DOI:
10.1039/D5RA05879A
(Paper)
RSC Adv., 2025,
15, 35859-35882
Electron extraction layer-driven performance enhancement in CaHfSe3 photovoltaics
Received
10th August 2025
, Accepted 23rd September 2025
First published on 29th September 2025
Abstract
Traditional solar cells – including those based on silicon or lead-halide perovskites – have a number of significant disadvantages, including long-term instability, costs, and toxicity. We demonstrate the suitability of CaHfSe3 as a promising next-generation lead-free thermally stable absorber material. We performed a comprehensive numerical simulation study using SCAPS-1D to consider several device topologies of the type FTO/TiO2 AZnO, WS2/CaHfSe3/MoO3/Au, and to investigate the different features of a system based on CaHfSe3. We conducted a full parametric study of the impacts of absorber thickness, defect density, acceptor doping and concentration, as well as carrier concentrations in the electron and hole transport layers. In addition, through experimentation we considered the operational characteristics of carrier generation-recombination methods, temperature and back contact effect current–voltage I–V characteristics, quantum efficiency, and the influence of series and shunt resistance. This allowed us to determine the optimized configuration. The top-performing structure, FTO/TiO2/CaHfSe3/MoO3/Au, had an outstanding PCE of 32.39%, VOC = 1.52 V, JSC = 23.17 mA cm−2, and FF = 91.41%. This research offers both fundamental insights and practical guidance for developing stable, efficient, and environmentally friendly CaHfSe3-based solar cells. It paves the path for further experimental realization and commercial application.
1. Introduction
Transition to renewable energy sources has emerged as a global imperative to address the challenges posed by climate change and rising electricity demand. Among energy conversion technologies, solar cells have experienced significant growth thanks to advances in developing new semiconductor materials.1,2 Although silicon-based photovoltaic cells currently dominate the market, they have certain limitations regarding manufacturing cost, efficiency, and stability. In particular, crystalline silicon suffers from intrinsic constraints as an indirect band gap semiconductor, its low absorption coefficient requires a minimum thickness to ensure sufficient light harvesting. Still, increasing thickness inevitably raises the series resistance.3 This trade-off between absorption and Joule losses fundamentally restricts the overall performance of silicon solar cells. These constraints have driven the exploration of alternative materials, especially perovskites, which are garnering increasing attention in this context due to their exceptional optical and electronic properties, cost-effectiveness, and sustainability for photovoltaic devices.4–6 Specifically, MAPbX3 or FABX3 lead halide perovskites have demonstrated remarkable power conversion efficiency (PCE) of 27%.7 However, conventional PSCs are weak in resilience to heat and humidity, have a high concentration of defects, are unstable due to the instability of organic cations, and contain poisonous lead (Pb), all of which pose significant health and environmental dangers, preventing their widespread commercialization. The primary obstacle is the search for lead (Pb)-free, stable, and eco-friendly materials. This is achieved by replacing cations such as Ge2+ or Sn2+, which have advantageous band gaps, for Pb2+ ions.8,9 However, the resultant perovskites are extremely unstable in air because of the strong oxidation propensity of Ge2+ and Sn2+ to Sn4+ and Ge4+, respectively, which is linked to the increased energy levels of their respective 5s and 4s orbitals.10 However, a major obstacle still stands in the way of these promising solar materials, poor durability.
This study is aimed at identifying materials with the positive attributes of lead halide perovskites while eliminating their disadvantages. Chalcogenide perovskites are one of the most promising alternatives due to their non-toxic nature, impressive structural stability, and good electronic properties. These compounds typically adopt the ABX3 structural formula, where A is an alkaline earth metal, B is a transition metal, and X is either sulfur (S) or selenium (Se).11 A broad range of ABX3 chalcogenide perovskites was recognized by Sun et al. in 2015 as potential absorber materials for solar applications. BaZrS3, CaZrS3, SrZrS3, BaHfS3, CaHfS3, SrHfS3, CaTiS3, BaZrSe3, CaZrSe3, SrZrSe3, BaHfSe3, SrHfSe3, and CaTiSe3 are some of the more prominent instances.12 Simulations highlight the strong potential of ABX3-based solar cells, but experimental realization is limited by the complexity of film growth and ambient stability issues, as defects and exposure to moisture or oxygen can reduce efficiency. Overcoming these challenges is essential to achieve the predicted performance. Crystalline chalcogenide perovskite (CP) thin films are typically synthesized at high temperatures, greater than 600 °C (1170 F), severely limiting the ability to fabricate both single and multi-junction solar cells. The production of high-quality CP films via physical deposition methods or chemical sulfidation processes typically involves temperatures as high as 700 to 1000 °C (1292 to 1832 F), where intermediate products frequently include impurity phases in the final composition.13 Additionally, fabricating CP thin films via solution-based methods remains a significant challenge, as no known solvent can withstand the extreme temperatures needed for their synthesis. Furthermore, in 2025, Hayes et al. devised a comprehensive method for synthesizing various ABS3 compounds, such as BaZrS3, BaTiS3, SrZrS3, and SrHfS3, together with colloidal phases synthesized within the Ba–Hf–S chemical system. These materials are composed of elements with a strong affinity for oxygen, notably group IV transition metals like Ti, Zr, and Hf.14 This study identifies deficiencies in current research, noting that while BaZrS3 nanoparticles have been manufactured in various phases by colloidal techniques, a consistent procedure for orthorhombic perovskite nanoparticles has not been established. To solve this, the authors introduce a high-temperature hot-injection technique, which allows precise control over the formation of colloidal BaZrS3 nanoparticles, leading to a more reliable and scalable synthesis process. Recent studies on the ABS3 chalcogenide perovskite show that replacing S with Se dramatically lowers the bandgap moving it away from the visible to the near-infrared wavelength. This transformation improves the capability for photovoltaic implementation and appear more favorable for transitioning solar energy technology.15 Hf based perovskites still suffer from bulk recombination mainly due to Hf vacancy defects.16 The lead-free perovskite CaHfSe3 emerges as a viable alternative, as it combines the advantageous characteristics of semiconductors with properties comparable to its lead-based counterparts. These include favorable charge-carrier effective masses, a direct and optimal band gap, high absorption coefficients, and exceptional thermal stability, making it a promising absorber material for next-generation photovoltaics. The optical spectrum of the CaHfSe3 reports a clear absorption edge which corresponds to a direct electronic bandgap of 1.65 eV.15 This optimization study of CaHfSe3-based perovskite solar cells (PSCs) focuses on various combinations of electron and hole transport layers (ETL–HTL), using SCAPS-1D simulations, a one-dimensional solar cell modelling software based on capacitance. To enhance device performance, MoO3 is employed as the HTL in combination with TiO2, WS2, and AZnO as ETL candidates. Key operational parameters such as back contact material, series resistance, and operating temperature are systematically examined. Additionally, the effects of absorber layer thickness and doping concentration are optimized to further improve cell efficiency. Analysis of current–voltage (J–V) characteristics and external quantum efficiency (EQE) confirms that the optimized configuration yields a significantly enhanced power conversion efficiency (PCE) compared to previously reported designs.
2. Device architecture and computational methodology
2.1. Numerical modeling technique
The numerical method employed in this research is based on simulations conducted with SCAPS-1D 3.3.11 (Solar Cell Capacitance Simulator – 1D), a widely used software for modeling the electrical behavior of thin-film solar cells, which solves the coupled Poisson's and continuity equations under steady-state conditions. As shown in eqn (1) Poisson equation, relates electrostatic potential to charge density in the material within the device:17 |
 | (1) |
With ε: dielectric constant, q: electron charge, ψ: electrostatic potential, n, and p: free electrons and holes, nt and np: trapped electrons and holes, ND+and NA−: represent the densities of ionized donors and acceptors, respectively.
Poisson's equation must be solved in order to analyze solar cells since it shows how the electric potential is dispersed throughout the device's many layers. This knowledge facilitates the analysis of charge carrier dynamics, which involves the movement and recombination of electrons and holes to generate energy. The continuity equations reflect the principle of charge conservation and describe the motion of charge carriers in semiconductors. In this case, eqn (2) and (3) represent continuity equations for electrons and holes, respectively.
|
 | (2) |
|
 | (3) |
where
G is the photogeneration rate,
Rn and
Rp are the rates of recombination of electrons and holes, respectively and
Jn and
Jp being the electrical current densities of electrons and holes. In semiconductor materials, electrons or holes can be migrating for two reasons—there is an electric field that produces a drift current or there is a concentration gradient of carriers so that there is a diffusion current.
Eqn (4) and
(5) describe the electrical current densities associated with electron and hole transport.
|
 | (4) |
|
 | (5) |
Using the optical absorption model, the optical absorption constant α is computed, as shown in eqn (6).
|
 | (6) |
where
υ is the radiation frequency,
Eg is the material's real band gap,
h is Planck's constant, and
A and
B are the model parameters.
2.2. Simulation settings and device structure
In the current work, SCAPS-1D Software was used to model a device with the structure of (back metals/MoO3/CaHfSe3/ETMs/FTO), as shown generally in Fig. 1, to simulate the one-dimensional performance of a thin films photovoltaic device. The ETMs WS2, AZnO, and TiO2 were used due to their properties that complement the structural and electronic fabric of the ETLs for improved extraction of electrons through the layers to the collecting circuits of solar cell architectures. The two different ETLs are used along with MoO3 as a hole transport layer or HTL layer and the studies explored the pairs of combinations that produced the best configuration for the device. The materials used for the CaHfSe3 absorber layer, the FTO-coated glasses, which are transparent conducting materials, and the back contact varied with Cu, Fe, Au, C, Ni, and Pt. The project is focused on improving the metrics and performance of CaHfSe3-based lead-free perovskite solar cells by simulating modifications to the HTLs (WS2, AZnO, TiO2), and ETL(MoO3), and the back contacts (Cu, Fe, Au, C, Ni, Pt). For each layer, key parameters were defined, such as thickness, bandgap energy, electron affinity, permittivity, doping concentration, carrier mobilities, and defect densities, are listed in Table 1.
 |
| Fig. 1 The device architecture of the PSC based on CaHfSe3. | |
Table 1 The key input values defining the architecture of the proposed solar cell
Parameters |
FTO18 |
TiO2 (ref. 16) |
CaHfSe3 (ref. 19) |
WS2 (ref. 20) |
MoO3 (ref. 21) |
AZnO8 |
Thickness (μm) |
0.1 |
0.03 |
Varied |
0.03 |
0.03 |
0.03 |
Bandgap Eg (eV) |
3.5 |
3.33 |
1.65 |
1.8 |
3.17 |
3.33 |
Electron affinity χ (eV) |
4 |
4.55 |
4 |
3.95 |
2.05 |
4.55 |
Dielectric permittivity (relative) |
9 |
8.12 |
7.93 |
13.6 |
12.5 |
8.12 |
Nc effective density of states in CB (cm−3) |
2.2 × 1018 |
4.1 × 1018 |
8.9 × 1018 |
1 × 1018 |
2.21 × 1019 |
4.1 × 1018 |
Nv effective density of states in VB (cm−3) |
1.8 × 1019 |
8.2 × 1019 |
1.1 × 1019 |
2.4 × 1019 |
1.8 × 1019 |
8.2 × 1019 |
Vth,e velocity of electrons (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
Vth,p velocity of holes (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
μn mobility of electrons (cm2 V−1 s−1) |
20 |
100 |
18.77 |
100 |
25 |
100 |
μp mobility of holes (cm2 V−1 s−1) |
10 |
20 |
6.81 |
100 |
100 |
20 |
ND shallow uniform donor density (cm−3) |
21 × 1019 |
11 × 1013 |
0 |
1 × 1018 |
0 |
11 × 1013 |
NA shallow uniform acceptor density (cm−3) |
0 |
0 |
1 × 1019 |
0 |
1 × 1019 |
0 |
Nt density of defect (cm−3) |
1 × 1015 |
1 × 1017 |
1 × 1013 |
1 × 1015 |
1 × 1015 |
1 × 1017 |
Table 2 presents the work functions of various materials, including Gold (Au), Nickel (Ni), Carbon (C), Iron (Fe), Copper (Cu), and Platinum (Pt). The simulations were conducted under standard AM1.5G illumination (1000 W m−2) at an operating temperature of 300 K. Additionally, series/shunt resistances and working temperature were incorporated to reflect realistic electrical behavior. Device performance was assessed using J–V characteristics, quantum efficiency (QE), and power conversion efficiency (PCE), with systematic variation of key parameters to determine the optimal configuration.
Table 2 The contact work functions applied in the configurations
Metal contacts |
Work function (eV) (ref. 8 and 18) |
Cu |
4.65 |
Fe |
4.81 |
C |
5 |
Au |
5.1 |
Ni |
5.5 |
Pt |
5.7 |
3. Outcome analysis and interpretation
3.1. The band diagram of the model cell
Simulated band diagrams for the proposed structures are shown in Fig. 2, highlighting the importance of energy level alignment in achieving efficient charge separation and high photovoltaic performance. To ensure effective electron transport, the electron affinity of the electron transport layer (ETLs) must exceed that of the CaHfSe3 absorber, facilitating downward conduction band alignment.22 Conversely, hole extraction is enhanced when the ionization potential of the hole transport layer (MoO3) is lower than that of the absorber, ensuring a smooth flow of holes.23
 |
| Fig. 2 CaHfSe3-based PSC energy diagram with MoO3 as the HTL and (a) AZnO, (b) TiO2, and (c) WS2 as the ETL. | |
The WS2/CaHfSe3 interface exhibits a small CBO (0.02 eV), indicating good energy level continuity and promoting efficient electron transport. In the case of TiO2, which has a wide Eg of 3.33 eV, the conduction band alignment (0.01 eV) also supports efficient electron extraction while acting as a barrier to hole backflow. The AZnO/CaHfSe3 interface shows a conduction band offset of approximately 0.03 eV, reflecting strong energetic compatibility and potential for high electron collection efficiency. On the hole transport side, the CaHfSe3/MoO3 interface demonstrates a favorable valence band offset (VBO), as the valence band maximum (VBM) of MoO3 lies below that of CaHfSe3. This ensures selective and efficient hole extraction while suppressing electron recombination at the back contact. Altogether, these band alignments confirm the suitability of WS2, TiO2, and AZnO as ETLs, and MoO3 as an HTL, in facilitating efficient charge separation and transport in CaHfSe3-based photovoltaic devices.
3.2. Coefficient of absorption
The coefficient of absorption (α) as a function of wavelength indicates a material's ability to absorb light across various regions of the electromagnetic spectrum (UV, visible, and infrared). Fig. 3 displays the theoretical absorption profiles for all layers. The absorbent material CaHfSe3 is notable for its strong Capability of absorbing radiation in the visible region, with values exceeding 105 cm−1 between 400 and 770 nm. This characteristic makes it an excellent choice for harnessing solar light and producing charge carriers. TiO2 and AZnO, used as electron transport layers (ETL), exhibit very low absorption coefficients (less than 102 cm−1 between 350 nm and 700 nm), allowing light to pass through with minimal absorption and facilitating optimal transmission to the active CaHfSe3 layer. Finally, MoO3, used as a hole transport layer (HTL), demonstrates intermediate behavior with a low absorption coefficient (around 103) in the 350–700 nm range, minimizing interaction with incident light. Its absorption increases slightly in the UV range but remains compatible with its role.24
 |
| Fig. 3 Absorption coefficients of CaHfSe3, TiO2, AZnO, WS2, and MoO3 as a function of wavelength. | |
3.3. Impact of ETL and absorber thickness on device performance
To improve light harvesting and promote efficient extraction of charge carriers, the absorption and ETL thicknesses must be carefully chosen and optimized. Maximum photon absorption in the active layer and efficient electron transport are ensured by a well-optimized design. In this study, contour plots were used to investigate the impact of absorber and ETL thickness variations on the performance of CaHfSe3 architectural performance has been investigated. Through this research, we can investigate how altering the thickness of the absorber and electron transport layers affects key performance metrics, including VOC, FF, JSC, and PCE. To identify the optimal thickness combinations for enhanced device performance, the contour plots display the predicted values of JSC, VOC, FF, and PCE across a range of absorber thicknesses (300 to 2100 nm) and ETL thicknesses (from 30 to 210 nm). In this investigation section, MoO3 was held constant as the HTM. Fig. 4 shows the impact of absorber layer thickness and ETL on the VOC. Under conditions where the absorber and ETL layers were about 300 nm and <150 nm in thickness, respectively, the perovskite employing TiO2 and AZnO as ETLs reached a peak VOC of 1.54 V, as illustrated in Fig. 4(a and b). As the thicknesses of both the absorber and ETL layers grew, the open-circuit voltage (VOC) exhibited a decline. This decrease can be attributed to a rise in the reverse saturation current caused by the thicker absorber, combined with the partial absorption of incident light within the thicker ETL layers. By contrast, when absorber and ETL thicknesses were approximately 1500 nm and ≤30 nm, respectively, the PSC with a WS2 ETL showed a maximum VOC of 1.34 V.
 |
| Fig. 4 VOC contour mapping for (a) AZnO, (b) TiO2, and (c) WS2 ETLs with CaHfSe3 absorber thickness. | |
The influence of changes in the absorber layer thickness and ETL on the JSC parameter of three selected perovskite solar cells is illustrated in Fig. 5. The PSC incorporating TiO2 as the ETL achieved a maximum current density of 23.33 mA cm−2 once the absorber layer exceeded 1800 nm and the ETL thickness was ≥30 nm. Under conditions where the absorber measured 2100 nm and the thickness of ETL ranged from 30 to 210 nm, the ETL AZnO demonstrated the greatest JSC of 23.4 mA cm−2. In addition, the PSCs using WS2 ETL exhibited maximum (JSC) of 23.4 mA cm−2 when the absorber thickness was 2100 nm and the ETL thickness ranged from 30 to 90 nm. Generally, an increase in absorber thickness led to higher JSC values in each device, as a thicker absorber enhances photon absorption, promotes greater electron–hole pair generation, and consequently boosts the photocurrent. This enhancement is also ascribed to the improved spectral responsiveness at extended wavelengths. Fig. 6 illustrates how different absorber and ETL layer thicknesses affect FF values. When WS2 is implemented as the electron transport layer (ETL) in solar cell structures, studies have revealed that the FF of BaHfSe3-based solar cells increase with absorber thickness up to 1.8–2.1 μm, while it slightly decreases as the WS2 layer becomes thicker. The highest FF value of 84.86% is recorded when the BaHfSe3 thickness ranges from 1.8 to 2.1 μm and the WS2 thickness is between 0.03 and 0.06 μm, Fig. 6(b).
 |
| Fig. 5 JSC contour mapping for (a) AZnO, (b) TiO2, and (c) WS2 electron transport layers (ETLs) as a function of CaHfSe3 absorber thickness. | |
 |
| Fig. 6 Contour mapping of the fill factor (FF %) for different ETL materials: (a) AZnO, (b) TiO2, and (c) WS2. | |
Using TiO2 as an ETL results in a maximum fill factor (FF) of 91.2% with a BaHfSe3 thickness of 0.9 μm and a TiO2 layer thickness ranging from 0.06 to 0.09 μm. In contrast, increasing the thickness of BaHfSe3 beyond 1.2 μm leads to a slight decline in FF, likely resulting from enhanced recombination losses or reduced efficiency in charge transport. Furthermore, Fig. 6(a) illustrates that the AZnO electron transport layer thickness significantly impacts the fill factor (FF), with ultrathin layers (≤0.03 μm) showing optimal values (∼90.7%). FF gradually decreases as AZnO thickness increases, primarily due to heightened recombination losses and series resistance. Conversely, variations in the BaHfSe3 absorber's thickness have minimal effect on the FF, indicating consistent device performance over a wide thickness range (0.3–2.1 μm), as charge carrier extraction and production remain unchanged. Fig. 7 displays the simulation findings showing distinct performance characteristics for the three ETLs when incorporated into BaHfSe3-based solar cells. AZnO-based devices reach a top PCE of 31.05% when configured with a 1.8 μm absorber and a 30 nm ETL, showcasing a favorable trade-off between efficiency and material economy. Notably, AZnO maintains good efficiency at reduced absorber thicknesses, making it a low-cost and easy-to-fabricate alternative. However, the devices with TiO2 exhibit competitive performance, achieving a slightly higher maximum PCE of 31.08% with a thicker absorber layer of 2.1 μm. Its reliance on increased thickness implies a greater dependence on photon absorption to compensate for its reduced electron mobility. Conversely, WS2 shows a lower overall efficiency, with a maximum PCE of 26.67% occurring at WS2 thicknesses less than 0.15 μm combined with absorber thicknesses ranging from 1.8 to 2.1 μm.
 |
| Fig. 7 Contour mapping of PCE (%) when ETL is (a) AZnO, (b) TiO2, and (c) WS2. | |
There are several reasons for the noticeable differences in VOC, JSC, FF, and PCE across perovskite with various ETL. The absorption coefficient is a crucial component that greatly impacts photovoltaic metrics, including VOC, JSC, FF, and PCE. This parameter affects the coupling efficiency of light photons with the underlying CaHfSe3 absorber layer and is directly correlated with the ETL's bandgap. The overall efficiency of the solar cell is significantly affected by the CBO, arising from the disparity in electron affinity and Fermi level positions between the absorber and the ETL.
3.4. Effect of absorber acceptor concentration and layer thickness on photovoltaic performance
Optimizing both the absorber thickness and NA within the absorber is essential for achieving improved efficiency. To examine the impact of these factors, we conducted simulations by adjusting the acceptor doping density (NA) from 1 × 1015 to 1 × 1020 cm−3 and absorber thickness from 0.3 to 1.5 μm across seven distinct PSC configurations. Fig. 8 shows how variations in absorber thickness and NA influence the (VOC) in the validated PSC configurations.
 |
| Fig. 8 Contour plots illustrate the influence of absorber doping concentration and thickness on VOC for ETL types (a) AZnO, (b) TiO2, and (c) WS2. | |
According to Fig. 8, under conditions where the absorber was less than 0.9 μm thick and the acceptor doping concentration (NA) was fixed at 1020 cm−3, AZnO and TiO2 achieved their maximum VOC values of 1.5 V. Notably, increasing the absorber thickness beyond this threshold resulted in a gradual decline in VOC for all ETLs under the same doping conditions. The data's shape makes it evident that while VOC tends to drop with increasing absorber thickness, it increases with increased acceptor concentration (NA). This behavior is explained by the stronger electric field and greater built-in potential at higher doping levels, which makes it easier to separate and remove photogenerated carriers and increase VOC. Nevertheless, the series resistance rises, and the greater sheet resistance at thicker absorber layers constrains the hole mobility towards the HTL. This leads to a decrease in carrier collection and, thus, a lower VOC.25
Fig. 9(a)–(c) show that the short-circuit current density JSC for devices using ZnO, TiO2, and WS2 as ETL shows a significant increase with absorber thickness and a modest increase with increasing acceptor density (NA). The maximum JSC values of 23.239 mA cm−2, 23.169 mA cm−2 and 23.232 mA cm−2 for ZnO, TiO2, and WS2, respectively, were observed at an absorber thickness of 1.5 μm with NA = 1020 cm−3. The increased light absorption in the thicker CaHfSe3 absorber layer, which permits a bigger production of photogenerated carriers, is primarily responsible for the progressive increase in JSC seen for all ETLs from 0.3 μm to 1.5 μm. Furthermore, for a constant thickness, the JSC values are rather stable over NA values and improve slightly with increasing NA. This is explained by the improved separation of photogenerated carriers in highly doped materials, leading to more effective electric field generation and less recombination. Fig. 10(a)–(c) illustrate that, for all three ETLs, the fill factor (FF) typically increases with the rising acceptor density (NA) throughout all examined thicknesses.
 |
| Fig. 9 Contour plots illustrate the influence of absorber doping concentration and thickness on JSC for ETL types (a) AZnO, (b) TiO2, and (c) WS2. | |
 |
| Fig. 10 Impact of (a) AZnO, (b) TiO2, and (c) WS2 ETLs on FF (%) contour plots in CaHfSe3 perovskite solar cells. | |
This enhancement results from minimized carrier recombination and enhanced charge transport as a result of increased doping levels. TiO2 often demonstrates the highest fill factor values, achieving 91.48% at a minimal thickness of 0.3 μm and elevated doping levels (NA > 1019 cm−3). This exceptional performance is attributed to its effective surface passivation characteristics and favorable band alignment with CaHfSe3, which reduces non-radiative recombination. At low acceptor densities, AZnO exhibits lower FF values than TiO2, but, when NA rises, it shows notable improvement, reaching up to 91.03–91.06% for NA = 1020 cm−3. However, we observe that the WS2-based solar cell structure exhibits a maximum FF value, reaching nearly 90% when the thickness of absorber exceeds 0.8 μm, and NA ranges from 1017 to 1018 cm−3.
Fig. 11(a) and (b) highlight the combined effect of absorber thickness and NA on the performance of CaHfSe3-based solar cells. The simulations show that the PCE reaches its peak at absorber thicknesses of 1.2 μm and 1.5 μm, achieving 32.30% for AZnO and 32.40% for TiO2 when NA exceeds 1019 cm−3. A reduction in absorber thickness leads to a noticeable drop in PCE, since thinner layers allow a considerable fraction of incident photons to pass through unabsorbed, thereby limiting electron–hole pair generation. Conversely, thicker absorbers enhance light absorption, increase carrier generation, and result in higher efficiencies. In addition, the PCE improves with increasing NA, as higher acceptor doping concentrations strengthen the built-in electric field, which facilitates the separation and transport of photogenerated carriers, reduces recombination losses, and ultimately enhances device performance.26 Fig. 11(c) shows that using WS2 as the ETL results in a PCE of 29.12% when the absorber thickness exceeds 1.2 μm and the NA is approximately 1018 cm−3. This phenomenon occurs because a high acceptor density increases the recombination rate, restricting carrier mobility and subsequently reducing the PCE.
 |
| Fig. 11 PCE (%) contour plots with ETL as (a)AZnO, (b) TiO2, and (c) WS2. | |
3.5. Influence of defect density and CaHfSe3 layer thickness on photovoltaic performance
Defect density is a crucial factor that significantly impacts perovskite solar cells (PSCs). Absorber layer defects, acting as recombination sites, reduce carrier lifetimes and decrease device efficiency. Consequently, achieving excellent performance necessitates reducing the absorber layer's defect density. Numerous studies have focused on the interplay between defect density and absorber thickness in CaHfSe3-based PSCs and their effect on device efficiency and the results demonstrate a significant relationship between the two factors. Optimizing both simultaneously can result in the most efficient device possible.
Hence, the thickness of the CaHfSe3 absorber ranged from 300 nm to 1500 nm, alongside Nt values adjusted from 1013 to 1015 cm−3. The simulation outcomes indicate that both the absorber thickness and Nt substantially affect the VOC of CaHfSe3-based solar cells, as shown in Fig. 12. For all evaluated ETLs (AZnO, TiO2, and WS2), In configurations (a) and (b), VOC exhibits a strong dependence on both thickness and defect density. When Nt is maintained below 1014 cm−3, VOC reaches values as high as 1.448 V, indicating minimal nonradiative recombination and efficient carrier extraction. The voltage increases with thickness up to approximately 0.9 μm, beyond which the gain plateaus, suggesting that additional thickness does not significantly enhance photogeneration or reduce recombination. This behavior reflects a balance between optical absorption and carrier diffusion length, where excessive thickness may introduce resistive losses or trap-limited transport without further improving voltage. As Nt increases beyond 1016 cm−3, a pronounced decline in VOC is observed, regardless of thickness. This trend highlights the detrimental impact of bulk and interfacial defects, which act as recombination centers and suppress quasi-Fermi level splitting. The steep voltage gradient across the defect density axis underscores the necessity of defect passivation strategies, such as surface treatments or compositional tuning, to preserve high VOC values. In contrast, configuration (c) yields significantly lower VOC values, ranging from 1.128 V to 1.148 V. The voltage remains relatively insensitive to thickness variations, and even at low Nt, the performance is markedly inferior to that of configurations (a) and (b). This suggests a fundamentally different device architecture or material interface, potentially characterized by poor band alignment, high interface recombination velocity, or inadequate carrier selectivity. The suppressed VOC may also reflect unfavorable energetics at the contact layers or the presence of deep-level traps within the CaHfS3 matrix. Particularly at low defect concentrations (Nt = 1013 cm−3), since the longer carrier transport pathways raise the possibility of recombination. Furthermore, the greatest VOC values reported are 1.504 V for both AZnO and TiO2 at 0.3 μm thickness and Nt = 1013 cm−3, and 1.342 V for WS2 at 1.2–1.5 μm thickness with the same low defect density.
 |
| Fig. 12 Contour graphs of VOC dependence on CaHfSe3 absorber thickness and defect density with ETLs as (a) AZnO, (b) TiO2, and (c) WS2. | |
Fig. 13 shows that the JSC of CaHfSe3-based solar cells is significantly influenced by both the absorber thickness and Nt. Because larger layers generate more carriers and absorb more photons, JSC grows continuously with absorber thickness. Beyond 1.2 μm, this gain tends to saturate, suggesting an ideal thickness range for balancing recombination and absorption losses. On the other hand, JSC slightly decreases when the defect density is increased from 1013 to 1015 cm−3 at a given thickness. This is because greater trap-assisted recombination lowers the carrier collection efficiency. The maximum JSC values recorded are 23.24, 23.02, and 23.24 mA cm−2 for AZnO, TiO2, and WS2, respectively. These values all correspond to a low defect density <1014 cm−3 and an absorber thickness greater than 1.2 μm. These results validate that adequate absorber thickness and defect passivation are both necessary to maximize photo-generated current.
 |
| Fig. 13 Contour graphs of JSC as a function of CaHfSe3 absorber thickness and defect density with ETL as (a) AZnO, (b) TiO2, and (c) WS2. | |
For the device utilizing AZnO as the electron transport layer, Fig. 14(a), the fill factor exhibited considerable stability across varying absorber thicknesses and defect concentrations. In particular, FF values were higher than 90.50% for thicknesses <900 nm and Nt < 1015 cm−3. The highest FF of 90.78% was observed at Nt = 1013 cm−3 and thickness ∼0.6 μm. At 1.2 μm thickness, FF decreased from 90.78% to 89.76% as Nt grew from 1013 to 1015 cm−3, indicating a modest drop with rising Nt. Due to accumulated recombination. Thus, low Nt and thickness allow for better fill factor, especially internal losses. Furthermore, the TiO2-based ETL device exhibited steady performance over diverse absorber thickness and Nt conditions, as presented in Fig. 14(b). A maximum fill factor (FF) of 91.20% was achieved at Nt = 1.0 × 1013 cm−3 and a thickness of 0.9 μm. As the defect density increased, the FF gradually declined; however, it remained above 90% under all tested conditions. This highlights the excellent compatibility of TiO2 with the CaHfSe3 absorber layer, as well as its effectiveness in minimizing series resistance and enabling efficient charge extraction.
 |
| Fig. 14 Contour graphs of FF as a function of CaHfSe3 absorber thickness and defect density with ETL as (a) AZnO, (b) TiO2, and (c) WS2. | |
In contrast, the device employing WS2 as the ETL displayed a lower but increasing FF trend with absorber thickness. Fig. 14(c) indicates that FF rose from 85.04% to 85.80% as thickness increased from 0.3 μm to 1.5 μm at Nt = 1.0 × 1015 cm−3, achieving a maximum FF of 85.80%. At low Nt values (<1014 cm−3), FF stayed below 85.00%, suggesting that WS2 is more suitable for thicker absorbers with higher defect tolerance. The lowest result, 84.74%, was recorded at 0.3 μm and Nt = 1.0 × 1013 cm−3.
Fig. 15 shows the PCE of CaHfSe3-based solar cells utilizing three different ETLs: AZnO, TiO2, and WS2, with various absorber thicknesses and defect densities. For all ETLs, the PCE increased with absorber thickness and dropped with increasing defect density (Nt), which corresponded to improved light absorption and lower carrier recombination at low Nt. The device achieved a maximum PCE of 31.00% when AZnO served as the ETL, combined with a carrier concentration of 1013 cm−3 and an absorber thickness of 1.5 μm. Moreover, Fig. 15(b) shows that the device using TiO2 as the ETL demonstrated high performance, achieving a maximum PCE of 30.81% under similar conditions. In contrast, Fig. 15(c) reveals that WS2 exhibited a lower peak efficiency of 26.46% at Nt equals 1.0 × 1013 cm−3 and an absorber thickness of 1.5 μm. While both AZnO and TiO2 showed high and comparable efficiencies across all examined conditions, WS2 consistently delivered lower efficiencies, particularly at low thicknesses and high defect densities. This behavior may be attributed to increased interfacial recombination or higher series resistance. These findings underscore the excellent performance and compatibility of AZnO and TiO2 as ETLs for CaHfSe3-based devices, particularly when combined with low defect densities and optimized absorber thickness.
 |
| Fig. 15 PCE contour maps showing the influence of CaHfSe3 absorber thickness and defect density for ETLs including (a) AZnO, (b) TiO2, and (c) WS2. | |
3.6. Effect of HTM and ETM thickness and carrier concentration on PV properties
3.6.1. Effect of HTM thickness. The HTL improves efficiency by collecting holes, restricting electron movement, and shielding the perovskite from environmental variables including moisture, heat, and oxygen.27 This section uses AZnO as the ETM and MoO3 as the HTM. HTL thickness was varied from 20–180 nm to evaluate how thickness influences the performance of a CaHfSe3-based PSC. Fig. 16 shows that all the significant metrics for performance, namely VOC, JSC, FF and PCE, were not significantly affected over the range of thicknesses and showed only slight differences in overall performance. Specifically, VOC = 1.482 V, JSC = 22.575 mA cm−2, FF = 90.73% and PCE = 30.37%. Overall, the MoO3 layer provided a suitable energy level for hole transport and sufficient holes across the range of HTL thickness without adding series resistance, or introducing any optical losses. Even the lowest thickness was suitable in ensuring effective hole extraction and hole transport (30 nm), as evidenced from the data showing that the performance metrics were not generally influenced by HTL thickness. Therefore, reduced HTL thickness is a solution to cut material usage or production cost while remaining similar in efficacy (Fig. 17).
 |
| Fig. 16 Influence of MoO3 layer thickness on the solar cell photovoltaic characteristics. | |
 |
| Fig. 17 Effect of MoO3 acceptor doping concentration (NA) on photovoltaic parameters. | |
3.6.2. Impact of the concentration of HTM carriers. The results indicate that VOC (1.482 V) and JSC (22.575 mA cm−2) remain almost unchanged with increasing NA. This stability arises because these parameters are primarily governed by the absorber's intrinsic optoelectronic properties, including bandgap, absorption coefficient, and carrier lifetime which are not significantly affected by moderate changes in the MoO3 HTL doping level. Consequently, carrier generation, separation, and transport within the absorber remain stable, in agreement with previous simulation studies on the limited effect of HTL doping on VOC and JSC. However, the FF shows a slight improvement, rising from 90.39% to 90.73% as NA increases. The improvement is probably caused by increased hole transport efficiency and reduced series resistance at the HTL, owing to higher doping levels, enabling better charge extraction. The PCE rises from 30.26% to 30.36%, plateauing above a doping concentration of 1015 cm−3.
3.6.3. Impact of the concentration of ETM carriers. Fig. 18 illustrates the influence of the ETL donor doping density (ND) on the photovoltaic performance of solar cells utilizing TiO2, ZnO, and WS2. TiO2 demonstrates consistent characteristics at all doping levels, maintaining a steady PCE of 30.41%, thus affirming its durability. Furthermore, AZnO shows a slight improvement in fill factor and efficiency as ND increases, achieving a maximum power conversion efficiency of 30.48%. However, WS2, with a maximum efficiency of 25.71%, demonstrates poorer and more doping-sensitive performance. In contrast VOC is mainly determined by the absorber bandgap and the splitting of the quasi-Fermi levels, which are controlled by carrier generation and recombination within the absorber. Since ETL doping does not directly affect these intrinsic mechanisms, VOC remains unchanged. Similarly, JSC is governed by photon absorption and the diffusion/collection of photogenerated carriers in the absorber. ETL doping does not alter the absorption spectrum or the intrinsic carrier generation, so JSC also stays constant (Fig. 19).
 |
| Fig. 18 Effect of ETM donor doping concentration (ND) on photovoltaic parameters. | |
 |
| Fig. 19 The impact of series resistance on solar cell performance under varied ETLs. | |
3.7. Impact of series resistance on performance
The series resistance (RS) in a solar cell generally arises from three sources: carrier transport within the emitter and base layers, contact resistance at the metal–semiconductor interfaces, and the resistance of the front and rear metallic contacts. Minimizing RS is crucial, as high values cause internal voltage drops that reduce the fill factor (FF) and overall power conversion efficiency (PCE). Excessive resistive losses can significantly reduce the fill factor. In the current investigation, when RS escalates from 0 to 6 Ω cm2, VOC and JSC stay rather stable throughout all ETLs, however, the FF and PCE diminish gradually. TiO2 and AZnO show excellent starting efficiencies of 30.93% and 30.80%, respectively, which decline to 27.83% and 27.72% at the maximum RS. WS2 starts at a lower PCE of 26.60%, dropping to 23.63%. This demonstrates that increasing series resistance affects device performance largely via lowering FF.
3.8. Impact of shunt resistance on performance
The shunt resistance (RSH) corresponds to leakage pathways through defects, pinholes, or imperfect interfaces. A low RSH leads to significant losses in open-circuit voltage (VOC) and a decrease in FF, whereas a high RSH ensures that the majority of photogenerated carriers contribute to power output. The influence of shunt resistance (RSH) on the photovoltaic performance of solar cells using AZnO, TiO2, and WS2 as electron transport layers (ETLs) was studied. Fig. 20 shows that raising RSH from 100 to 900 Ω cm2 significantly improved FF and PCE across all configurations. This behavior is explained by suppressing leakage currents, which are usually linked to manufacturing flaws, and avoiding the junction using low-resistance shunt routes. In particular, when RSH rose, FF for the AZnO-based device improved from 40.59% to 84.71%, and PCE improved from 13.59% to 28.88%. At RSH = 900 Ω cm2, the TiO2- and WS2-based cells showed comparable improvements, with final PCE values of 28.74% and 25.05%, respectively. The fill factor is the metric most sensitive to changes in shunt resistance, as seen by the relatively consistent VOC and JSC over the RSH range.
 |
| Fig. 20 Effect of RSH on the PSCs parameters. | |
3.9. Effect of temperature
One of the most difficult issues with PSC is guaranteeing its long-term stability, particularly at high temperatures. High temperatures can induce chemical and structural changes in perovskite materials, leading to a significant reduction in device performance. Furthermore, thermal stress may damage the integrity of the interfaces between distinct layers, causing greater charge carrier recombination and impeding charge transfer.28 To better understand the thermal response of PSC in actual working situations, a rigorous performance study was performed across a temperature range of 300 K to 400 K. Fig. 21. It has been demonstrated that increasing the temperature (from 300 to 400 K) reduces the VOC, fill factor (FF), and PCE, while the JSC remains essentially constant. This drop is mostly because of more recombination and energy loss at high temperatures, making removing charges and lowering the potential difference across the cell. TiO2 and AZnO have similar performance, with VOC values of 1.48 V, PCEs surpassing 30% at 300 K, and strong thermal stability. WS2 had lower results (VOC of 1.34 V, PCE ∼25.7%) and more noticeable deterioration with temperature, indicating inefficient charge extraction and higher recombination. In contrast, the simulated JSC remains nearly constant between 300 K and 400 K because photon absorption and carrier generation in the absorber are not significantly affected by temperature, while temperature mainly influences VOC through the increase of the saturation current density.
 |
| Fig. 21 Impact of temperature on solar cell performance using TiO2, AZnO, and WS2 as ETLs. | |
3.10. Examination of the rates of recombination and generation
Elevating electrons from the BV to the BC causes holes to develop in the valence band, which in turn creates electron–hole pairs. This process is the essential mechanism for creating charge carriers in the substance. In SCAPS-1D, the generation rate G(x) is determined by the input photon flux Nphot (λ,x), which indicates the quantity of photons accessible for absorption at each depth x and wavelength λ. This photon flux is utilized to estimate the spatial generation profile within the absorber layer according to eqn (7): |
G(λ,x) = α(λ, x)·Nphot(λ, x)
| (7) |
It is essential to integrate insights from carrier generation-recombination mechanisms and band structure studies to optimize absorber thickness, doping levels, and interface engineering, while selecting suitable ETLs and HTLs. This comprehensive knowledge enhances the efficiency, stability, and overall performance of CaHfSe3-based solar cells. Fig. 22 illustrates carrier generation and recombination rates at various depths (0 to 1.1 μm) in three CaHfSe3-based CP solar cells. Investigations indicate that generation rates for all devices peak between 0.15 and 0.3 μm. In contrast, the recombination rate acts as a limiting factor, neutralizing these electron–hole pairs and preventing them from increasing the photocurrent. Carrier density and lifespan are the primary factors influencing the recombination rate. Moreover, defects present within the absorber and at its interfaces considerably increase the rate of electron–hole recombination. These imperfections, often caused by impurities, structural faults, or grain boundaries, lead to a non-uniform distribution of the recombination rate across the material. The PSC with WS2 and AZnO ETL had the greatest recombination rates in the 1 to 1.1 μm range, where AZnO as the ETL exhibited the highest rate. This occurs when additional electrons in the BC bridge the Eg and join the BV, becoming stable and taking the position of a hole.
 |
| Fig. 22 Generation and recombination profiles in CaHfSe3 absorber using TiO2, AZnO, and WS2 as ETLs. | |
3.11. Back contact effect on PV parameters
Eight distinct metals were used as rear electrodes (Table 3). In simulations for structure based on FTO/AZnO/CaHfSe3/MoO3, exhibiting work function values spanning from 4.65 to 5.70 eV, to assess the impact of the back contact Work function on device performance. The results are summarized in Fig. 23. It has been discovered that the PCE improves with an increasing work function, eventually reaching a saturation point of roughly 5.2 eV. At lower levels, the reduced PCE results from creating a Schottky barrier at the metal/absorber interface, which prevents effective hole extraction and promotes interfacial recombination. Furthermore, as the work function exceeds ∼5.0 eV, the energy alignment at the back contact becomes more favorable for hole transport, lowering barrier height and improving carrier extraction. This leads to a fast rise in efficiency, reaching a maximum of around 30.37%. Beyond this level, the device's performance plateaus, indicating that the contact performs ohmically. Significantly, while gold ensures superior device performance, it accounts for around 20% of the overall production cost in perovskite solar cells. Electrodes composed of Ni, Pd, and Pt provide cost-effective alternatives, demonstrating promising efficiency levels that render them feasible substitutes for Au without significantly compromising device performance.
Table 3 The contacts used in the setups and their work function
Back electrode |
Cu |
Fe |
C |
Au |
W |
Ni |
Pd |
Pt |
Work function (eV) (ref. 29) |
4.65 |
4.82 |
5.00 |
5.30 |
5.22 |
5.50 |
5.60 |
5.70 |
 |
| Fig. 23 Influence of the back electrode work function on solar cell PCE. | |
3.12. Quantum efficiency and J–V properties
The current–voltage characteristics of the simulated devices are illustrated in Fig. 24(a). All three devices exhibited relatively high short-circuit current densities (JSC ∼22.49–22.58 mA cm−2), indicating that the CaHfS3 absorber effectively captures light and generates carriers. However, there are significant differences in overall current stability and open-circuit voltage (VOC). The cell utilizing AZnO as the ETL maintains a stable current plateau across a wide voltage range and achieves the highest VOC (≈1.46 V). This suggests reduced interfacial recombination losses and excellent band alignment at the AZnO/CaHfS3 interface, indicating that AZnO is the most effective ETL among those studied. Fig. 24(b) shows how wavelength affects quantum efficiency (QE) for our investigation's three most successful devices. According to the data, AZnO and WS2 outperform TiO2 in the UV (300–360 nm) region, with QE values exceeding 90% compared to TiO2, approximately 60%. This improvement may be attributed to the superior electrical and optical characteristics of AZnO and WS2, such as their broad band alignment compatibility with CaHfS3 and high optical transparency in the ultraviolet region, resulting in more effective charge carrier extraction in the high-energy area. Within the visible spectrum (360–700 nm), all three ETL designs demonstrate virtually optimal quantum efficiency, with values around 100%. This suggests that the CaHfS3 absorber efficiently transforms visible light photons into charge carriers and that the ETLs retain good charge transport with negligible recombination losses throughout this range. However, after 700 nm, the QE starts to drop dramatically for all setups. CaHfS3's bandgap limits its ability to absorb low-energy photons in the near-infrared range, causing a noticeable drop at 770–800 nm.
 |
| Fig. 24 (a) J–V characteristics and (b) quantum efficiency for every device under study. | |
3.13. Comparing SCAPS-1D results with earlier research
A comparison with previously reported perovskite-based devices is summarized in Table 4. The FTO/TiO2/CaHfSe3/MoO3/Au and FTO/AZnO/CaHfSe3/MoO3/Au devices developed in this study exhibit outstanding photovoltaic performance, with VOC ≈ 1.52 V, JSC ≈ 23.17–23.23 mA cm−2, and PCE values exceeding 32%. The TiO2-based structure achieves a slightly higher fill factor (91.41% vs. 91.02%) and overall efficiency (32.39% vs. 32.20%) compared to the AZnO-based device, reflecting lower resistive losses and more efficient charge extraction. These PCEs are significantly higher than those reported in previous studies, particularly for systems using CaHfSe3 as the absorber layer. The top-performing FTO/TiO2/CaHfSe3/MoO3/Au device, achieving 32.39% efficiency, highlights the superior optoelectronic properties of CaHfSe3 and its favorable energy alignment with MoO3. These results confirm the strong potential of the proposed materials and device architecture for high-performance, lead-free thin-film solar cells.
Table 4 Device comparison: VOC, JSC, FF, and PCE
Structure |
VOC (V) |
JSC (mA cm−2) |
FF (%) |
PCE (%) |
Year |
Ref. |
FTO/TiO2/CaHfSe3/MoO3/Au |
1.52 |
23.169 |
91.41 |
32.39 |
— |
— |
FTO/AZnO/BaHfSe3/MoO3/Au |
1.52 |
23.23 |
91.02 |
32.20 |
— |
— |
FTO/WS2/BaHfSe3/MoO3/Au |
1.34 |
23.22 |
84.85 |
29.12 |
— |
— |
FTO/TiO2/CaHfSe3/NiO2/Au |
1.30 |
20.62 |
83.78 |
22.58 |
2025 |
19 |
FTO/TiO2/BaZrS3/SpiroOMeTAD/Au |
0.70 |
22.00 |
79.40 |
12.42 |
2023 |
30 |
FTO/TiO2/CaZrS3/CuO |
0.60 |
35.73 |
80.88 |
17.53 |
2024 |
18 |
3.14. Validation against the ideal cell model
A comparison was made between the simulated photovoltaic properties of CaHfSe3-based solar cells and those of an ideal single-junction solar cell. The optimized combination, FTO/TiO2/CaHfSe3/MoO3/Au, attained a remarkable PCE of 32.39%, with a VOC of 1.52 V, a JSC of 23.17 mA cm−2, and an FF of 91.41%. This indicates that the results conform to the theoretical boundaries of an ideal single-junction solar cell.31 These results demonstrate the great promise of CaHfSe3 as an absorber material for lead-free perovskite solar cells, since they approach the theoretical Shockley Queisser limit.
4. Conclusion
This makes our study unique as it intends to examine the photovoltaic characteristics of CaHfSe3 absorber material in state-of-the-art solar cells. We explored the potential of devices constructed FTO/TiO2, AZnO and WS2 layers with CaHfSe3/MoO3/Au using SCAPS-1D simulations. We then analysed the effects of the absorber thickness, the density of defects in the absorber, carrier concentration in electron & hole transport layers, levels of acceptor doping, etc., and how these factors impact the efficiency of the devices. The effects of RS and RSH, and temperature of operation in addition to other variables explored, were also examined in relation to overall photovoltaic performance. A detailed study of the band structure, carrier generation and recombination rates, J–V characteristics, and quantum efficiency (QE) was presented herein. The optimized device structure FTO/TiO2/CaHfSe3/MoO3/Au yielded a power conversion efficiency (PCE) of 32.39%, with a VOC of 1.52 V, JSC of 23.17 mA cm−2, and fill factor (FF) of 91.41%. In contrast, the FTO/AZnO/CaHfSe3/MoO3/Au structure had a PCE of 32.20% with near identical electrical properties. Overall, the results suggest the high optoelectronic quality of CaHfSe3, and its compatibility with multiple ETLs. Additionally, cheap substitutes for gold in the back contact were examined, including carbon, with only modest losses in efficiency. Overall, this detailed and systematic work contributes to the understanding of CaHfSe3-based perovskite solar cells. It represents a potential pathway towards a lead-free, efficient, low-cost, and sustainable photovoltaic technology, and will support clean energy targets of the future.
Conflicts of interest
The authors declare no conflict of interest.
Data availability
The data will be available from the corresponding author on reasonable request. SCAPS-1D is available from https://scaps.elis.ugent.be/.
References
- A. Basem, S. Opakhai, Z. M. S. Elbarbary, F. Atamurotov and N. E. Benti, A comprehensive analysis of advanced solar panel productivity and efficiency through numerical models and emotional neural networks, Sci. Rep., 2025, 15(1), 259 Search PubMed.
- H. Togun, A. Basem, M. J. Jweeg, N. Biswas, A. M. Abed, D. Paul, H. I. Mohammed, A. Chattopadhyay, B. K. Sharma and T. Abdulrazzaq, Advancing organic photovoltaic cells for a sustainable future: The role of artificial intelligence (AI) and deep learning (DL) in enhancing performance and innovation, Sol. Energy, 2025, 291, 113378 Search PubMed.
- W. Li, Z. Xu, Y. Yan, J. Zhou, Q. Huang, S. Xu, X. Zhang, Y. Zhao and G. Hou, Passivating contacts for crystalline silicon solar cells: an overview of the current advances and future perspectives, Adv. Energy Mater., 2024, 14(18), 2304338 Search PubMed.
- F. A. Nelson, A. Basem, D. J. Jasim, T. E. Gber, M. T. Odey, A. F. Al Asmari and S. Islam, Chemical effect of alkaline-earth metals (Be, Mg, Ca) substitution of BFe2XH hydride perovskites for applications as hydrogen storage materials: A DFT perspective, Int. J. Hydrogen Energy, 2024, 79, 1191–1200 Search PubMed.
- A. Kumari, A. Ali, J. A. Abraham, A. K. Mishra, M. Kallel, W. M. Shewakh, S. Formanova and R. Sharma, A Comprehensive DFT Analysis of Novel Vacancy-Ordered Double Perovskites Na2SnX6 (X= Br, I) for the Opto-electronic and Thermoelectric Properties Applications, J. Inorg. Organomet. Polym. Mater., 2025, 1–13 Search PubMed.
- B. P. Finkenauer, Y. Zhang, K. Ma, J. W. Turnley, J. Schulz, M. Gómez, A. H. Coffey, D. Sun, J. Sun and R. Agrawal, Amine-Thiol/Selenol Chemistry for Efficient and Stable Perovskite Solar Cells, J. Phys. Chem. C, 2023, 127(2), 930–938 Search PubMed.
- Best Research-Cell Efficiency Chart, 2025, https://www.nrel.gov/pv/cell-efficiency.html.
- M. Alla, S. Bimli, V. Manjunath, E. Choudhary, S. Sharma, G. R. Wakale, A. Miglani, M. Rouchdi and B. Fares, Examining the potential of non-toxic stable double perovskite solar cells based on Cs2CuSbX6, Mater. Today Commun., 2023, 36, 106608 Search PubMed.
- P. P. Boix, S. Agarwala, T. M. Koh, N. Mathews and S. G. Mhaisalkar, Perovskite solar cells: beyond methylammonium lead iodide, J. Phys. Chem. Lett., 2015, 6(5), 898–907 Search PubMed.
- S. A. Dar and B. S. Sengar, Breakthrough in sustainable photovoltaics: Achieving 30.86% efficiency with innovative lead-free bilayer perovskite solar cells using SCAPS-1D and DFT framework, Sol. Energy Mater. Sol. Cells, 2025, 282, 113352 Search PubMed.
- B. K. Bareth and M. N. Tripathi, High photovoltaic performance of an emerging lead-free chalcogenide perovskite BaHfS3 under high pressure, Sol. Energy Mater. Sol. Cells, 2025, 282, 113445 Search PubMed.
- Y.-Y. Sun, M. L. Agiorgousis, P. Zhang and S. Zhang, Chalcogenide perovskites for photovoltaics, Nano Lett., 2015, 15(1), 581–585 Search PubMed.
- S. Karthick, S. Velumani and J. Bouclé, Chalcogenide BaZrS3 perovskite solar cells: A numerical simulation and analysis using SCAPS-1D, Opt. Mater., 2022, 126, 112250 Search PubMed.
- D. C. Hayes, S. Agarwal, K. C. Vincent, I. M. Aimiuwu, A. A. Pradhan, M. C. Uible, S. C. Bart and R. Agrawal, A reliable, colloidal synthesis method of the orthorhombic chalcogenide perovskite, BaZrS 3, and related ABS 3 nanomaterials (A= Sr, Ba; B= Ti, Zr, Hf): a step forward for earth-abundant, functional materials, Chem. Sci., 2025, 16(3), 1308–1320 Search PubMed.
- N. A. Moroz, C. Bauer, L. Williams, A. Olvera, J. Casamento, A. A. Page, T.
P. Bailey, A. Weiland, S. S. Stoyko and E. Kioupakis, Insights on the synthesis, crystal and electronic structures, and optical and thermoelectric properties of Sr1–x Sb x HfSe3 orthorhombic perovskite, Inorg. Chem., 2018, 57(12), 7402–7411 Search PubMed.
- R. O. Balogun, M. A. Olopade, O. O. Oyebola and A. D. Adewoyin, First-principle calculations to investigate structural, electronic and optical properties of MgHfS3, Mater. Sci. Eng., B, 2021, 273, 115405 Search PubMed.
- H. R. Abdul Ameer, A. N. Jarad, K. H. Salem, H. S. Hadi, M. A. Alkhafaji, R. S. Zabibah, K. A. Mohammed, K. Kumar Saxena, D. Buddhi and H. Singh, A role of back contact and temperature on the parameters of CdTe solar cell, Adv. Mater. Process. Technol., 2024, 10(2), 497–505 Search PubMed.
- N. Chawki, M. Rouchdi, M. Alla and B. Fares, Numerical modeling of a novel solar cell system consisting of electron Transport material (ETM)/CaZrS3-based chalcogenide perovskites using SCAPS-1D software, Sol. Energy, 2024, 274, 112592 Search PubMed.
- D. Srinivasan, A.-D. Rasu Chettiar, E. N. Vincent Mercy and L. Marasamy, Scrutinizing the untapped potential of emerging ABSe3 (A= Ca, Ba; B= Zr, Hf) chalcogenide perovskites solar cells, Sci. Rep., 2025, 15(1), 3454 Search PubMed.
- M. K. Hossain, M. H. K. Rubel, G. I. Toki, I. Alam, M. F. Rahman and H. Bencherif, Effect of various electron and hole transport layers on the performance of CsPbI3-based perovskite solar cells: a numerical investigation in DFT, SCAPS-1D, and wxAMPS frameworks, ACS Omega, 2022, 7(47), 43210–43230 Search PubMed.
- H. El-assib, M. Alla, S. Tourougui, M. Alla, F. Elfatouaki, S. A. Dar, A. Chauhan, N. Chawki, N. Shrivastav and V. Manjunath, High-performance optimization and analysis of Cs2CuSbCl6-Based lead-free double perovskite solar cells with theoretical efficiency exceeding 27%, Renew. Energy, 2025, 239, 122092 Search PubMed.
- B. K. Ravidas, M. K. Roy and D. P. Samajdar, Investigation of photovoltaic performance of lead-free CsSnI3-based perovskite solar cell with different hole transport layers: First Principle Calculations and SCAPS-1D Analysis, Sol. Energy, 2023, 249, 163–173 Search PubMed.
- M. S. Uddin, M. K. Hossain, M. B. Uddin, G. F. Toki, M. Ouladsmane, M. H. Rubel, D. I. Tishkevich, P. Sasikumar, R. Haldhar and R. Pandey, An in-depth investigation of the combined optoelectronic and photovoltaic properties of lead-free Cs2AgBiBr6 double perovskite solar cells using DFT and SCAPS-1D frameworks, Adv. Electron. Mater., 2024, 10(5), 2300751 Search PubMed.
- S. T. Jan and M. Noman, Comprehensive analysis of heterojunction compatibility of various perovskite solar cells with promising charge transport materials, Sci. Rep., 2023, 13(1), 19015 Search PubMed.
- K. Deepthi Jayan, Design and Comparative Performance Analysis of High-Efficiency Lead-Based and Lead-Free Perovskite Solar Cells, Phys. Status Solidi, 2022, 219(7), 2100606 Search PubMed.
- J. Liang, Y. Wang, X. Liu, J. Chen, L. Peng and J. Lin, Theoretical analysis of doping of perovskite light-absorbing layer for highly efficient perovskite solar cells, J. Phys. Chem. Solids, 2024, 188, 111901 Search PubMed.
- X. Cai, T. Hu, H. Hou, P. Zhu, R. Liu, J. Peng, W. Luo and H. Yu, A review for nickel oxide hole transport layer and its application in halide perovskite solar cells, Mater. Today Sustain., 2023, 23, 100438 Search PubMed.
- G. Pindolia, S. M. Shinde and P. K. Jha, Optimization of an inorganic lead free RbGeI3 based perovskite solar cell by SCAPS-1D simulation, Sol. Energy, 2022, 236, 802–821 Search PubMed.
- N. Chawki, M. Rouchdi, M. Alla and B. Fares, Simulation and analysis of high-performance hole transport material SrZrS3-based perovskite solar cells with a theoretical efficiency approaching 26%, Sol. Energy, 2023, 262, 111913 Search PubMed.
- N. Thakur, P. Kumar and P. Sharma, Simulation study of chalcogenide perovskite (BaZrSe3) solar cell by SCAPS-1D, Mater. Today: Proc., 2023 DOI:10.1016/j.matpr.2023.01.012.
- A. Morales-Acevedo, Fundamentals of solar cell physics revisited: Common pitfalls when reporting calculated and measured photocurrent density, open-circuit voltage, and efficiency of solar cells, Sol. Energy, 2023, 262, 111774 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.