressed all three dimensions such as efficiency, cost, and environmental safety, in an integrated optimization framework.
Recent advancements in both Pb-based and Pb-free perovskite solar cells have demonstrated significant progress in optimizing the efficiency, stability, and device architecture. For example, MAPbI3-based devices with polymer dual passivation have reached efficiencies above 20%, while thermally stable FAPbI3 and interface-engineered CsPbI3 cells have achieved up to 26%. , On the Pb-free front, MASnI3 and FASnI3 devices have been optimized using novel HTL/ETL combinations and plasmonic structures, reporting PCEs close to 27%. − These findings provide strong evidence of performance enhancement across absorber types, highlighting the need for a unified, simulation-based comparisonan area that this study directly addresses.
In this work, we introduce a comprehensive simulation-based methodology to design perovskite solar cells (PSCs) that simultaneously optimizes power conversion efficiency (PCE), environmental compatibility, and economic feasibility. Using the SCAPS-1D simulation platform, we systematically evaluate 14 perovskite absorber materials, comprising both Pb-based and Pb-free classes, under a fixed, low-cost, and scalable device architecture. This architecture features a novel material combination: ZnO as a stable and inexpensive electron transport material (ETM), PEDOT:PSS + WO3 as a dual-function hole transport layer (HTL), and nickel (Ni) as a cost-effective and earth-abundant back contact. Unlike conventional simulation studies that vary transport layers across absorber types, our work ensures direct and fair comparison by fixing all nonabsorber layers and maintaining identical simulation parameters, thus isolating absorber-specific effects.
While the environmental concerns surrounding Pb-based perovskites are well-established, we intentionally include MAPbI3, MAPbI3(1–x)Br3x , and MAPbI3–x Cl x in our analysis to serve as high-performance reference benchmarks. These lead-containing absorbers represent the current state of the art in PSC efficiency and allow us to critically assess whether Pb-free candidates, such as MASnI3, CsSnI3, and FASnI3, can match or exceed their performance under consistent architectural and operational constraints. The comparative scope of 14 absorbers simulated under uniform interfacial conditions is a key novelty of this work and provides a level of material-to-material comparability that is often missing in the literature.
Beyond the absorber comparison, this study introduces several innovations that distinguish it from previous simulation efforts. First, we demonstrate, for the first time in the context of MASnI3 and MAPbI3 devices, the synergistic effects of combining PEDOT:PSS + WO3 as a cost-effective dual HTL, which improves hole extraction and band alignment while maintaining low toxicity and high scalability. Second, we integrate material pricing from commercial suppliers (e.g., Sigma-Aldrich) directly into our design process to guide the selection toward low-cost architectures. The total material cost of our optimized cells remains below 4 euro/g, in stark contrast to conventional structures that often exceed 500 euro/g due to the use of gold and Spiro-OMeTAD.
Finally, we perform a multidimensional optimization across key physical parametersabsorber thickness, defect density, electron affinity, series resistance, and temperaturewhile analyzing conduction/valence band offsets (CBO/VBO), generation–recombination profiles, quantum efficiency spectra, and the impact of back contact metal work function. The MASnI3-based architecture achieves a record-level simulated efficiency of 27.98%, outperforming the Pb-based MAPbI3 (27.10%) under identical conditions. Notably, these high efficiencies are achieved using only earth-abundant and non-noble materials, making the design viable for industrial-scale production.
In summary, this study establishes a new simulation benchmark for eco-friendly and economically viable PSCs. It provides not only a novel device structure but also a robust comparative platform for evaluating the true potential of lead-free perovskites. The findings offer a practical roadmap toward the commercialization of PSCs that are highly performing, scalable, and sustainable.
2. Material Selection
Determining the highly efficient cell structure is a significant focus of this study. The efficiency of the cell structure is mainly dependent on the absorber materials. Although a bunch of perovskite absorber materials are available, all of them cannot exhibit a higher efficiency. Hence, we have chosen highly efficient absorber materials from previous studies. Different research groups have introduced several new perovskite materials and cell structures. Each cell structure shows different power conversion efficiencies (η (%)). The following is a list of common absorbers; their relative efficiency and references are given in Table .
1. PCE of Some Perovskite Materials and the Relevant References.
| absorber material | η (%) | reference |
|---|---|---|
| MASnI3 | 23.36% | |
| FASnI3 | 14.03% | |
| MAPbI3(1–x)Br3 x | 17.9% | |
| Cs2AgBiBr6 | 5.44% | |
| Cs2TiX6 | 6.75% | |
| MAGeI3 | 13.5% | |
| CsPbI3 | 21.06% | |
| MAPbI3 | ∼24% | , |
Table indicates that each perovskite absorber material has different η values (%) from others. Additionally, a perovskite absorber material can exhibit different η (%) under different conditions with different HTM and ETM. It is also clear from the table that there are mainly two types of absorber materials available. They are Pb-based absorber materials and other cation-based absorber materials. However, Pb induces toxicity that harms the human body and the environment, obstructing its commercialization. −
Determining highly efficient cell structures with inexpensive material is another key aspect of the study. The importance of a cost-effective cell structure can be realized by the cell structure and their relative cost represented in Table .
2. Common Cell Structure, Efficiency, and Cost of HTM, ETM, and Back Contact Materials .
| cell structure | ETM cost (euro/g) | HTM cost (euro/g) | back contact cost (euro/g) | total cost (euro) | η (%) |
|---|---|---|---|---|---|
| FTO/TiO2/CsGeI3/Spiro-OMeTAD/Ag | 27.2 | 528 | 9.01 | 564.21 | 18.30% |
| ZnO(nr)/CH3NH3SnI3/Cu2O/Au | 13.6 | 15.76 | 264 | 293.36 | 20.23% |
| ZnO(nr)/CH3NH3SnI3/Spiro-OMeTAD/Au | 13.6 | 528 | 264 | 805.6 | 20.21% |
| ITO/NiO/MASnI3/PCBM/Ag | 19.7 | 1130 | 9.01 | 1158.71 | 29.19% |
| FTO/Perovskite/Spiro-OMeTAD/Au | 0 | 528 | 264 | 792 | 19.16% |
| ITO/PEDOT:PSS/MAGeI3/PCBM/Ag | 1.324 | 1130 | 9.01 | 1140.334 | 11.16% |
| ITO/PEDOT:PSS/MAGeI3/C60/Ag | 1.324 | 677 | 9.01 | 687.334 | 13.5% |
| FTO/TiO2/MAPbI3/Spiro-OMeTAD/Au | 27.2 | 528 | 264 | 819.2 | 23.18% |
| FTO/TiO2/MAPbI3/NPB/Au | 27.2 | 725 | 264 | 1016.2 | 22.56% |
| FTO/TiO2/MAPbI3/P3HT/Au | 27.2 | 749 | 264 | 1040.2 | 19.38% |
| FTO/TiO2/MAPbI3/ MEH-PPV/Au | 27.2 | 798 | 264 | 1089.2 | 21.23% |
| ITO/NiO/CH3NH3PbI3–x Cl x /PCBM/Ag | 19.7 | 1130 | 9.01 | 1158.71 | 22% |
| FTO/TiO2/CH3NH3PbI3–x Cl x /Spiro-OMeTAD/Ag | 27.2 | 528 | 9.01 | 564.21 | 29.56% |
| FTO/ZnO/MASnI3/PEDOT:PSS/Au | 0.172 | 1.324 | 264 | 265.496 | 21.22% |
The price of each material is taken from Sigma-Aldrich.
Table indicates that the cells are either expensive or not optimized. The cells are costly, for either the high cost of HTM and ETM or the high cost of back contact materials. Hence, it is necessary to choose low-cost HTM, ETM, and back contact materials initially to simulate cost-effective and efficient perovskite solar cells. As simulation can reveal the output characteristics of the cell and is effective for designing and fabricating perovskite solar cells, it is necessary to choose low-cost HTM, ETM, and back contact to develop low-cost and highly efficient perovskite solar cells.
First, we collected common HTM and ETM for highly efficient cell structures from previous studies and their relative price to determine low-cost HTM and ETM. The following is a list of some common HTM and ETM and their relative cost. We have calculated the per gram price to make the comparison easier.
Table indicates that PEDOT:PSS and WO3 are the least expensive HTM, and ZnO is the least expensive ETM. Hence, we have chosen PEDOT:PSS + WO3 as the HTM and ZnO as the ETM for the study. Second, we have taken 14 common perovskite absorber materials for our research. We have divided those materials according to toxicity. Three of them are Pb-based absorber materials, while 11 of them do not contain Pb. We have taken more absorber materials as different absorber materials exhibit other output characteristics for different HTM and ETM. Hence, all common absorber materials are considered to determine the highly efficient cell structure. The cost of back contact material is also crucial for determining low-cost cells. However, we did not choose the back contact materials initially. Instead, we have simulated the best two cells from two categories with all possible back contact materials to reduce the risk of efficiency drop due to the impact of back contact. We have chosen the least expensive back contact materials from best performing back contact materials. In this process, the cell can retain efficiency with minimal expense.
3. Common HTM and ETM for Highly Efficient Perovskite Solar Cells and Their Cost .
| material name | material type | price (euro/g) |
|---|---|---|
| Spiro-OMETAD | HTM | 528 |
| CuI | HTM | 21.6 |
| NiO | HTM | 19.7 |
| PCBM | HTM | 1130 |
| PEDOT:PSS | HTM | 1.324 |
| WO3 | HTM | 1.26 |
| graphene | HTM | 1070 |
| P3HT | HTM | 749 |
| CdS | ETM | 10.35 |
| ZnO | ETM | 0.172 |
| SnO2 | ETM | 38.4 |
| WS2 | ETM | 23.8 |
The price of each material is taken from Sigma-Aldrich.
3. Device Physics and Simulation
Perovskite solar cells generally consist of HTM, ETM, an absorber, and contacts. However, several researchers have also introduced HTM-free and ETM-free perovskite solar cells. In this study, we chose a common cell structure with an ETM, absorber, HTM, and contacts. The light is illuminated from the ETM side of the cell. The following is the cell structure used in the study.
Simulations are performed by the SCAPS-1D simulator, a single-dimensional solar cell capacitance simulator developed by Burgelman and his co-workers at the University of Gent, Belgium. This prime solar cell simulator has several advantages, including the possibility of simulating the environmental and cell parameters and representing data graphically and numerically. Furthermore, the cell can be simulated by changing each parameter; the data can be saved and used later. Due to all of these advantages and user-friendly interfaces, we have chosen SCAPS-1D to model and simulate the cell structure in this study.
We have chosen PEDOT:PSS + WO3 as the HTM, and ZnO is used as the ETM for the current study. The general perovskite solar cell structure is shown as Figure . These materials are considered due to outstanding characteristics and the cost of these materials. Table lists the fundamental specifications of these materials.
1.

Device structure for a perovskite solar cell.
4. Basic Parameters of PEDOT:PSS + WO3 and ZnO.
| material | thickness (μm) | E g (eV) | χ (eV) | ε r | N c (cm –3 ) | N v (cm –3 ) | μ n (cm2/(V s)) | μ p (cm2/(V s)) | N D (cm –3 ) | N A (cm –3 ) | N t (cm –3 ) |
|---|---|---|---|---|---|---|---|---|---|---|---|
| PEDOT:PSS + WO3 | 0.1 | 1.8 | 3.4 | 18 | 2.2 × 1018 | 1.8 × 1019 | 4.5 × 10–2 | 4.5 × 10–2 | 0 | 1 × 1018 | 1 × 1014 |
| ZnO | 0.05 | 3.3 | 4.1 | 9 | 4 × 1018 | 1 × 1019 | 1 × 102 | 2.5 × 101 | 1 × 1018 | 1 × 105 | 1 × 1014 |
In Table , E g, χ, εr, N c, N v, μn, μp, N D, N A, and N t represent bandgap energy, electron affinity, relative permittivity, effective conduction band density, effective valence band density, electron mobility, hole mobility, donor concentration, acceptor concentration, and defect density, respectively.
During simulation, the interface layer between the HTM/perovskite absorber material and perovskite absorber material/ETM is also considered to significantly impact the cell’s performance. The interface layer can introduce defects in the interface between the HTM/absorber material and absorber material/ETM, reducing the cell’s performance. The temperature is 300 K, while the solar radiation spectrum is AM 1.5G. AM 1.5G provides an illumination power of 1000 W m–2. The hole and electron thermal velocities are 1 × 107 cm/s, while band-to-band recombination is 2.3 × 10–9 cm3/s.
This study uses two types of perovskite materials to design and simulate low-cost, highly efficient perovskite solar cells. First, Pb-based toxic materials such as MAPbI3, MAPbI3–x Cl x , and MAPbI3(1–x)Br3 x are analyzed. The simulation parameters for these Pb-based toxic materials are listed in Table .
5. Simulation Parameters of the Pb-Based Perovskite Absorber Materials.
| material | thickness (μm) | E g (eV) | χ (eV) | εr | N c (cm–3) | N v (cm–3) | μn (cm2/(V s)) | μp (cm2/(V s)) | N D (cm–3) | N A (cm–3) | N t (cm–3) |
|---|---|---|---|---|---|---|---|---|---|---|---|
| MAPbI3 | 0.5 | 1.5 | 3.9 | 10 | 2.75 × 1018 | 3.9 × 1018 | 1 × 101 | 1 × 101 | 0 | 1 × 1015 | 1 × 1014 |
| MAPbI3–x Cl x | 0.5 | 1.55 | 3.93 | 6.5 | 2.5 × 1020 | 2.5 × 1020 | 2 × 100 | 2 × 100 | 0 | 1 × 1015 | 1 × 1014 |
| MAPbI3(1–x)Br3 x | 0.5 | 1.7 | 3.79 | 6.5 | 2.2 × 1018 | 1.8 × 1019 | 2 × 100 | 2 × 100 | 0 | 1 × 1015 | 1 × 1014 |
In contrast, the other category of perovskite absorber materials consists of nontoxic perovskite materials, such as MASnI3, CsSnI3, FASnI3, MAGeI3, Cs3Bi2I9, Cs2AgBiBr6, Cs2TiI6, Cs2TiBr6, Cs2TiCl6, MASnBr3, and Cs2TiF6. Table provides the simulation parameters and values for the nontoxic perovskite absorber material.
6. Simulation Parameters of Lead-Free Nontoxic Perovskite Absorber Materials.
| material | thickness (μm) | E g (eV) | χ (eV) | εr | N c (cm–3) | N v (cm–3) | μn (cm2/(V s)) | μp (cm2/(V s)) | N D (cm–3) | N A (cm–3) | N t (cm–3) |
|---|---|---|---|---|---|---|---|---|---|---|---|
| MASnI3 | 0.5 | 1.3 | 4.2 | 10 | 1 × 1018 | 1 × 1018 | 1.6 × 100 | 1.6 × 100 | 0 | 1 × 1015 | 1 × 1014 |
| CsSnI3 | 0.5 | 1.27 | 4.47 | 18 | 1.8 × 1019 | 2.2 × 1018 | 4.37 × 100 | 4.37 × 100 | 0 | 1 × 1015 | 1 × 1014 |
| FASnI3 | 0.5 | 1.41 | 3.52 | 8.2 | 1 × 1018 | 1 × 1018 | 2.2 × 101 | 2.2 × 101 | 0 | 1 × 1015 | 1 × 1014 |
| MAGeI3 | 0.5 | 1.9 | 3.98 | 10 | 1 × 1016 | 1 × 1015 | 1.62 × 101 | 1.01 × 101 | 1 × 1015 | 0 | 1 × 1014 |
| Cs3Bi2I9 | 0.5 | 2.03 | 3.4 | 9.68 | 4.98 × 1019 | 2.11 × 1019 | 4.3 × 100 | 1.7 × 100 | 0 | 1 × 1015 | 1 × 1014 |
| Cs2AgBiBr6 | 0.5 | 2.05 | 4.19 | 5.8 | 1 × 1016 | 1 × 1016 | 1.181 × 101 | 4.9 × 10–1 | 0 | 1 × 1015 | 1 × 1014 |
| Cs2TiI6 | 0.5 | 1.8 | 4.2 | 18 | 1 × 1019 | 1 × 1019 | 4.4 × 100 | 2.5 × 100 | 1 × 1015 | 0 | 1 × 1014 |
| Cs2TiBr6 | 0.5 | 1.6 | 4.47 | 10 | 1 × 1019 | 1 × 1019 | 4.4 × 100 | 2.5 × 100 | 1 × 1015 | 0 | 1 × 1014 |
| Cs2TiCl6 | 0.5 | 2.23 | 4 | 19 | 1 × 1019 | 1 × 1019 | 4.4 × 100 | 2.5 × 100 | 1 × 1015 | 0 | 1 × 1014 |
| MASnBr3 | 0.5 | 2.15 | 4.17 | 10 | 2.5 × 1019 | 2.5 × 1019 | 1.6 × 100 | 1.6 × 100 | 1 × 1015 | 0 | 1 × 1014 |
| Cs2TiF6 | 0.5 | 1.9 | 4.3 | 18 | 1 × 1019 | 1 × 1019 | 4.4 × 100 | 2.5 × 100 | 1 × 1015 | 0 | 1 × 1014 |
The selection of PEDOT:PSS + WO3 as the hole transport layer (HTL) and ZnO as the electron transport layer (ETL) in this study is grounded in both economic and scientific rationale. While their affordability significantly reduces total device cost, these materials were primarily chosen for their favorable energy level alignment (Figure ) with a broad range of perovskite absorber materialsboth toxic (Pb-based) and nontoxic. Band alignment is a critical criterion for efficient charge extraction and reduced interface recombination, and both PEDOT:PSS + WO3 and ZnO offer conduction and valence band positions that are well-matched to the energy levels of the 14 absorber materials considered in this work. This compatibility ensures efficient hole and electron transport across the absorber interfaces, as supported by the simulated band alignment diagrams included in the study. Finding HTM/ETM combinations that are simultaneously low-cost and capable of maintaining optimal energy band offsets (minimal VBO and CBO) across diverse absorber types is highly nontrivial. PEDOT:PSS + WO3 and ZnO provide this rare combination of energetic compatibility, material stability, and economic feasibility, justifying their consistent use in all simulated architectures. This strategic constraint also ensures that absorber performance comparisons are fair, isolating the effects of the absorber material itself without variations introduced by transport layers.
2.
Band alignment diagram of materials used in this study.
4. Results and Discussion
In this study, optimal values were used for the simulation, making the process of identifying a highly efficient absorber layer more reliable. The output parameters, including open-circuit voltage (V oc), short-circuit current density (J sc), fill factor (FF), and efficiency (η) for different absorber-based perovskite solar cells (PSCs), are presented in Tables and . These parameters are crucial, as the efficiency is directly dependent on them, and they share an interdependent relationship, as expressed in eq . The equation includes P in as the input power:
| 1 |
7. Output Parameters of Perovskite Absorber Materials Containing Pb.
| absorber | V oc (V) | J sc (mA/cm2) | FF (%) | η (%) |
|---|---|---|---|---|
| MAPbI3 | 1.05 | 28.54 | 84.99 | 25.49 |
| MAPbI3(1–x)Br3 x | 1.11 | 22.38 | 84.48 | 20.97 |
| MAPbI3–x Cl x | 0.91 | 26.28 | 82.58 | 19.78 |
The results of this study were evaluated by considering the presence of lead (Pb). It is widely recognized that Pb is a toxic heavy metal that poses severe health risks to humans and is detrimental to all living organisms, including plants and animals. , The output parameters such as open-circuit voltage (V oc), short-circuit current density (J sc), fill factor (FF), and eta (η) obtained from the simulation for different Pb-based absorber PSC are shown in Table .
Table shows that among the lead-based perovskites, MAPbI3(1–x)Br3 x achieves the highest open-circuit voltage (V oc) of 1.11 V, while MAPbI3–x Cl x exhibits the lowest V oc at 0.91 V. The higher V oc observed in MAPbI3(1–x)Br3 x is attributed to bandgap widening resulting from the substitution of iodide with bromide ions, which raises the conduction band and lowers the valence band. This increased bandgap enhances the built-in potential and reduces recombination, thus elevating V oc. However, the same bandgap increase leads to reduced absorption in the longer-wavelength region of the solar spectrum, causing a significant drop in short-circuit current density (J sc), as seen in the lower J sc of 22.38 mA/cm2 for MAPbI3(1–x)Br3 x compared to 28.54 mA/cm2 for MAPbI3. This reflects the classic V oc–J sc trade-off governed by the Shockley–Queisser limit. Although fill factor (FF) values are comparable across compositions, MAPbI3 records the highest FF at 84.99%, likely due to optimal band alignment and minimal series resistance in the device configuration. In contrast, MAPbI3–x Cl x has a slightly lower FF (82.58%), potentially due to greater energy mismatch and increased interface recombination. Based on its balanced performance, with high J sc, V oc, and FF values, MAPbI3 is selected for further optimization and simulation.
The performance of lead-free perovskite solar cells (PSCs), summarized in Table , reveals a broad range of photovoltaic characteristics. V oc values span from 0.579 V (Cs2TiBr6) to 1.24 V in MAGeI3, primarily influenced by differences in the bandgap energy and energy level alignment. While MAGeI3’s wide bandgap accounts for its high V oc, it limits light absorption in the lower-energy spectrum, resulting in a reduced J sc. In contrast, CsSnI3 achieves the highest J sc (33.90 mA/cm2) owing to its narrower bandgap (∼1.3 eV), which facilitates broad-spectrum absorption and efficient photogeneration. However, despite its high current density, CsSnI3 exhibits a significantly lower fill factor (65.51%). This is primarily attributed to its poor carrier transport properties and higher recombination rates as simulated in the SCAPS-1D environment. Specifically, CsSnI3 has a relatively high intrinsic carrier concentration and shallow energy levels, which can enhance carrier recombination under operating conditions. Additionally, the simulated defect density and low built-in potential can lead to an increased ideality factor, causing early voltage saturation and limiting the maximum power point. These factors collectively degrade the FF, despite a high photocurrent. In contrast, Cs3Bi2I9 shows a higher FF (85.50%) due to better band alignment and simulated interface properties, though with a lower J sc. Among all candidates, MASnI3 provides the most balanced performance, with high J sc, moderate V oc, and stable FF, resulting in the highest simulated efficiency (24.08%). It is therefore selected for further optimization.
8. Output Parameters of a Perovskite Absorber that Does Not Contain Toxic Pb.
| absorber | V oc (V) | J sc (mA/cm2) | FF (%) | eta (%) |
|---|---|---|---|---|
| MASnI3 | 0.93 | 35.14 | 73.54 | 24.08 |
| MAGeI3 | 1.24 | 17.81 | 73.06 | 16.13 |
| FASnI3 | 0.63 | 31.42 | 78.39 | 15.60 |
| Cs3Bi2I9 | 1.04 | 16.40 | 85.50 | 14.54 |
| CsSnI3 | 0.62 | 35.90 | 65.51 | 14.52 |
| Cs2AgBiBr6 | 1.04 | 15.35 | 76.25 | 12.18 |
| Cs2TiCl6 | 1.07 | 14.20 | 74.41 | 11.31 |
| Cs2TiI6 | 0.87 | 18.78 | 69.05 | 11.30 |
| Cs2TiBr6 | 0.61 | 24.82 | 63.44 | 9.59 |
| MASnBr3 | 0.87 | 14.80 | 72.57 | 9.36 |
| Cs2TiF6 | 0.77 | 17.88 | 67.27 | 9.27 |
Figures and present the initial quantum efficiency (QE) and J–V characteristics of the MAPbI3 and MASnI3-based cells, respectively. In Figure , the MAPbI3 cell achieves a peak QE of approximately 99% at a wavelength near 570 nm, with QE remaining above 85% initially. The QE declines sharply beyond 820 nm, limiting the carrier collection from longer-wavelength photons. Similarly, Figure shows the MASnI3-based cell starting at ∼85% QE, peaking at around 99% near 550–600 nm, and then dropping to nearly zero beyond 950 nm, consistent with its narrower bandgap. These trends indicate that both cells are optimized for midvisible spectrum absorption but suffer from low edge-of-spectrum QE due to recombination losses, weaker internal electric field, and incomplete absorption near the band edges. The relatively low initial QE at the spectral extremes can also be linked to nonoptimized layer thickness and interface mismatch, which reduce photon-to-electron conversion efficiency. These observations reinforce the need for parameter tuning in absorber and transport layers to improve the overall spectral response.
3.
(a) Initial J–V curve and (b) QE curve of the MAPbI3-based cell.
4.
(a) Initial J–V curve and (b) QE curve of the MASnI3-based cell.
In this phase of the study, the P+PIN+ structure was implemented. A thin P+ layer was utilized as a back surface field (BSF) to create near-ohmic contact and enhance the cell performance by reducing the surface recombination velocity (SRV). Additionally, a thin intrinsic layer of the absorber material was employed to form an interface between the absorber and the N+ layer, effectively minimizing recombination losses. The P+P layer was engineered by grading the materials based on doping density, whereas the N+ region was achieved by increasing the doping concentration.
These modifications are expected to enhance the overall performance of the PSCs, leading to higher efficiencies and improved stability, thus making lead-free PSCs a viable alternative to their lead-based counterparts.
4.1. Effect of Thickness
The absorber layer is a key factor for achieving high efficiency in perovskite solar cells (PSCs). There is a crucial relationship between the output parameters of the cell and the thickness of the absorber layer. As the thickness increases, more light is absorbed, leading to the generation of a greater number of electron–hole pairs. This enhances the short-circuit current density (J sc) due to increased photogeneration. However, an increase in thickness also introduces competing effects: it can reduce recombination at the back contact due to a stronger back surface field, but it simultaneously increases the bulk recombination probability due to longer carrier diffusion paths, which may ultimately reduce efficiency. In this part of the study, both MAPbI3 and MASnI3-based cells were simulated with thickness values ranging from 0.1 to 1 μm.
According to Figure , for the MAPbI3-based cell, the efficiency increases with a thickness up to 0.325 μm, where it peaks and then decreases beyond this point. J sc shows a steady increase with thickness, achieving a maximum of 28.6 mA/cm2 at 1 μm due to more complete absorption of incident photons. On the other hand, V oc decreases from 1.125 V at 0.1 μm to 1.07 V at 1 μm because thicker absorbers increase the likelihood of bulk recombination, which lowers the quasi-Fermi level splitting and suppresses V oc. The fill factor (FF) initially decreases due to increased resistance and carrier transport inefficiency but shows a mild recovery at higher thicknesses as the electric field strength improves. The FF peaks at 86.5% at 0.1 μm and dips to a minimum of 85.07% at 0.55 μm. The optimal balance of these parameters is achieved at 0.325 μm, where V oc = 1.10 V, J sc = 28.47 mA/cm2, FF = 85.42%, and efficiency = 26.71%.
5.
Open-circuit voltage V oc (V), short-circuit current J sc (mA/cm2), fill factor FF (%), and efficiency η (%) curves for the MAPbI3-based cell according to thickness.
From the simulation results presented in Figure , it is evident that at a thickness of 0.1 μm, the fill factor and open-circuit voltage are maximized, with values of 0.99 V and 82.88%, respectively. However, the efficiency and J sc are minimized at this thickness, with values of 31.32 mA/cm2 and 25.74%, respectively. The short-circuit current density increases with thickness, while the open-circuit voltage decreases. Efficiency initially increases with thickness and then decreases beyond a certain point, following the observed variation in the thickness. The fill factor initially decreases with increasing thickness but increases again after a specific threshold. The optimal thickness for the MASnI3-based cell is found to be 0.25 μm, yielding maximum efficiency. The output parameters for this optimized thickness are a V oc of 0.97 V, a J sc of 34.81 mA/cm2, an FF of 80.91%, and an efficiency of 27.95%.
6.
Open-circuit voltage V oc (V), short-circuit current J sc (mA/cm2), fill factor FF (%), and efficiency η (%) curves for the MASnI3-based cell according to thickness.
These observations underline that optimizing absorber thickness is essential to achieving the best trade-off between light absorption and recombination loss. Thin absorbers suffer from insufficient generation, while overly thick absorbers introduce resistive and recombination penalties. Therefore, the identified optimal thicknesses0.325 μm for MAPbI3 and 0.25 μm for MASnI3maximize carrier collection while minimizing performance losses from recombination and transport resistance, aligning well with previous theoretical findings. ,
4.2. Effect of Temperature
Previous simulations were conducted at a fixed temperature of 300 K. However, in real-world conditions, the operating temperature of perovskite solar cells (PSCs) varies due to environmental and thermal loads. Therefore, in this study, temperature-dependent simulations were performed over a range of 260–400 K to investigate the thermal behavior of both MAPbI3 and MASnI3-based devices.
As illustrated in Figure , MAPbI3-based cells exhibit linear performance degradation with increasing temperature. Specifically, the V oc, fill factor (FF), and efficiency decrease consistently, while J sc initially drops slightly and then stabilizes. This behavior is attributed to enhanced carrier recombination, thermal activation of traps, and ion migration at elevated temperatures, which reduce the built-in potential and quasi-Fermi level splitting. ,
7.
Open-circuit voltage V oc (V), short-circuit current J sc (mA/cm2), fill factor FF (%), and efficiency η (%) curves for the MAPbI3-based cell according to temperature.
The increased lattice vibration also raises series resistance and impedes charge extraction, further reducing the FF. These trends are typical for hybrid perovskites and validate the need for temperature-tolerant cell architectures.
In contrast, the MASnI3-based cell exhibits nonlinear temperature dependence, as shown in Figure . Meanwhile, V oc and efficiency also decline steadily with temperature, consistent with thermal recombination dynamics. , The trends in FF and J sc are distinctly nonlinear. At a lower temperature (260 K), FF is relatively low, likely due to insufficient thermal energy for effective charge extraction and activation of shallow defect states. As the temperature increases, FF improves, reaching a peak at an intermediate temperature (∼320–330 K), before declining again due to thermally induced recombination and resistive losses. This rise and fall in FF suggest a temperature-activated transport regime, where moderate thermal energy enhances carrier mobility and reduces interface resistance but excessive heat reverses these gains.
8.
Open-circuit voltage V oc (V), short-circuit current J sc (mA/cm2), fill factor FF (%), and efficiency η (%) curves for the MASnI3-based cell according to temperature.
Similarly, the short-circuit current density (J sc) follows an inverted parabolic trend: it starts high at 260 K, decreases to a local minimum at intermediate temperature, and then increases again at higher temperatures, peaking near 400 K. This is due to two competing effects: at lower temperatures, the carrier lifetime is sufficient to support high J sc, but poor mobility limits extraction. At intermediate temperatures, increased recombination and interface barriers reduce the level of J sc. At higher temperatures, thermally assisted transport enhances carrier collection again, despite increased recombination.
These nonlinear behaviors in MASnI3 arise from its higher sensitivity to defect activation, self-doping effects, and narrower bandgap, which make it more responsive to thermal changes compared to MAPbI3. The complex temperature dependence of J sc and FF underscores the importance of designing MASnI3-based PSCs with careful thermal management and defect passivation to ensure a stable performance under operational fluctuations.
In summary, while both cell types degrade in efficiency as temperature increases, the nonmonotonic behavior of FF and J sc in MASnI3 reveals a more complex interplay of thermal activation, recombination, and transport phenomena. These insights emphasize the necessity for thermal optimization in device engineering, especially for Sn-based, lead-free perovskites, to enhance the long-term stability and performance reliability under real-world conditions.
4.3. Effect of Absorber Layer Electron Affinity
Electron affinity (χ) is a critical parameter in device modeling as it directly influences the energy band alignment between various layers in a device. Specifically, it plays a pivotal role in determining the valence band offset (VBO) and conduction band offset (CBO), which are key factors for optimizing charge carrier transport and minimizing recombination losses. These parameters are mathematically expressed as
| 2 |
| 3 |
Here, χPER is the electron affinity of the absorber material; χHTL and χETL are the electron affinities of the hole transport layer (HTL) and electron transport layer (ETL), respectively. E g HTL is the bandgap of the HTL, and E g PER denotes the bandgap of the absorber layer.
The electron affinity of absorber materials significantly impacts the overall device performance by aligning the energy levels of adjacent layers, thus optimizing charge transport and reducing potential energy barriers. These observations align with the findings Wang et al., who demonstrated the correlation between energy level alignment and device efficiency.
Simulation results in Figure show that by varying the electron affinity (χ) of the absorber material between 3.0 and 4.5 eV, the performance of perovskite-based solar cells can be substantially tuned. For the MAPbI3-based cell, the efficiency increases with electron affinity and reaches a maximum of 27.10% at 3.75 eV. Similarly, the MASnI3-based cell peaks at 27.98% efficiency at 4.00 eV. These optimized points confirm that proper alignment of energy levels between the absorber and transport layers is essential for maximizing charge extraction and minimizing recombination losses, in agreement with a previous study.
9.

Efficiency (%) curves of MAPbI3 and MASnI3-based cells according to electron affinity (eV).
The electron affinity of the absorber directly influences the conduction band offset (CBO) at the interface with the electron transport material (ETM), here ZnO. This band offset determines whether the device exhibits a “spike-type” or “cliff-type” alignment. A spike-type alignment occurs when the ETM conduction band is lower than that of the absorber, creating a small energy barrier that can block minority carrier injection and suppress recombinationbeneficial if the spike is moderate (<0.3 eV). A cliff-type alignment, where the ETM conduction band lies above that of the absorber and promotes majority carrier flow but increases the risk of interface recombination due to a lack of an energy barrier.
In this study, both MAPbI3 and MASnI3 exhibit changes in performance, consistent with transitions between these two alignment types. For MAPbI3, the optimal electron affinity of 3.75 eV yields a near-zero or slightly positive CBO with ZnO, suggesting a mild spike-type alignment, which favors efficient electron extraction, while preventing recombination. At lower electron affinities (<3.5 eV), the alignment shifts toward a cliff, increasing recombination and reducing performance. However, MAPbI3 maintains a stable output due to its relatively robust interfacial electronic structure and defect tolerance.
In contrast, MASnI3 is far more sensitive to the electron affinity shift. As the electron affinity increases beyond 4.00 eV, the conduction band of MASnI3 sinks below that of ZnO, creating a larger spike that begins to obstruct electron transport and reduce performance. Below this value, the system trends toward a cliff-type alignment, which again promotes recombination. This explains the sharper performance peak in MASnI3 and its nonlinear response: small deviations in χ can quickly move the CBO into an unfavorable range, either by introducing a too-steep spike or a recombination-prone cliff. Therefore, the optimal point (χ = 4.00 eV) likely corresponds to a minimal positive CBO, ensuring a balanced band alignment that maximizes extraction, while limiting interfacial losses.
These findings highlight the importance of band alignment engineering. For MAPbI3, performance is relatively resilient across a range of CBOs, but for MASnI3, precise electron affinity tuning is critical to avoid spike/cliff extremes and to ensure device stability and efficiency.
4.4. Effect of Back Contact
Back contact plays a vital role in the device performance of PSCs. The efficiency of the cell increases with the increment of the metal work function of back contact. In this section, two cells are simulated by 10 different back contact materials. The metal work function of different back contact materials is given in Table .
9. Different Back Contact Materials and Their Metal Work Function.
| back contact material , | Cu | Ag | Fe | C | Au | W | Ni | Pt | Pd | Se |
|---|---|---|---|---|---|---|---|---|---|---|
| metal work function, Φm (eV) | 4.65 | 4.74 | 4.81 | 5 | 5.1 | 5.22 | 5.5 | 5.6 | 5.7 | 5.9 |
By investigation of Figure , it can be said that the efficiency is increasing with the metal work function (eV) of back contact, and after a particular value of the metal work function (eV), the efficiency is almost constant.
10.

Efficiency Curve of MAPbI3 and MASnI3-based cells according to back contact metal work function (eV).
Here, six back contacts can retain maximum efficiency for both optimized cells. As we aim to determine a low-cost and highly efficient cell structure, we choose the back contact material according to the price of the materials from these six back contact materials.
Table clearly identifies the least expensive back contact materials. Although all six back contacts can provide maximum efficiency, Ni is the least costly back contact among them. Hence, we chose Ni as the back contact for our study to improve the cost-effectiveness of the cells. Contact parameters after optimization are given in Table .
10. Back Contact Materials that Provide Maximum Efficiency and Their Price .
| metal name | metal work function (eV) | price (euro/g) |
|---|---|---|
| Au | 5.1 | 264 |
| W | 5.22 | 13.5 |
| Ni | 5.5 | 2.232 |
| Pd | 5.6 | 635.714 |
| Pt | 5.7 | 230 |
| Se | 5.9 | 11.32 |
The price of each material is taken from Sigma-Aldrich.
11. Back Contact and Front Contact Parameters.
| parameters | back contact | front contact |
|---|---|---|
| surface recombination velocity of electrons (cm/s) | 1.000 × 105 | 1.000 × 107 |
| surface recombination velocity of holes (cm/s) | 1.000 × 107 | 1.000 × 105 |
| metal work function (eV) | 5.5 | 4.3 |
| majority carrier barrier height relative to E f (eV) | ||
| MAPbI3 | –0.3000 | 0.2000 |
| MASnI3 | –0.3000 | 0.2000 |
| majority carrier barrier height relative to E v (eV) | ||
| MAPbI3 | –0.3749 | 0.1641 |
| MASnI3 | –0.3749 | 0.1641 |
4.5. Effect of the Absorber Layer Defect Density
Defect density optimization is an optimization technique to increase cell performance. There is a direct effect of the defect density on device output parameters. The recombination rate is increased with an increment of defect density, which reduces the carrier lifetime and diffusion length of the carrier. Hence, the efficiency is also deduced. In this section, both cells are simulated with a defect density of 1.00 × 1010 (1/cm3) to 1.00 × 1018 (1/cm3).
In the simulated MAPbI3-based perovskite solar cell (where L p = L n and τp = τn), both the diffusion length and carrier lifetime decreased with an increase in defect density. The highest power conversion efficiency was observed when the diffusion length ranged from 510.0 to 16.0 μm and the carrier lifetime ranged from 1.00 × 107 to 1.00 × 104 s. The optimal performance was achieved at a defect density of 1.00 × 1013 cm–3, corresponding to an efficiency of 27.1% (Figure ).
11.
(a) Impact of the defect density on the diffusion length, (b) impact of the defect density on the carrier lifetime, and (c) impact of defect density on the power conversion efficiency of the MAPbI3-based cell.
Figure shows that at a low defect density of 1.00 × 1010 cm–3, the values of L p, L n, τp, and τn were 64.0 μm, 640.0 μm, 1.00 × 106 s, and 1.00 × 108 s, respectively, with a maximum efficiency of 27.99%. As the defect density increased, all these parameters declined consistently, confirming that higher defect densities negatively impact device performance in the MASnI3-based perovskite system.
12.
(a) Impact of defect density on the diffusion length, (b) impact of the defect density on the carrier lifetime, and (c) impact of the defect density on the power conversion efficiency of the MASnI3-based cell.
The efficiency is the lowest (9.56%) for the MASnI3-based cell when N t is at maximum. We have chosen the defect density value as 1.00 × 1013 (1/cm3), where the efficiency is 27.99%.
4.6. Effect of Series Resistance
Series resistance (R s) is a fundamental parameter that significantly impacts the performance of photovoltaic devices. It primarily affects the short-circuit current density (J sc) and fill factor (FF), which collectively determine the power conversion efficiency (PCE). An increase in R s introduces resistive losses that hinder charge carrier transport, reduce the fill factor, and limit the overall charge extraction efficiency. These effects ultimately lead to a noticeable decline in device efficiency, as supported by previous studies. ,
Simulation results, as illustrated in Figure , reveal that the efficiency (η) of the device consistently decreases with increasing R s. Under ideal conditions (R s = 0 Ω), a maximum efficiency of 27.87% is achieved. As R s increases to 6 Ω, the efficiency declines to 23.31%. Notably, the open-circuit voltage (V oc) appears to show a slight increase, which may seem counterintuitive. In practical devices, V oc is predominantly determined by the separation of quasi-Fermi levels and is typically not influenced by R s, especially under open-circuit conditions where the current is zero and series resistance has no direct electrical effect.
13.

Efficiency and open-circuit voltage curves of the MAPbI3-based cell according to series resistance.
The minor increase in simulated V oc with increasing R s is likely a numerical artifact or a consequence of changes in internal recombination dynamics captured by the SCAPS-1D simulation environment rather than a physically meaningful behavior. It may result from slight shifts in internal carrier distributions or reduced recombination current under zero current conditions. However, in real photovoltaic devices, V oc either remains constant or slightly decreases with increasing R s due to indirect effects, such as elevated recombination caused by higher internal resistive heating. Therefore, the observed trend in V oc should not be interpreted as a performance benefit.
As shown in Figure , both J sc and FF decrease steadily with increasing the value of R s. At R s = 6 Ω, J sc drops to 28.47 mA/cm2, and FF falls to 73.34%, clearly demonstrating the impact of resistive losses on current extraction and voltage efficiency.
14.

Short-circuit current and fill factor curves of a MAPbI3-based cell according to series resistance.
Recognizing that eliminating R s entirely is not feasible in real devices, a practical value of R s = 1 Ω was selected for further simulation and optimization. Under this condition, MAPbI3-based cells yield optimized values of V oc = 1.10 V, J sc = 28.47 mA/cm2, FF = 86.42%, and η = 27.10%. For MASnI3-based cells, the corresponding values are V oc = 0.97 V, J sc = 34.89 mA/cm2, FF = 82.51%, and η = 27.98%.
Figure further provides the outcome of impact of series resistance on the performance of the MASnI3 based cell. When R s = 0 Ω, the device achieves a maximum efficiency of 29.1%, while with R s = 6 Ω, the efficiency sharply drops to 22.45%, primarily due to reductions in FF and J sc. V oc remains mostly unchanged across the R s range, as expected.
15.

Efficiency and open-circuit voltage curve of a MASnI3-based cell according to series resistance.
These findings underscore the necessity of minimizing series resistance in PSCs to maintain high FF and efficiency, in agreement with established experimental and theoretical research. ,
The simulation results presented in Figure indicate that both short-circuit current density (J sc) and fill factor (FF) decline consistently as the series resistance (R s) increases from 1 to 6 Ω. This trend reflects the growing impact of resistive losses, which hinder charge carrier extraction and reduce the device’s power output. Given that real-world photovoltaic devices always exhibit some finite resistance, a practical value of R s = 1 Ω is selected for optimization, representing a realistic balance between ideal behavior and manufacturability. Under these conditions, the MASnI3-based cell achieves optimal output parameters: V oc = 0.97 V, J sc = 34.89 mA/cm2, FF = 82.51%, and power conversion efficiency (PCE) = 27.98%. These values confirm that while zero-resistance conditions are useful for theoretical benchmarking, low but realistic R s values must be targeted to ensure both high performance and practical feasibility in perovskite solar cell deployment.
16.

Short-circuit current and fill factor curve of a MASnI3-based cell according to series resistance.
From the above discussion, it can be said that both cells show optimized efficiency, which is also practically achievable. According to the Shockley–Queisser limit, the maximum efficiency achievable for a single-junction solar cell is 33.14%, and the efficiency of both MAPbI3 and MASnI3-based cells does not exceed the limit. The previous study also confirmed the achievability of these efficiencies as MAPbI3-based achieved a power conversion efficiency of 26.74%, and MASnI3-based cells achieved a 27.43% power conversion efficiency in the previous study.
Here, both cells are optimized and can provide optimal efficiency. The quantum efficiency of both cells in Figure clearly identifies that both cells are optimized. Here, both cells show a nearly 87% quantum efficiency. However, the quantum efficiency of both cells reaches almost 100% at the midrange of the wavelength and drops significantly at the final range of wavelength.
17.

Quantum efficiency of the optimized cells.
Therefore, the simulation result and price of the materials confirm that the proposed cell can provide a higher efficiency with the least expensive ETM, HTM, and back contact materials.
4.7. Performance Analysis
This section of the study discusses performance improvement and the reasons behind the improvement. Here, the reason behind the overall improvement is shown and analyzed through initial and optimized cell properties such as generation current, recombination current, generation profile, recombination profile, and band diagram of the cell.
For MAPbI3-based cells, according to Figure , the initial and optimized generation and recombination curve shows that the overall performance of the optimized cell is increased compared to the initial cell. Although the recombination current did not drop significantly, generation current has grown considerably on the optimized cell. Hence, the overall cell performance of the optimized cell performance has also increased.
18.
Generation and recombination current of the MAPbI3-based cell.
The reason for overall performance, generation current, and recombination current ratio improvement can be described in Figure . In Figure , the X (μm) indicates the depth within the cell structure, starting from the front contact and extending through the absorber to the back contact. Figure reveals the electron–hole pair generation and recombination on the initial and optimized cells. It can be observed that, although the generation rates did not rise significantly, the ratio between generation and recombination improved significantly. A broader picture in Figure , described by Figure , defines that.
19.
Generation–recombination profile of the MAPbI3-based cell.
20.

(a) Generation–recombination profile of the initial MAPbI3-based cell and (b) generation–recombination profile of the optimized MAPbI3-based cell.
The reason for the improvement in the ratio between generation and recombination ratio can be described in Figure . In Figure , X (μm) indicates the depth within the cell structure, starting from the front contact and extending through the absorber to the back contact. On point A, the back surface field is generated between the back contact and HTM due to the optimized back contact of the cell. This back surface field reduces the recombination between the HTM and back contact. On points B and C, VBO and CBO are induced due to the introduction of a thin (30 nm) intrinsic absorber layer and optimization of the electron affinity of the absorber layer. These VBO and CBO reduce the recombination on the interface between the absorber and ETM. Additionally, the interface between the HTM and absorber is also improved. Hence, the performance of the optimized cell is improved compared with the initial cell.
21.
Energy band diagram of the MAPbI3-based cell.
The MASnI3-based cell also has similar impacts as the MAPbI3-based cell for optimizing parameters. Initial cells have comparatively lower generation currents. However, the generation current improved significantly after optimizing material and cell parameters (Figure ). Hence, the overall cell performance has also improved considerably. The reason behind the generation’s current improvement can be described by the generation–recombination profile of the cell. The generation–recombination profile of the cell is shown in Figure . In Figure , X (μm) indicates the depth within the cell structure, starting from the front contact and extending through the absorber to the back contact. It can be seen from Figure that overall carrier recombination has dropped significantly compared with carrier generation. A broader picture of the generation recombination profile in Figure identifies the numerical comparison between the initial and optimized cell generation recombination profiles. The reason for the generation recombination profile can be described by the initial and optimized cell band diagram. In Figure , X (μm) indicates the depth within the cell structure, starting from the front contact and extending through the absorber to the back contact. Figure shows that the initial cell has no optimized VBO and CBO, which results in higher recombination after optimizing the cell by varying cell parameters and materials parameters and introducing intrinsic absorber materials between absorber materials. The interface recombination points B, C, D, and E can be reduced significantly. While the probability of recombination on point A is reduced by optimizing back contact materials.
22.
Generation and recombination current of the MASnI3-based cell.
23.
Generation–recombination profile of the MASnI3-based cell.
24.

(a) Generation–recombination profile of the initial MASnI3-based cell and (b) generation–recombination profile of the optimized MASnI3-based cell.
25.
Energy band diagram of the MASnI3-based cell.
After continuous simulation and optimization, the following cell structure of Figure is finalized for the study.
26.

Optimized cell structures for the study.
Table clearly indicates that cell structures of the current study clearly outplay previous cell structures as these two cells can provide high efficiency while the materials cost is lowest for these cells. Hence, these cells are noble cell structures for further study and fabrication.
12. Cost of Materials and Efficiency Comparison of Cell Structures (Current and Previous Studies) .
| cell structure | ETM cost (euro/g) | HTM cost (euro/g) | back contact cost (euro/g) | total cost (euro) | η (%) |
|---|---|---|---|---|---|
| FTO/TiO2/CsGeI3/Spiro-OMeTAD/Ag | 27.2 | 528 | 9.01 | 564.21 | 18.30% |
| ZnO(nr)/CH3NH3SnI3/Cu2O/Au | 13.6 | 15.76 | 264 | 293.36 | 20.23% |
| ZnO(nr)/CH3NH3SnI3/Spiro-OMeTAD/Au | 13.6 | 528 | 264 | 805.6 | 20.21% |
| ITO/NiO/MASnI3/PCBM/Ag | 19.7 | 1130 | 9.01 | 1158.71 | 29.19% |
| FTO/Perovskite/Spiro-OMeTAD/Au | 0 | 528 | 264 | 792 | 19.16% |
| ITO/PEDOT:PSS/MAGeI3/PCBM/Ag | 1.324 | 1130 | 9.01 | 1140.334 | 11.16% |
| ITO/PEDOT:PSS/MAGeI3/C60/ Ag | 1.324 | 677 | 9.01 | 687.334 | 13.5% |
| FTO/TiO2/MAPbI3/Spiro-OMeTAD/Au | 27.2 | 528 | 264 | 819.2 | 23.18% |
| FTO/TiO2/MAPbI3/NPB/Au | 27.2 | 725 | 264 | 1016.2 | 22.56% |
| FTO/TiO2/MAPbI3/P3HT/Au | 27.2 | 749 | 264 | 1040.2 | 19.38% |
| FTO/TiO2/MAPbI3/MEH-PPV/Au | 27.2 | 798 | 264 | 1089.2 | 21.23% |
| ITO/NiO/CH3NH3PbI3–x Cl x /PCBM/Ag | 19.7 | 1130 | 9.01 | 1158.71 | 22% |
| FTO/TiO2/CH3NH3PbI3–x Cl x /Spiro-OMeTAD/Ag | 27.2 | 528 | 9.01 | 564.21 | 29.56% |
| FTO/ZnO/MASnI3/PEDOT:PSS/Au | 0.172 | 1.324 | 264 | 265.496 | 21.22% |
| ZnO/MAPbI3/PEDOT:PSS + WO3/Ni | 0.172 | 1.292 | 2.232 | 3.696 | 27.10 (this study) |
| ZnO/MASnI3/PEDOT:PSS + WO3/Ni | 0.172 | 1.292 | 2.232 | 3.696 | 27.98 (this study) |
The price of each material is taken from Sigma-Aldrich.
5. Conclusions
This study presents a comprehensive, simulation-guided framework for designing high-efficiency, low-cost, and environmentally sustainable perovskite solar cells (PSCs). Using SCAPS-1D, we evaluated 14 absorber materials that are both Pb-based and lead-free within a fixed, low-cost device architecture featuring ZnO (ETM), PEDOT:PSS + WO3, and nickel (Ni) as the back contact. The optimized configurations yielded power conversion efficiencies of 27.10% for MAPbI3 and 27.98% for MASnI3, confirming that nontoxic MASnI3 can rival conventional lead-based absorbers in performance.
A key contribution of this work is the simultaneous optimization of efficiency, affordability, and environmental safetythree factors typically explored in isolation. By tuning the absorber thickness, electron affinity, defect density, and series resistance, we achieved enhanced carrier generation, reduced recombination, and favorable energy level alignment across a wide spectral range.
These results establish a practical blueprint for building scalable PSCs using abundant and low-cost materials. The simulation-driven approach also reduces experimental trial and error, accelerating the discovery and validation of efficient perovskite architectures.
Looking ahead, future research should focus on experimental validation, long-term operational stability, and scalable manufacturing. By addressing performance, cost, and toxicity in a unified design, this study offers a crucial step toward next-generation solar technologies that are both commercially viable and environmentally responsible.
Acknowledgments
The authors gratefully acknowledge the use of SCAPS-1D, a one-dimensional solar cell simulation program developed by Marc Burgelman, Philippe Nollet, and Staf Degrave at the University of Ghent. Their contribution has been instrumental in enabling the simulations conducted in this study.
The parameter data used to support the findings of this study are included in the article.
The authors did not receive any external funding for the study.
The authors declare no competing financial interest.
References
- Kojima A., Teshima K., Shirai Y., Miyasaka T.. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009;131(17):6050–6051. doi: 10.1021/ja809598r. [DOI] [PubMed] [Google Scholar]
- Min H., Lee D. Y., Kim J., Kim G., Lee K. S., Kim J., Paik M. J., Kim Y. K., Kim K. S., Kim M. G., Shin T. J., Seok S. II. Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes. Nature. 2021;598:444–450. doi: 10.1038/s41586-021-03964-8. [DOI] [PubMed] [Google Scholar]
- Habibi M., Zabihi F., Ahmadian-Yazdi M. R., Eslamian M.. Progress in emerging solution-processed thin film solar cells – Part II: Perovskite solar cells. Renew. Sustain. Energy Rev. 2016;62:1012–1031. doi: 10.1016/j.rser.2016.05.042. [DOI] [Google Scholar]
- Luo S., Daoud W. A.. Recent progress in organic-inorganic halide perovskite solar cells: Mechanisms and material design. J. Mater. Chem. A. 2014;3:8992. [Google Scholar]
- Zheng K., Zhu Q., Abdellah M., Messing M. E., Zhang W., Generalov A., Niu Y., Ribaud L., Canton S. E., Pullerits T.. Exciton Binding Energy and the Nature of Emissive States in Organometal Halide Perovskites. J. Phys. Chem. Lett. 2015;6:2969–2975. doi: 10.1021/acs.jpclett.5b01252. [DOI] [PubMed] [Google Scholar]
- Kim H.-S., Lee C.-R., Im J.-H., Lee K.-B., Moehl T., Marchioro A., Moon S.-J., Humphry-Baker R., Yum J.-H., Moser J. E., Grätzel M., Park N.-G.. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9% Sci. Rep. 2012;2:591. doi: 10.1038/srep00591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Noh J. H., Im S. H., Heo J. H., Mandal T. N., Seok S. I.. Chemical Management for Colorful, Efficient, and Stable Inorganic–Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013;13:1764–1769. doi: 10.1021/nl400349b. [DOI] [PubMed] [Google Scholar]
- Leijtens T., Stranks S. D., Eperon G. E., Lindblad R., Johansson E. M. J., McPherson I. J., Rensmo H., Ball J. M., Lee M. M., Snaith H. J.. Electronic Properties of Meso-Superstructured and Planar Organometal Halide Perovskite Films: Charge Trapping, Photodoping, and Carrier Mobility. ACS Nano. 2014;8:7147. doi: 10.1021/nn502115k. [DOI] [PubMed] [Google Scholar]
- Stranks S., Eperon G., Grancini G., Menelaou C., Alcocer M., Leijtens T., Herz L., Petrozza A., Snaith H.. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science. 2013;342:341–344. doi: 10.1126/science.1243982. [DOI] [PubMed] [Google Scholar]
- Chen C.-W., Hsiao S.-Y., Chen C.-Y., Kang H.-W., Huang Z.-Y., Lin H.-W.. Optical Properties of Organometal Halide Perovskite Thin Films and General Device Structure Design Rules for Perovskite Single and Tandem Solar Cells. J. Mater. Chem. A. 2014;3:9152. [Google Scholar]
- Huang L., Sun X., Li C., Xu R., Xu J., Du Y., Wu Y., Ni J., Cai H., Li J., Hu Z., Zhang J.. Electron transport layer-free planar perovskite solar cells: Further performance enhancement perspective from device simulation. Sol. Energy Mater. Sol. Cells. 2016;157:1038–1047. doi: 10.1016/j.solmat.2016.08.025. [DOI] [Google Scholar]
- Zhang P., Xiong J., Chen W.-H., Du P., Song L.. Air-processed MAPbI 3 perovskite solar cells achieve 20.87% efficiency and excellent bending resistance enabled via a polymer dual-passivation strategy. Dalton Trans. 2023;52:15974–15985. doi: 10.1039/D3DT02080K. [DOI] [PubMed] [Google Scholar]
- Noman M., Shahzaib M., Jan S. T., Shah S. N., Khan A. D.. 26.48% efficient and stable FAPbI 3 perovskite solar cells employing SrCu 2 O 2 as hole transport layer. RSC Adv. 2023;13:1892–1905. doi: 10.1039/D2RA06535E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu N., Xu Z., Pan Y., Wang J., Liu Z., Zhang J., Huang L., Hu Z., Zhu Y., Liu X.. Efficient inverted CsPbI3 inorganic perovskite solar cells achieved by facile surface treatment with ethanolamine. Chem. Commun. 2023;59:8452–8455. doi: 10.1039/D3CC02084C. [DOI] [PubMed] [Google Scholar]
- Jayan K. D., Sebastian V.. Comprehensive device modelling and performance analysis of MASnI3 based perovskite solar cells with diverse ETM, HTM and back metal contacts. Sol. Energy. 2021;217:40–48. doi: 10.1016/j.solener.2021.01.058. [DOI] [Google Scholar]
- Rehman A ul, Afzal S., Naeem I., Bibi D., Sarwar S. G., Nabeel F., Obodo R. M.. Performance optimization of FASnI3 based perovskite solar cell through SCAPS-1D simulation. Hybrid Adv. 2024;7:100301. doi: 10.1016/j.hybadv.2024.100301. [DOI] [Google Scholar]
- Hossain M. K., Uddin M. S., Toki G. I., Mohammed M. K., Pandey R., Madan J., Rahman M. F., Islam M. R., Bhattarai S., Bencherif H., Samajdar D. P., Amami M., Dwivedi D. K.. Achieving above 24% efficiency with non-toxic CsSnI3 perovskite solar cells by harnessing the potential of the absorber and charge transport layers. RSC Adv. 2023;13:23514–23537. doi: 10.1039/d3ra02910g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karimi E., Ghorashi S. M. B.. Investigation of the influence of different hole-transporting materials on the performance of perovskite solar cells. Optik. 2017;130:650–658. doi: 10.1016/j.ijleo.2016.10.122. [DOI] [Google Scholar]
- Raj A., Kumar M., Bherwani H., Gupta A., Anshul A.. Evidence of improved power conversion efficiency in lead-free CsGeI3 based perovskite solar cell heterostructure via scaps simulation. J. Vac. Sci. Technol., B. 2021;39:012401. doi: 10.1116/6.0000718. [DOI] [Google Scholar]
- Anwar F., Mahbub R., Satter S. S., Ullah S. M.. Effect of Different HTM Layers and Electrical Parameters on ZnO Nanorod-Based Lead-Free Perovskite Solar Cell for High-Efficiency Performance. Int. J. Photoenergy. 2017;2017:9846310. doi: 10.1155/2017/9846310. [DOI] [Google Scholar]
- Du H.-J., Wang W.-C., Zhu J.-Z.. Device simulation of lead-free CH3NH3SnI3 perovskite solar cells with high efficiency. Chin. Phys. B. 2016;25:108802. doi: 10.1088/1674-1056/25/10/108802. [DOI] [Google Scholar]
- Abdelaziz S., Zekry A., Shaker A., Abouelatta M.. Investigating the performance of formamidinium tin-based perovskite solar cell by SCAPS device simulation. Opt. Mater. 2020;101:109738. doi: 10.1016/j.optmat.2020.109738. [DOI] [Google Scholar]
- Rolland, A. ; Pedesseau, L. ; Beck, A. ; Kepenekian, M. ; Katan, C. ; Huang, Y. ; Wang, S. ; Cornet, C. ; Durand, O. ; Even, J. . Computational design of high performance hybrid perovskite on silicon tandem solar cells. arXiv 2015. [Google Scholar]
- Alam I., Mollick R., Ashraf M. A.. Numerical simulation of Cs2AgBiBr6-based perovskite solar cell with ZnO nanorod and P3HT as the charge transport layers. Phys. B Condens. Matter. 2021;618:413187. doi: 10.1016/j.physb.2021.413187. [DOI] [Google Scholar]
- Chakraborty K., Choudhury M. G., Paul S.. Numerical study of Cs2TiX6 (X = Br–, I–, F– and Cl−) based perovskite solar cell using SCAPS-1D device simulation. Sol. Energy. 2019;194:886–892. doi: 10.1016/j.solener.2019.11.005. [DOI] [Google Scholar]
- Lakhdar N., Hima A.. Electron transport material effect on performance of perovskite solar cells based on CH3NH3GeI3. Opt. Mater. 2020;99:109517. doi: 10.1016/j.optmat.2019.109517. [DOI] [Google Scholar]
- Yue M., Su J., Zhao P., Lin Z., Zhang J., Chang J., Hao Y.. Optimizing the Performance of CsPbI3-Based Perovskite Solar Cells via Doping a ZnO Electron Transport Layer Coupled with Interface Engineering. Nano-Micro Lett. 2019;11:91. doi: 10.1007/s40820-019-0320-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hossain M., Alharbi F., Tabet N.. Copper Oxide as Inorganic Hole Transport Material for Lead Halide Perovskite Based Solar Cells of Enhanced Performance. Sol. Energy. 2015;120:370–380. doi: 10.1016/j.solener.2015.07.040. [DOI] [Google Scholar]
- Lenka T. R., Soibam A. C., Dey K., Maung T., Lin F.. Numerical analysis of high-efficiency lead-free perovskite solar cell with NiO as hole transport material and PCBM as electron transport material. CSI Trans. ICT. 2020;8:111–116. doi: 10.1007/s40012-020-00291-7. [DOI] [Google Scholar]
- Zhao P., Liu Z., Lin Z., Chen D., Su J., Zhang C., Zhang J., Chang J., Hao Y.. Device simulation of inverted CH3NH3PbI3–xClx perovskite solar cells based on PCBM electron transport layer and NiO hole transport layer. Sol. Energy. 2018;169:11–18. doi: 10.1016/j.solener.2018.04.027. [DOI] [Google Scholar]
- Isoe W., Mageto M., Maghanga C. M., Mwamburi M., Odari B., Awino C.. 2020 Thickness Dependence of Window Layer on CH 3 NH 3 PbI 3-X Cl X Perovskite Solar Cell. Int. J. Photoenergy. 2020;2020:1–7. doi: 10.1155/2020/8877744. [DOI] [Google Scholar]
- Deepthi Jayan K., Sebastian V.. Comprehensive device modelling and performance analysis of MASnI3 based perovskite solar cells with diverse ETM, HTM and back metal contacts. Sol. Energy. 2021;217:40–48. doi: 10.1016/j.solener.2021.01.058. [DOI] [Google Scholar]
- Anon Sigma-Aldrich. [Google Scholar]
- Uddin, R. ; Bhowmik, S. ; Ali, M. E. ; Alam, S. U. . Hole Transport Layer Free Non-toxic Perovskite Solar Cell Using ZnSe Electron Transport Material. In Machine Intelligence and Emerging Technologies; Satu, M. S. ; Moni, M. A. ; Kaiser, M. S. ; Arefin, M. S. . Springer Nature Switzerland: Cham, 2023, pp 486–498. [Google Scholar]
- Zhao P., Han M., Yin W., Zhao X., Kim S. G., Yan Y., Kim M., Song Y. J., Park N. G., Jung H. S.. Insulated Interlayer for Efficient and Photostable Electron-Transport-Layer-Free Perovskite Solar Cells. ACS Appl. Mater. Interfaces. 2018;10:10132–10140. doi: 10.1021/acsami.8b00021. [DOI] [PubMed] [Google Scholar]
- Burgelman M., Verschraegen J., Degrave S., Nollet P.. Modeling thin-film PV devices. Prog. Photovolt. Res. Appl. 2004;12:143–153. doi: 10.1002/pip.524. [DOI] [Google Scholar]
- Alam I., Ashraf M. A.. Effect of different device parameters on tin-based perovskite solar cell coupled with In 2 S 3 electron transport layer and CuSCN and Spiro-OMeTAD alternative hole transport layers for high-efficiency performance. Energy Sources, Part A. 2024;46:17080–17096. doi: 10.1080/15567036.2020.1820628. [DOI] [Google Scholar]
- Alipour H., Ghadimi A.. Optimization of lead-free perovskite solar cells in normal-structure with WO3 and water-free PEDOT: PSS composite for hole transport layer by SCAPS-1D simulation. Opt. Mater. 2021;120:111432. doi: 10.1016/j.optmat.2021.111432. [DOI] [Google Scholar]
- Azri F., Meftah A., Sengouga N., Meftah A.. Electron and hole transport layers optimization by numerical simulation of a perovskite solar cell. Sol. Energy. 2019;181:372. doi: 10.1016/j.solener.2019.02.017. [DOI] [Google Scholar]
- Bansal, S ; Aryal, P . Evaluation of new materials for electron and hole transport layers in perovskite-based solar cells through SCAPS-1D simulations. In 2016 IEEE 43rd Photovoltaic Specialists Conference (PVSC); IEEE: 2016, 0747-0750. [Google Scholar]
- Islam M. T., Jani M. R., Shorowordi K. M., Hoque Z., Gokcek A. M., Vattipally V., Nishat S. S., Ahmed S.. Numerical simulation studies of Cs3Bi2I9 perovskite solar device with optimal selection of electron and hole transport layers. Optik. 2021;231:166417. doi: 10.1016/j.ijleo.2021.166417. [DOI] [Google Scholar]
- Mandadapu U., Vedanayakam S. V., Thyagarajan K., Babu B. J.. Optimisation of high efficiency tin halide perovskite solar cells using SCAPS-1D. Int. J. Simul. Process Model. 2018;13:221. doi: 10.1504/IJSPM.2018.093097. [DOI] [Google Scholar]
- Gupta U. C., Gupta S. C.. Trace element toxicity relationships to crop production and livestock and human health: implications for management. Commun. Soil Sci. Plant Anal. 1998;29:1491–1522. doi: 10.1080/00103629809370045. [DOI] [Google Scholar]
- Påhlsson A-M B. Toxicity of heavy metals (Zn, Cu, Cd, Pb) to vascular plants Water. Air. Soil Pollut. 1989;47:287–319. doi: 10.1007/BF00279329. [DOI] [Google Scholar]
- Khoshsirat N., Md Yunus N. A., Hamidon M. N., Shafie S., Amin N.. Analysis of absorber layer properties effect on CIGS solar cell performance using SCAPS. Optik. 2015;126:681–686. doi: 10.1016/j.ijleo.2015.02.037. [DOI] [Google Scholar]
- Kim G.-H., Kim D. S.. Development of perovskite solar cells with> 25% conversion efficiency. Joule. 2021;5:1033–1035. doi: 10.1016/j.joule.2021.04.008. [DOI] [Google Scholar]
- Boyd C. C., Cheacharoen R., Leijtens T., McGehee M. D.. Understanding degradation mechanisms and improving stability of perovskite photovoltaics. Chem. Rev. 2019;119:3418–3451. doi: 10.1021/acs.chemrev.8b00336. [DOI] [PubMed] [Google Scholar]
- Mesquita I., Andrade L., Mendes A.. Temperature impact on perovskite solar cells under operation. ChemSusChem. 2019;12:2186–2194. doi: 10.1002/cssc.201802899. [DOI] [PubMed] [Google Scholar]
- Stranks S. D., Snaith H. J.. Metal-halide perovskites for photovoltaic and light-emitting devices. Nat. Nanotechnol. 2015;10:391–402. doi: 10.1038/nnano.2015.90. [DOI] [PubMed] [Google Scholar]
- Kumar A., Bansode U., Ogale S., Rahman A.. Understanding the thermal degradation mechanism of perovskite solar cells via dielectric and noise measurements. Nanotechnology. 2020;31:365403. doi: 10.1088/1361-6528/ab97d4. [DOI] [PubMed] [Google Scholar]
- Ferdous Utsho K. I., Mostafa S. M. G., Tarekuzzaman Md, Al-Saleem M. S. M., Nahid N. I., Al-Humaidi J. Y., Rasheduzzaman Md, Rahman M. M., Hasan Md Z. Optimizing Cs2CuBiBr6 double halide perovskite for solar applications: the role of electron transport layers in SCAPS-1D simulations. RSC Adv. 2025;15:2184–2204. doi: 10.1039/D4RA08515A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang S., Sakurai T., Wen W., Qi Y.. Energy level alignment at interfaces in metal halide perovskite solar cells. Adv. Mater. Interfaces. 2018;5:1800260. doi: 10.1002/admi.201800260. [DOI] [Google Scholar]
- Hussain B., Aslam A., Khan T. M., Creighton M., Zohuri B.. Electron affinity and bandgap optimization of zinc oxide for improved performance of ZnO/Si heterojunction solar cell using PC1D simulations. Electronics. 2019;8:238. doi: 10.3390/electronics8020238. [DOI] [Google Scholar]
- Biswas S. K., Sumon M. S., Sarker K., Orthe M. F., Ahmed M. M.. A Numerical Approach to Analysis of an Environment-Friendly Sn-Based Perovskite Solar Cell with SnO2 Buffer Layer Using SCAPS-1D. Adv. Mater. Sci. Eng. 2023;2023:4154962. doi: 10.1155/2023/4154962. [DOI] [Google Scholar]
- Wang S., Peng Y., Li L., Zhou Z., Liu Z., Zhou S., Yao M.. Impact of loss mechanisms on performances of perovskite solar cells. Phys. B Condens. Matter. 2022;647:414363. doi: 10.1016/j.physb.2022.414363. [DOI] [Google Scholar]
- Green M. A., Bremner S. P.. Energy conversion approaches and materials for high-efficiency photovoltaics. Nat. Mater. 2017;16:23–34. doi: 10.1038/nmat4676. [DOI] [PubMed] [Google Scholar]
- Da Y., Xuan Y., Li Q.. Quantifying energy losses in planar perovskite solar cells. Sol. Energy Mater. Sol. Cells. 2018;174:206–213. doi: 10.1016/j.solmat.2017.09.002. [DOI] [Google Scholar]
- Servaites J. D., Yeganeh S., Marks T. J., Ratner M. A.. Efficiency enhancement in organic photovoltaic cells: consequences of optimizing series resistance. Adv. Funct. Mater. 2010;20:97–104. doi: 10.1002/adfm.200901107. [DOI] [Google Scholar]
- Araujo G. L., Cuevas A., Ruiz J. M.. The effect of distributed series resistance on the dark and illuminated currentVoltage characteristics of solar cells. IEEE Trans. Electron Devices. 1986;33:391–401. doi: 10.1109/T-ED.1986.22500. [DOI] [Google Scholar]
- Rühle S.. Tabulated values of the Shockley-Queisser limit for single junction solar cells Sol. Energy. 2016;130:139–147. doi: 10.1016/j.solener.2016.02.015. [DOI] [Google Scholar]
- Raoui Y., Ez-Zahraouy H., Tahiri N., El Bounagui O., Ahmad S., Kazim S.. Performance analysis of MAPbI3 based perovskite solar cells employing diverse charge selective contacts: Simulation study. Sol. Energy. 2019;193:948–955. doi: 10.1016/j.solener.2019.10.009. [DOI] [Google Scholar]
- Singh A. K., Srivastava S., Mahapatra A., Baral J. K., Pradhan B.. Performance optimization of lead free-MASnI3 based solar cell with 27% efficiency by numerical simulation. Opt. Mater. 2021;117:111193. doi: 10.1016/j.optmat.2021.111193. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The parameter data used to support the findings of this study are included in the article.















