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

Performance evaluation of Eu2NiMnO6-based lead-free perovskite solar cells: a SCAPS-1D study

Md. Abu Bakkar Siddiqueab, Nazmul Shahadathab, Md. Tarekuzzamanab, Md. Raihan Kabirab, Sohail Ahmadc, Rashel Mohammad Khokan*abd, Md. Rasheduzzamanab, S. M. G. Mostafab, Mohammad Jalal Uddinb and Md. Zahid Hasan*ab
aMaterials Research and Simulation Lab, Department of Electrical and Electronic Engineering, International Islamic University Chittagong, Kumira, Chittagong, 4318, Bangladesh. E-mail: rashel.eee13@gmail.com; zahidhasan.02@gmail.com
bDepartment of Electrical and Electronic Engineering, International Islamic University Chittagong, Kumira, Chittagong, 4318, Bangladesh
cDepartment of Physics, College of Science, King Khalid University, P O Box 9004, Abha, Saudi Arabia
dFaculty of Engineering, Kitami Institute of Technology, 165 Kouen-cho, Kitami, Hokkaido 090-8507, Japan

Received 24th July 2025 , Accepted 18th September 2025

First published on 25th September 2025


Abstract

Lead-free double Perovskite materials are currently attracting considerable research interest owing to their environmentally friendly attributes. In this investigation, we have analyzed a tremendous double Perovskite material Eu2NiMnO6 (ENMO) as the absorber layer of a solar cell with the help of SCAPS-1D (a solar cell capacitance simulator). The material has become remarkable because of its narrow experimental band gap of 1.01 eV. Throughout the study, we investigated the effect of appropriate ETLs (Electron Transport Layers) and HTLs (Hole Transport Layers) with the absorber layer. For optimizing the device, tungsten disulfide (WS2), C60 (Buckminsterfullerene), and PCBM (Phenyl-C61-butyric acid methyl ester) are used as ETLs, and Copper Ferrite Tin Sulfide (CFTS) is used as the HTL. Besides evaluating the effects of ETL and HTL, other important factors like absorber thickness, shunt and series resistance, temperature, capacitance, Mott–Schottky characteristics, recombination and generation rates, current density–voltage (JV), and quantum efficiency are also analyzed. The simulation demonstrates that the optimal output parameters (VOC, JSC, FF, and PCE) for the WS2 ETL based device are 0.720 V, 45.287 mA cm−2, 81.02%, and 26.45%. It is the most detailed investigation with the highest reported efficiency, significantly higher than previous research work. Using this extensive simulation study, researchers will be able to create Perovskite Solar Cells (PSCs) that are both affordable and effective while also expanding the possibilities for solar technology.


1 Introduction

The increasing reliance on conventional energy resources and the repeated depletion of fossil fuels, resulting from intense industrial activities, reveals the urgent environmental and economic challenges of today.1–5 The conventional method of generating electricity with fossil fuels is frequently seen as unsustainable over time due to the limited availability of these resources and the environmental problems resulting from their emissions.6 Environmental concerns have led several related organizations to promote extensive research on technologically advanced, sustainable power facilities. The substitution of fossil fuels with sustainable energy alternatives is a fundamental goal of science and technology. Solar energy is the most suitable solution to this challenge because it has the capability to meet global energy demand. Solar cells may be an effective method for converting the sun's plentiful energy into productive, low-cost, and environmentally friendly electric power.7,8

In regard to power conversion efficiency (PCE), research into perovskite photovoltaic (PV) technology has demonstrated significant potential as an economical substitute for silicon (Si)-based solar cell technology. As a significant advancement in third-generation solar cells, perovskite solar cells (PSCs) have a photoelectronic conversion efficiency (PCE) of 25.7%, which is equivalent to that of silicon-based models.9–11 The remarkable photophysical and optical characteristics of perovskite material have been widely studied,11–14 together with collective efforts to improve interfacial engineering methods, optimize materials, and fine-tune device architecture,15–20 all of which have contributed to the significant rise in Power Conversion Efficiency (PCE). Generally, lead-based PSCs produce higher efficiency.21–25 However, lead-based cells face performance challenges in the presence of moisture and light, and lead poisoning also poses a serious hurdle to commercialization.26–28 For this reason, initiatives to investigate stable, lead-free perovskite compounds with effective photovoltaic performance are continuing.

Perovskite oxide, commonly known as ABO3, has been intensively explored due to its unique geometry and physics.29–31 Perovskites with matching positive and negative charges are regarded as optimal perovskite compounds. However, perovskite compounds exhibit ferroelectricity due to their non-linear positive and negative charge centers, resulting in a net dipole moment.32 Moreover, investigations into these substances have unveiled a group of closely associated and advanced materials known as double perovskites. Double perovskite materials were discovered in the 1950s33 and are symbolized by the formula A2BB'O6, with A representing alkaline earth and alkali metals, and B and B′ representing transition, alkaline or alkali metals.33,34 These double perovskite materials are highly coveted for integration into heterostructures for perovskite solar cells (PSCs), particularly for their application in absorbent layers.

Among various double perovskite structures, Cs2AgBiX6 (where X is Cl, Br, or I) has been extensively studied, while La2NiMnO6 systems have also earned significant attention, particularly for their application in perovskite solar cells.32,35 Recent studies, such as those by Hossain et al.,32 have highlighted key insights into the design of La2NiMnO6-based devices, focusing on different charge transport layers and utilizing DFT and SCAPS-1D frameworks for performance optimization. These investigations emphasize the material's promising photovoltaic properties and potential for enhancing power conversion efficiency. Sheikh et al.36 discovered a lead-free inorganic double perovskite material, Ln2NiMnO6 (where Ln stands for La, Dy, Eu and Du), with a narrow band-gap range of 1.08 eV to 1.19 eV. All things considered, the materials' narrow band gap, ability to be deposited via chemical solutions, and high dielectric constant make them appealing for photovoltaic research.36

The experimental results of this study have shown commendable device performance for La2NiMnO6 (LNMO), Eu2NiMnO6 (ENMO), and Dy2NiMnO6 (DNMO) -based solar cells due to their optimized material properties and improved efficiency in terms of key parameters such as open-circuit voltage (VOC), current density (JSC), fill factor (FF), and power conversion efficiency (PCE).36 The efficiency and performance of perovskite solar cells (PSCs) are enhanced using appropriate electron and hole-transport layers. Eu2NiMnO6 (ENMO) is a superior perovskite material compared to others due to its lead-free structure, non-toxic nature, and its potential for eco-friendly applications. When compared with other rare-earth compounds, ENMO has a smaller band gap.36 These properties make it a promising candidate for photovoltaic research, as it can absorb a wide range of light, be deposited from solution, and has a high dielectric constant. The ETL is a vital component of PSCs, performing the dual function of removing electrons from the absorber and obstructing holes.37 Conversely, the HTL affects the manufacturing cost, stability, and efficiency of solar devices.38 Primary considerations in choosing an HTL for PSCs include the valence band offset with the absorber, hole mobility, and the associated cost,39 whereas the ETL needs to possess a conduction band offset between the absorber and ETL that is suitable for maintaining compatibility with the high electron mobility of other layers, while also being cost-effective.39 An effective transfer of charge carriers produced by light from the absorber to their designated contacts in PSCs is greatly influenced by ETL and HTL. Moreover, they prevent electrons and holes from moving toward their respective electrodes. As a result, it prevents charge recombination at the ETL/absorber and absorber/HTL interfaces. Meanwhile, it separates and directs the electrically charged particles to their specified sites of contact for collection.40

The present work examines the effectiveness of lead-free Eu2NiMnO6 PSCs utilizing the SCAPS-1D framework and various ETLs and HTLs for the first time. Throughout the investigation, the performance is assessed using WS2, PCBM, and C60 as ETLs, and CFTS as the HTL, with gold (Au) employed as the back-metal contact. Furthermore, we examined the performance of the HTL and ETL layers, as well as the influence of the absorber and ETL thickness, JV curves, generation and recombination rates, operational temperature, series resistance, shunt resistance, capacitance, Mott–Schottky analysis, and quantum efficiency.

As a double perovskite oxide, Eu2NiMnO6 (ENMO) is a promising alternative to lead-based perovskites because of its suitable bandgap, high stability, and environmental safety. Although ENMO has not yet been widely tested in solar cells, its successful synthesis has been reported using common oxide deposition methods. For example, sol–gel and solvothermal techniques have been used to produce ENMO with good structural quality and controlled composition.41 In addition, pulsed laser deposition (PLD) is a well-established method for preparing high-quality oxide thin films, such as EuO, NiO, and MnO, and can also be applied to ENMO.42 Similarly, spray pyrolysis has been demonstrated for making uniform oxide thin films, including Eu-doped TiO2, and is recognized as a simple and scalable deposition technique.43 Together, these methods show that ENMO can be prepared with reliable thin-film quality and could be integrated into stable, lead-free photovoltaic devices.

2 Numerical simulations

2.1. Numerical analysis using SCAPS-1D

The SCAPS-1D simulator was used within the computational model framework, applying Poisson's equation (eqn (1)) and continuity equations for holes (eqn (2)) and electrons (eqn (3)) to derive the photovoltaic parameters of the PSCs.44–50 For the purpose of calculating the PV parameters, the simulation algorithm additionally accounts for loss processes using the Shockley–Read–Hall (SRH) recombination.51,52 The symbols used in eqn (1)–(3) are as follows.53
 
image file: d5ra05366h-t1.tif(1)

For this case, the electronic potential is represented by ψ, the relative permittivity by εr, the permittivity of free space by ε0, the densities of ionized donors and acceptors by ND and Na, the electron and hole densities by n and p, the distributions of electrons and holes by ρp and ρn, and e is the electronic charge.

 
image file: d5ra05366h-t2.tif(2)
 
image file: d5ra05366h-t3.tif(3)

According to eqn (2) and (3), Jn and Jp denote the electron and hole current densities, respectively. Un and Up refer to the net recombination rates for electrons and holes, and G represents the generation rate.

The overall current density, which is influenced by both concentration gradients and electric fields, can be determined by applying the drift and diffusion current formulas, as described in eqn (4) and (5).54

 
Jn = qnµnE + qDnn (4)
 
JP = qpµpEqDPp (5)
Here, Dn and DP refer to the diffusion coefficients for electrons and holes. Moreover, the film's absorption constant was calculated using the new Eg-sqrt model, a revised form of the standard sqrt (hvEg) model. The correlation between these variables is shown by eqn (6),53 which follows the “Tauc laws”.
 
image file: d5ra05366h-t4.tif(6)

The photon energy is represented by hv, the bandgap by Eg, and the absorption coefficient by. Eqn (7) and (8) (ref. 53) provided below establishes the relationship between the model constants α0 and βo and the traditional constants A and B:

 
image file: d5ra05366h-t5.tif(7)
 
image file: d5ra05366h-t6.tif(8)

Fig. 1 outlines the SCAPS-1D simulation process in six key steps. It begins by launching the software, followed by identifying the research problem. The subsequent step involves setting the device's material properties and simulation conditions. The specific outputs to be calculated—such as JV curves or QE—are then defined. Once configured, the simulation is initiated. Finally, the results are visualized and analyzed through simulated output curves to gain insights into device performance.


image file: d5ra05366h-f1.tif
Fig. 1 Workflow for SCAPS-1D.

2.2. Device structure of Eu2NiMnO6

The layout of the optimized SC is outlined in Fig. 2(a). In this analysis, the Perovskite, along with the HTL, is chosen as the p-region, while the ETL functions as the n-region in Eu2NiMnO6-based devices. In this device setup, CFTS was used as the HTL, indium-doped tin oxide (ITO) for the front contact, Au as the back-metal contact (BMC), and WS2, C60, and PCBM as the ETL, with ENMO serving as the absorber layer. Eu2NiMnO6 crystallizes in a monoclinic double perovskite structure (space group (P21/n)), with ordered Ni2+/Mn4+ at the B-sites and a narrow band gap near 1.1 eV—attributes highly favorable for photovoltaic absorption.55,56 The material exhibits a high value of the room temperature, a relatively high dielectric constant (εᵣ ≈ 300 at ∼50 kHz experimentally, and ∼6.3 from DFT), which serves to reduce recombination, and extend carrier diffusion length.36,55,56 Furthermore, the monoclinic symmetry and optimized Ni–O–Mn bond lengths enhance orbital overlap and super exchange interactions, improving charge transport and carrier lifetime.57 Moreover, the robust oxide perovskite framework offers enhanced chemical and thermal stability compared with halide perovskites, making Eu2NiMnO6 a highly attractive absorber material for solar cells. Fig. 2(a) presents the schematic construction of the main device, and Fig. 2(b) displays the ITO/ETL (WS2, C60, PCBM)/ENMO/CFTS/Au device's energy band alignment. ITO/WS2/ENMO/CFTS/Au was determined to be the best computationally effective SC among all configurations. Table 1. contains the simulation's parameters for the absorber, ETLs, HTL, and front contact. Additionally, the interfacial defect layers' input parameters are given in Table 2. As the temperature is 300 K and the frequency is 1 MHz, A 1000 W m−2 power density characteristic of the AM1.5 G solar spectrum has been employed for all the simulations.
image file: d5ra05366h-f2.tif
Fig. 2 (a) Device layout of the Eu2NiMnO6 – based Perovskite solar cell, (b) energy band alignment of various ETL and HTL materials with Eu2NiMnO6 absorber.
Table 1 Input data for ITO, ETL, HTL, and absorber layers used in this work
Parameters (unit) ITO32 WS2 (ref. 32) C60 (ref. 32) PCBM32 Eu2NiMnO6 (ref. 41 and 62) CFTS32
Thickness (µm) 0.5 0.1 0.05 0.05 0.8 0.1
Bandgap, Eg (eV) 3.5 1.8 1.7 2 1.01 1.3
EA (eV) 4 3.95 3.9 3.9 3.52 3.3
εr 9 13.6 4.2 3.9 9 9
NC (cm−3) 2.2 × 1018 1 × 1018 8 × 1019 2.5 × 1021 1 × 1018 2.2 × 1018
NV (cm−3) 1.8 × 1019 2.4 × 1019 8 × 1019 2.5 × 1021 1 × 1018 1.8 × 1019
Electron thermal velocity (cm s−1) 1 × 107 1 × 107 1 × 107 1 × 107 1 × 107 1 × 107
Hole thermal velocity (cm s−1) 1 × 107 1 × 107 1 × 107 1 × 107 1 × 107 1 × 107
µn (cm2 V−1 s−1) 20 100 8 × 10−2 0.2 22 21.98
µh (cm2 V−1 s−1) 10 100 3.5 × 10−3 0.2 22 21.98
ND (cm−3) 1 × 1021 1 × 1018 1 × 1017 2.93 × 1017 0 0
NA (cm−3) 0 0 0 0 7 × 1016 1 × 1018
Nt (cm−3) 1 × 1015 1 × 1015 1 × 1015 1 × 1015 1 × 1015 1 × 1015


Table 2 Data for interface parameters used in the Eu2NiMnO6-based solar cell32
Interface Defect type Capture cross section: electrons/holes (cm2) Energetic distribution Reference for defect energy levels, Et Interface defect density (cm−2)
ETL/Eu2NiMnO6 Neutral 1 × 10−17 Single Above the VB maximum 1 × 1010
1 × 10−18
Eu2NiMnO6/HTL Neutral 1 × 10−18 Single Above the VB maximum 1 × 1010
1 × 10−19


2.3. Band alignment of Eu2NiMnO6 based solar cell

Fig. 2(b). shows a variety of solar cell structures, each employing different types of ETL, HTL, absorbers, and front and back contact materials. An exhaustive analysis of three ETLs and one HTL was conducted in our research. Analyzing various combinations in the ITO/ETL/Eu2NiMnO6/HTL/Au structure identifies the optimal theoretical configuration for the Eu2NiMnO6 (ENMO) absorber layer, as shown in Fig. 2(b). Our findings, illustrated in Fig. 2(b) demonstrates that WS2, which has an energy gap of 1.8 eV, provided superior performance as an ETL in conjunction with CFTS HTL in double perovskite ENMO devices. For optimal performance, the front electrode at the incident light plane needs to provide both high transmittance and superior electrical conductivity. Metal materials, including Au, are commonly used to compose the back electrode. The device's stability and efficiency can be improved by utilizing a high-quality back electrode, which also helps with carrier collection.58 The Au electrode (WF ∼5.1 eV) is deemed the most suitable for most Eu2NiMnO6 PSCs because of their mesoporous or planar structure, as depicted in Fig. 2(b). The performance gain observed in Fig. 2(b) arises from the favorable energy alignment of WS2 with the absorber. Specifically, WS2 introduces a small positive conduction band offset (CBO, −0.43 eV), which forms a moderate spike at the ETL/absorber interface. Such alignment is beneficial because it suppresses interfacial recombination by raising the barrier for electron back-transfer, while still allowing efficient electron extraction. Previous SCAPS-based studies have shown that this CBO within the optimal range significantly enhances device performance, whereas negative offsets (“cliffs”) lead to increased recombination losses, and excessively large spikes (>0.5 eV) can obstruct electron transport59,60 Similarly, the valence band offset (VBO) at the absorber/HTL interface plays a complementary role: a moderate positive VBO (+0.07 eV) ensures selective hole extraction while blocking electron leakage, thereby reducing recombination and maintaining high device efficiency. In consideration of these findings, the favorable CBO of WS2 and an optimized VBO enable efficient charge carrier separation, suppression of recombination, and enhanced photovoltaic performance, consistent with earlier SCAPS modeling reports.61

3 Result & discussion

3.1. Influence of VBO and CBO

Exposure of the solar cell to sunlight generates electrons and holes within the perovskite absorber layer. The conduction and valence band offsets (CBO and VBO) at the interfaces of ETL/absorber and absorber/HTL primarily dictate the efficiency of separating charge carriers.63 These offsets directly influence the device's performance.

The CBO for the ETL/absorber interface is expressed as.64

 
CBO = XAbsorber − XETL (9)
In the preceding instance, XAbsorber and XETL stand for the absorber's and ETL's electron affinities, subsequently, while CBO stands for conduction band offsets.

Three distinct barrier types are observed at the ETL/absorber interface: virtually flat, cliff-like, and spike-like.65 A negative CBO forms as a cliff-like barrier when XETL exceeds XAbsorber. This implies that the ETL possesses a lower conduction band minimum (CBM) compared to the absorber. In the absence of a CBO, a flat barrier results in no energy differences and, as a result, no barrier to charge transfer.

On the other hand, when ETL's CBM exceeds the absorber's (XETL < XAbsorber), a positive CBO corresponds to the appearance of a spike-like barrier. The VBO shown in Table 3 at the contact between the absorber and the HTL is defined as.64

 
VBO = XHTL − XAbsorber + Eg,HTLEg,Absorber (10)
In this context, VBO represents Valence Band Offsets, XHTL HTL indicates the electron affinity of the HTL, and Eg,HTL and Eg,Absorber denotes the bandgaps of the HTL and absorber.

Table 3 Provides the VBO and CBO values for each ETL
Absorber ETLs CBO VBO
Eu2NiMnO6 WS2 −0.43 0.07
C60 −0.38 0.07
PCBM −0.38 0.07


The calculations for CBO and VBO were performed with eqn (9) and (10).64 The CBO, as well as the VBO for WS2 is:

The CBO at the ETL/absorber interface is defined as = XAbsorber − XETL = 3.52–3.95 = −0.43 eV.In this instance, the CBO is negative, and the nature of the barrier is cliff-like.

The VBO at the absorber/HTL interface is defined as = XHTL − XAbsorber + Eg,HTLEg,Absorber = 3.3 − 3.52 + 1.3 − 1.01 = 0.07 eV. Here, there is a spike-like barrier that has a positive CBO. We can also determine the CBO and VBO of other ETLs in a similar way as presented in Table 3.

3.2. Band diagram

Fig. 3(a–c) displays the optimized Eu2NiMnO6(ENMO)-based PSCs' energy band diagrams. The Electron affinity of the ETL has to be greater than that of the ENMO in order to transfer the electron to the absorber-ETL interface, and the ionization energy has to be lower than that of the Eu2NiMnO6 (ENMO) in order to close the gaps in the material's contact. Energy level alignment has a major impact on the efficiency and performance of PSCs. WS2, C60, and PCBM ETLs have bandgaps of 1.8, 1.7, and 2 eV, accordingly, their performances with the same heterostructure are very similar to each other. The variation in energy levels among WS2, C60, and PCBM arises from differences in their electron affinities and band gaps, which influence how their conduction bands align with the ENMO absorber. WS2, with an electron affinity of 3.95 eV and a bandgap of 1.8 eV, creates a conduction band offset of −0.43 eV, promoting efficient electron extraction but potentially increasing interfacial recombination. C60 and PCBM, with slightly lower electron affinities (∼3.90 eV), generate smaller offsets (−0.38 eV), which reduce recombination risk but may slightly limit carrier transfer. PCBM's wider bandgap also improves hole blocking at the interface. These differences in band alignment directly affect carrier transport and device efficiency, helping to explain why WS2-based cells achieved the highest performance in our simulations.66,67 In general, the thickness of C60 and PCBM layers is typically less than 100 nm.68 However, in this theoretical study, we observed that when the thickness of WS2 is set to 50 nm, the efficiency is significantly lower than when a 100 nm thickness is used. Therefore, to achieve better performance and more accurate results, we have chosen to use a WS2 thickness of 100 nm for this investigation. In Fig. 3(a–c), the quasi-Fermi levels Fn and Fp are aligned with the valence band energy of each device. Both the conduction band's (EC) and valence band's (EV) energy, accordingly. For each ETL, Fp was positioned over the EV, while Fn and EC kept up their harmonically similar operations.
image file: d5ra05366h-f3.tif
Fig. 3 Energy band diagrams for (a) WS2, (b) C60, and (c) PCBM.

3.3. Influence of absorber and ETL thickness on the performance of solar cells

The thickness of both the ETL and absorber layers plays a crucial role in enhancing the PV output features of the SCs. To accomplish the best-performing solar collectors, PV outputs must be optimized.32 To attain maximum efficiency in SCs, optimizing the performance of photovoltaic (PV) systems is necessary.32 The first and most important stage of creating high-performance SCs is selecting the appropriate absorber, ETL, and HTL combination. During this analysis, we selected WS2, C60, and PCBM as ETL, Eu2NiMnO6 as an absorber, and CFTS as HTL. Contour maps for VOC, JSC, FF, and PCE of Eu2NiMnO6 (ENMO)-based PSCs are shown in Fig. (4–7), with variations plotted against absorber and ETL thickness of (0.4–1.2) and (0.025–0.125) µm. The thickness of both the absorber and ETL strongly affects the rates of carrier generation and recombination in the device. A thicker absorber allows more photons to be absorbed, which increases the number of electron–hole pairs generated. However, when the absorber becomes too thick, carriers generated deep inside face longer transport paths, which increases bulk recombination before they reach the junction.66,69 On the other hand, the ETL thickness mainly controls charge extraction and interfacial recombination. An ETL that is too thin may not effectively block holes, leading to interfacial recombination, while an excessively thick ETL increases resistance to electron transport, which also promotes recombination losses.70–72 Therefore, as shown in Fig. 7, the contour mapping of PCE reflects a balance: sufficient absorber thickness is needed for maximum carrier generation, while optimized ETL thickness ensures efficient extraction and minimal recombination.
image file: d5ra05366h-f4.tif
Fig. 4 Contour mapping of VOC shows effects of varying ETL and absorber thicknesses for ETLs including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f5.tif
Fig. 5 Contour mapping of JSC shows effects of varying ETL and absorber thicknesses for ETLs including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f6.tif
Fig. 6 Contour mapping of FF shows effects of varying ETL and absorber thicknesses for ETLs including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f7.tif
Fig. 7 Contour mapping of PCE shows effects of varying ETL and absorber thicknesses for ETLs including (a) WS2, (b) C60, and (c) PCBM.

The contour plots in Fig. 4(a–c) demonstrate the effect of altering both ENMO absorber layer and ETL thickness on the open-circuit voltage (VOC) of the solar cells. According to Fig. 4(a), VOC levels reached their maximum at ETL thicknesses of 0.025–0.125 µm and absorber thicknesses of around 0.4–0.45 µm. Compared to all other structures, the ITO/WS2/ENMO/CFTS/Au PSC structure recorded the highest VOC value of 0.7299 V. Out of all the PSCs under study, the ETL of 0.025–0.50 µm and absorber of 0.4–0.6 µm is the optimum thickness range to obtain best VOC ∼0.71 V for ITO/C60/ENMO/CFTS/Au device and The PCBM ETL registered the lowest VOC, which was 0.6865 V, its absorber thickness was between 0.4-0.55 µm and ETL thickness between 0.025-0.037 µm, according to Fig. 4(c). Since VOC increased as the ETL layer's thickness reduced, as graph Fig. 4(a–c) illustrates. It happens due to, enhanced absorber layer thickness results in increased carrier recombination rates, which raise the saturation current affecting the photocurrent.32

Fig. 5 illustrates the different thicknesses of the ENMO and ETL layers in the tested SC setups impact JSC. Fig. 5(a) demonstrates WS2 as ETL in the configuration, featuring absorber and ETL thicknesses near (0.95–1.2) µm and (0.025–0.125) µm, accordingly, resulting in a greater JSC value of 45.67 mA cm−2. Observations were made when the absorber and ETL thicknesses were around (0.9–1.2) and (0.025–0.095) µm. A JSC of 45.51 mA cm−2 was achieved with PCBM as ETL. The minimum JSC value of 45.45 mA cm−2 was found when C60 as ETL. The spectral response at longer wavelengths causes the JSC values for each SC to go up as the thickness of the absorber grows, whereas partial light absorption causes the JSC values to drop with an increase in ETL thickness.73

The variations in FF are driven by the interplay of material properties, energy level alignment, charge extraction efficiency, and interface quality.74,75 The contour diagrams in Fig. 6 exhibit the FF changes when the absorber and ETL thickness are changed. Fig. 6(a) shows that the WS2 ETL-based device had an FF of 81.10%, which is the highest among these three SC configurations for absorber and ETL thicknesses of nearly 0.52–0.68 µm and 0.1–0.0125 µm. While the PCBM-based ETL device had an FF of 80.80% with absorber and ETL thicknesses of around 0.4–0.55 µm and 0.01–0.037 µm, respectively, and among the ETLs, C60 had the lowest FF, recorded at 79.06%. WS2 outperforms PCBM and C60 due to its superior electronic properties, better energy alignment, and lower recombination losses, while the impact of ETL thickness remains minimal in these configurations.76 Interestingly, ETL thickness is not a major factor in maximizing the FF values, for all three different solar configurations.

The thickness of the absorber layer, which is determined by the carriers generated through photosynthesis, has been carefully optimized to the ideal level for creating a solar cell with improved efficiency.77 Contour plots in Fig. 7 depict PCE changes due to absorber and ETL thickness variations. The solar structure WS2-based-ETL, with an absorber thickness ranging from 0.75 to 1.05 µm and an ETL thickness of around 0.03 to 0.125 µm, achieved the maximum PCE among all the modified solar structures. It achieved a PCE of approximately 26.47% as shown in Fig. 7(a). Regarding the thickness of the absorber and ETL, C60 and PCBM based ETL device exhibit a similar PCE of around 25.02% and 24.34%, respectively (Fig. 7(b and c)). Increasing the absorber thickness indicates higher efficiency because the thick absorber layers improve carrier recombination, whereas excessively thin layers cannot generate carriers efficiently, which lowers the overall device efficiency. However, the PCBM ETL-based solar setup, with an absorber layer thickness of around 0.55–1 µm and an ETL thickness of around 0.025–0.032 µm, displays the smallest PCE of about 24.34% Fig. 7(c).

3.4. Influence of acceptor density and absorber thickness on the performance of solar cells

This study investigates the effects of absorber thickness and acceptor density (NA) in ENMO-based SCs. The observed effect's statistical significance is presented in (Fig. 8–11). To explore the effect of these parameters on the PV performance characteristics of the three optimized PSCs throughout the simulation in Fig. 8–11, the absorber thickness was adjusted between 0.4 to 1.2 µm, and NA varied from 7 × 1014–7 × 1018 cm−3. The influence of absorber thickness and NA on VOC is shown in Fig. 8(a–c). Fig. 8(a) illustrates that WS2-based ETLs produce the largest VOC of 0.7480 V. Whenever the absorb thickness is determined, it varies from 0.4 to 1.2 µm, and the NA varies between 7 × 1017–7 × 1018 cm−3. The VOC for the C60 ETL-based structure was 0.7164 V, and for the PCBM ETL-based structure, it was 0.6912 V (Fig. 8(b and c)), with the absorber thickness from 0.4 to 1.2 µm and 0.4 to 1.2 µm and the NA from 7 × 1017–7 × 1018 cm−3 for both cases. However, it should be noted that A similar response in SC structures with ETLs was observed when the absorber thickness was adjusted by varying the NA of WS2, C60, and PCBM.
image file: d5ra05366h-f8.tif
Fig. 8 Contour mapping of VOC showing the effects of varying in absorber thickness and NA for ETLs, including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f9.tif
Fig. 9 Contour mapping of JSC showing the effects of varying in absorber thickness and NA for ETLs including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f10.tif
Fig. 10 Contour mapping of FF showing the effects of varying in absorber thickness and NA for ETLs including (a) WS2, (b) C60, and (c) PCBM.

image file: d5ra05366h-f11.tif
Fig. 11 Contour mapping of PCE showing the effects of varying in absorber thickness and NA for ETLs including (a) WS2, (b) C60, and (c) PCBM.

Fig. 9 illustrates how the three enhanced SC structures under consideration's JSC values vary in response to adjustments in the absorber layer's thickness and NA. Out of these three solar configurations, the WS2 ETL-based device exhibits the highest JSC value, reaching 46.16 mA cm−2, when NA is approximately between a higher of 7.0 × 1014–7.0 × 1015 cm−3, and the thickness of the absorber falls between 1 to 1.2 µm Fig. 9(a). The lowest JSC values are seen in C60, an ETL-based solar structure, which is 44.98 mA cm−2 in cases when the absorber thickness ranges around 0.9 to 1.2 µm and the NA value is around 7 × 1016 to 7 × 1018 cm−3 (Fig. 9(b)). PCBM as ETL shows JSC value of 46.1 cm−3 during a thickness of 1–1.2 µm for the absorber, and the NA value is around 7 × 1014–7 × 1015 cm−3 (Fig. 9(c)). Remarkably, this shares a similarity with WS2 (ETL).

Fig. 10(a–c) illustrates the influence of absorber thickness and NA on FF. WS2 ETL-related PSC indicates 81.10% for FF when absorber thickness is (0.4–1.2) µm and NA is in the range from around 7 × 1015 to 7 × 1017 cm−3 on the basis of Fig. 10(a). C60 (ETL)-based device displays an FF of 76.00% when absorber thickness is 0.4–1.2 µm, which is the lowest. PCBM (ETL)-associated PSC displays an FF of 81.20% when NA varies between 7 × 1017–7 × 1018 cm−3. Moreover, with PCBM-based ETL, the greatest value of FF can be attained. The higher FF of PCBM-based PSCs compared to WS2-based PSCs is due to PCBM's superior electron mobility, better energy level alignment, and more favourable interface characteristics, which reduce recombination losses and enhance charge extraction.78 Additionally, WS2 has a higher defect density, which hampers charge transport.79 These combined factors contribute to the slightly higher FF observed in PCBM-based devices.

The effects of varying absorber thickness and NA on PCE for the three PSCs can be seen in Fig. 11(a–c). Fig. 11(a–c) illustrates that the highest PCE values for WS2, C60, and PCBM ETLs are 27%, 24.34%, and 25.35%, respectively, as the ENMO thickness is varied from 0.6 µm to 1.2 µm for WS2, C60, and PCBM. The NA falls between 7 × 1017–7 × 1018 cm−3. Among these three, WS2 (ETL) shows the maximum PCE and the C60 based ETL configuration shows the minimum PCE.

3.5. Influence of varying absorber and HTL layer thickness on PV performance

The performance of the device was improved by raising the absorber thickness from 400 nm to 1400 nm, since it affected the ITO/ETL (WS2, C60, PCBM)/ENMO/CFTS/Au structure's performance. Fig. 12(a) illustrates how the PSC's performance changes with varying absorber thickness for different ETLs. During the optimization process, higher reverse saturation current and absorber thickness caused a decrease in the PSC's VOC.80 The WS2-based ETL design exhibits the highest value of VOC compared to other configurations, which is ∼0.73 V, and the PCBM ETL-based device displays the smallest value of VOC is ∼0.68 V. In the case of JSC, all three structures followed the same pattern, while C60 ETL-based device revealed the lowest value ∼44.5 mA cm−2. However, the PCBM-based ETL structure shows a nearly linear decreasing pattern. The WS2 ETL-based PSC has the greatest FF value at 81%. In terms of PCE, each configuration shows the same scenario of increasing except PCBM ETL-associated structure. The maximum efficiency is ∼26.3% displayed by the WS2 ETL-based structure at 0.8 µm and the lowest value is ∼22% displayed by the C60-based structure. Thicker absorber layers enhance carrier recombination, while very thin layers are less efficient at generating carriers, which reduces the overall device efficiency.81 To improve VOC (∼0.72 V), JSC (∼45.5 mA cm2), FF (81%), and PCE (∼26.3%), the optimal thickness for the Eu2NiMnO6 absorber was determined to be 0.8 µm in the investigation.
image file: d5ra05366h-f12.tif
Fig. 12 Impact of varying (a) absorber thickness and (b) HTL thickness on PV parameters.

The effect of varying CFTS HTL thickness on PV parameters is shown in Fig. 12(b). CFTS is exclusively considered the HTL in thickness optimizations due to its highest PCE, with the effect of increasing CFTS thickness shown in Fig. 12(b), suggesting that the values of PCE, FF, JSC, and VOC for every ETL stayed constant. The optimal HTL thickness of 0.1 µm results from a balance between efficient hole extraction, minimal series resistance, and reduced recombination at the absorber/HTL interface. When the HTL is thinner than 0.1 µm, it may not fully cover the absorber surface, which can create incomplete contact and pathways for interfacial recombination, thereby reducing VOC and FF.66,82 On the other hand, when the HTL is thicker than 0.1 µm, holes must travel longer distances through the transport layer. This increases series resistance and reduces carrier mobility, which limits charge extraction and lowers JSC and overall PCE.71,83 In addition, an overly thick HTL can introduce additional trap states at the interface and increase the probability of recombination before carriers reach the electrode.84 Therefore, at 0.1 µm, the HTL is thick enough to ensure complete coverage and good band alignment with the absorber, but still thin enough to minimize transport losses, giving the best trade-off in photovoltaic performance (VOC ≈ 0.72 V, JSC ≈ 45.3 mA cm−2, FF ≈ 81%, and PCE ≈ 26.5%) for ITO/WS2/ENMO/CFTS/Au structure. The VOC value stays constant at around 0.72 V for WS2, 0.70 V for C60, and PCBM as ETLs at nearly 0.68 V for the increased thickness of CFTS. During the thickness of the CFTS improved the JSC value of C60 indicated a lower value of 44.5 mA cm−2, while WS2 displayed a higher value of 45.3 mA cm−2. Out of all configurations, the WS2-based structure achieves the best FF and PCE values, at about 81.1% and 26.5%, respectively. The C60 ETL-associated cell provides the smallest PCE and FF value with enhanced CFTS thickness, clocking in at around 23.5% and 75%, respectively. In the earlier study, it was noted that when the HTL thickness grew, the PCE value increased as well.85 Regarding the change, It was found that a thickness of 0.1 µm for the HTL was optimal for achieving higher PCE, so 0.1 µm was selected to be the optimized thickness of the HTL for further examination, which was also aligned with the earlier study.86

3.6. Influence of temperature, shunt, and series resistance on Eu2NiMnO6

3.6.1 Effects of series resistance. The right and left side metal contacts, connections among the layers of the solar cell, and manufacturing flaws are the main sources of the series (Rs) and shunt (Rsh) resistances, which strongly influence the efficiency of solar cells.53 The shunt resistance did not change from 105 Ω cm2, the influence of Rs changed from 0 and 6 Ω cm2, as indicated in Fig. 13(a) regarding the three (ITO/ETL/ENMO/CFTS/Au) structures. The declared figure shows that the PCE was decreasing for all three structures with a fluctuation of Rs. For the WS2 ETL-based structure, the PCE value fell from about 26% to 17.5%. On the other hand, the PCE of structures with C60 and PCBM ETLs decreased from about 23% to 15%, similar to another study of double perovskite SCs.86 It has been seen that the PCE value of C60 and PCBM ETL-based solar cells decreased similarly. For each of the three structures, the value of RS also had an impact on the FF value. The FF value of WS2 ETL-associated solar device decreased from around 82% to 50%. In contrast, the FF value of C60 and PCBM ETL-based devices decreased from around 75% to 48% and from 76 to 46%. The fill factor (FF) declined consistently due to the enhanced series resistance.87 Consequently, throughout the device's manufacture, RS must be reduced to a minimum to maximize performance and optimize FF. The JSC value of WS2 and PCBM ETL-based designs was slightly decreased with the variation of RS, which was around 45.3 to 45.2 mA cm−2 for WS2 ETL-based cells and around 45.2 to 45 mA cm−2 for PCBM-based cells. However, regarding the C60-based structure, the JSC value decreased gradually from 44.5 to 43.8 mA cm−2 with the variation of RS. The VOC value remained constant for all three structures with the variation of series resistance, demonstrating no impact on the VOC of RS for all three studied configurations. It is also seen in previous double Perovskite-based studies.
image file: d5ra05366h-f13.tif
Fig. 13 Effects of (a) series resistance and (b) shunt resistance on PV parameters.
3.6.2 Effects of shunt resistance. The device shunt resistance (Rsh) is an essential internal electrical component that influences the efficiency of SCs. It considers current leakage across the donor–acceptor and active layer-electrode boundaries.86 In our study, Fig. 13(b) represents the effects of Rsh in the case of three separate ETL-based SC structures. In Fig. 13(b), Rsh varied in ranges of 101 to 107 Ω cm2 for all three configurations. As Rsh climbed, the VOC, PCE, and FF readings all displayed a similar pattern except for JSC. It is also seen in previous studies.88 It was noticed that the value of VOC, PCE, and FF instantly increased in the range of 101 to 102 Ω cm2 of Rsh value. In the case of VOC, the WS2 ETL-associated solar configuration displayed a maximum of ∼0.71 V at an Rsh value around 102 Ω cm2 and remained constant after 103 Ω cm2. In comparison with the other PSCs, the Structures based on PCBM-ETL provided a minimum voltage of close to 0.635 V (Fig. 13(b)). The JSC for all three configurations is about the same, with the C60 ETL-based PSC displaying a minimum of around 44.5 mA cm−2 and the WS2 ETL-associated solar structure showing a maximum of around 45.25 mA cm−2. Among all the configurations, the FF of the WS2 ETL-based device PSC possessed the greatest at ∼81% and the C60 ETL-associated solar cell indicated the smallest value of ∼70%. In the case of PCE, the WS2 ETL-based PSC illustrated the greatest value of ∼25% and the remaining two PSCs C60 and PCBM, ETL-based structure indicated almost a similar value of ∼20%. Because of the fluctuation in Rsh, a pattern of variation was seen with the various PV parameters, which agreed with the results of the earlier investigation.89 To achieve optimal performance for the device, it is crucial to minimize the series resistance and maximize the shunt resistance.90
3.6.3 Effects of temperature. An increase in temperature from (275–320) K in our investigation is shown in Fig. 14 as an effect on the device's performance characteristics. The temperature impacts for three distinct PSC setups are illustrated in the preceding Fig. 14. At the time of changing temperature, we found variations in VOC, PCE, and FF for all three configurations. In the case of PCE, all three configurations showed similar trends of declining efficiency considering the rising temperature, where ITO/WS2/ENMO/CFTS/Au PSC indicated the highest value of ∼27% and C60 ETL-associated solar device showed the lowest value of ∼23.8%. The FF of the WS2 ETL-based solar configuration increased with the increase in temperature, where the largest value is ∼81%. The FF of the C60 ETL-associated solar device is also in an equivalent shape. On the other hand, the FF of PCBM ETL-associated solar configuration decreased with the increase in temperature. JSC stays constant regardless of temperature changes in all three configurations. Which states that there is no impact of temperature on the JSC of the PSCs of our study. All three structures showed a similar trend in VOC, decreasing with higher temperatures. As the temperature increases, VOC decreases due to bandgap narrowing and increased recombination.91 While JSC shows minor changes due to a balance between enhanced carrier generation and reduced mobility. The WS2 ETL-based PSC indicated a maximum value of ∼0.735 V, and the PCBM ETL-associated solar configuration displayed the smallest value of ∼0.676 V. Moreover, rising temperatures have an impact on diffusion length and raise RS, which have an immediate impact on the device's FF and PCE.92,93
image file: d5ra05366h-f14.tif
Fig. 14 Effects of temperature on (a) VOC, (b) JSC, (c) FF, and (d) PCE for (ITO/ETL/Eu2NiMnO6/CFTS/Au) double PSCs using ETLs (WS2, C60, PCBM).

3.7. Influence of capacitance and Mott–Schottky

The capacitance per unit area (C) displayed with Mott–Schottky (MS) and bias voltage (V), respectively, for three distinct configurations are presented in Fig. 15(a) and (b). In both instances shown in Fig. 15, the frequency stayed at 1 MHz, while the voltage ranged from −0.8 V to 0.8 V. For all configurations, capacitance stays zero as voltage ranges between −0.8 V and 0.4 V, but when the voltage fluctuated between ∼0.4 to 0.8 V, all three PSCs showed an exponential increase, while the WS2 ETL-associated solar configuration exhibited a late increase. The PSC with PCBM ETL demonstrated the peak capacitance of about 16[thin space (1/6-em)]000 nF cm−2, whereas the PSC with WS2 ETL had the lowest capacitance at approximately 4000 nF cm−2. Earlier research shows that the current is considerably less than the saturation current at low voltages and only reaches the saturation current at the peaks of voltage at the contact.94
image file: d5ra05366h-f15.tif
Fig. 15 Variation of (a) capacitance (b) Mott–Schottky (c) generation and (d) recombination for Eu2NiMnO6.

Conversely, the built-in potential (Vbi) of a device reflects the difference in performance between the electrodes and the degree of doping, may be found using MS, a well-used and trustworthy technique.80 Fig. 15(b) of our investigation showed an almost exact reversal of the preceding Figure behavior, with each of the three PSCs exhibiting a linear drop while the voltage varied between −0.8 and 0.4 V and all three remaining constants when the voltage ranged between ∼0.4 and 0.8 V, when the value was zero. Here the C60 ETL-associated structure indicated the largest MS value around 0.005 1/C2 and the WS2 ETL-associated structure revealed the smallest MS value around 0.0003 1/C2.

3.8. Effects of generation rate and recombination rate

Fig. 15(c) and (d) provide the graphs illustrating the rates of generation and recombination for three distinct structures. As carriers are produced, an electron shifts to the conduction band, creating an electron–hole pair.95 In Fig. 15(c), all three configurations show peak generation rates at about 0.8–0.9 µm. The computation of the electron–hole pair production, denoted as G(x), is performed utilizing SCAPS-1D and the incoming photon flux, Nphot (λ, x), according to eqn 11:
 
G(λ, x) = α(λ, x) × Nphot(λ, x) (11)

The reverse of generation, known as recombination, is the coupling and annihilation of conduction band electrons and holes.95 There is an impact on the defect state of every layer in the recombination process. After that, the energy state is constructed, which has a great impact on the recombination process. Defects at interfaces and grain boundaries cause uneven recombination rates in PSCs.86 Fig. 15(d) shows a slower start to recombination, with a peak at 0.9–1.0 µm in the C60 ETL structure. The convexity observed in the C60 and PCBM curves between 0.9–1.0 µm occurs due to the higher electron mobility of these materials, which results in increased charge buildup at the interface, leading to enhanced recombination.96,97 In contrast, the WS2 curve does not exhibit this convexity, as WS2 demonstrates more uniform charge transport with reduced recombination effects, resulting in smoother behavior.98 In the time range of 0.1–0.8 µm, the C60 and PCBM ETL-based structures showed almost similar recombination rates, at that time the WS2 ETL-based PSC showed a slightly lower recombination rate. But, within the bounds of 1.0–1.2 µm, the recombination rates are almost zero for all three configurations.

3.9. JV and QE properties of Eu2NiMnO6

Fig. 16(a) shows the JV curve for an ITO/ETL/ENMO/CFTS device structure with three distinct ETLs. In this case, the voltage varies between 0-0.8 V. In the beginning, all three configurations exhibit almost similar photocurrent. The process is continuous in the range of around 0.0–0.6 V for every structure, after that, the photocurrent of all PSCs start to decrease in the period of ∼0.6–0.72 V. Initially, the photocurrent of all three PSCs is nearly 45 mA cm−2. Moreover, the WS2 ETL-associated structure showed a good photocurrent in the presented Fig. 16(a), and the C60 ETL-based structure demonstrated a slightly reduced photocurrent as the voltage changed. The superior performance of WS2 ETL-based devices in terms of photocurrent is due to better energy level alignment, higher charge mobility.99 Conversely, the slightly reduced photocurrent with C60 is attributed to less optimal energy alignment, lower charge mobility, and possibly higher recombination rates.
image file: d5ra05366h-f16.tif
Fig. 16 (a) JV curve and (b) QE curve optimization for Eu2NiMnO6.

The plots of QE for every device under study are displayed in Fig. 16(b). The wavelengths range from 300 to 1300 nm in this case. In the plot, we observed an exponential increase for all configurations in the wavelength of 300–400 nm. It remains constant from ∼400–1000 nm, which is a long period. It can demonstrate that during that period, there is no impact of wavelength on the QE of studied PSCs. After that, the QE of all configurations starts to decrease with the variation of wavelength. In figure C60 the C60 ETL-based structure revealed a minor reduction in QE relative to other structures. However, the WS2 and PCBM ETL-based configurations have displayed almost a similar kind of efficiency with the variation of wavelength. The reduced photon absorption in C60 might be the cause of the decreased QE.88

3.10. Effect of interface defect density

Interface defects play a crucial role in charge transport and overall device performance. At the ETL/ENMO interface, defect states can trap photogenerated electrons and enhance Shockley–Read–Hall (SRH) recombination, while at the ENMO/HTL interface, hole trapping similarly increases interfacial recombination, which negatively affects the VOC and FF.100 More defects at the interface lead to higher recombination and lower charge transport efficiency. Energy level alignment is also critical. A small positive conduction band offset (CBO) at the ETL/ENMO interface promotes efficient electron transfer, whereas a negative CBO encourages interface-assisted recombination.101 Similarly, a slight positive valence band offset (VBO) at the ENMO/HTL interface enables hole extraction and suppresses electron leakage, while misalignment magnifies recombination losses.102 In summary, minimizing interfacial defects, controlling defect density, and ensuring favorable energy alignment at both junctions are essential for efficient carrier transport and enhanced device performance.

Fig. 17(a) and (b) illustrate the influence of defect density (Nt) on the effects of the ETL/Eu2NiMnO6 and HTL/Eu2NiMnO6 interface on multiple photovoltaic parameters, including VOC, FF, JSC, and PCE, within the range of 1010 to 1018 cm−2. According to the figure, recombination rates increase with rising Nt, reducing PCE, and ultimately leading to a fall in the performance parameters of PSCs. In this case, the performance parameters VOC, FF, and PCE exhibit a declining trend as the defect density increases. The JSC value remains almost constant for C60 and PCBM-based structures. In terms of the WS2-based ETL device, the JSC value remained almost unchanged between defect density values of 1010 and 1016 cm−2. Then, it displays a declining nature. The VOC drops significantly from around 0.74 V to 0.22 V, the JSC drops from nearly 45.3 to 43.81 mA cm−2, and the FF decreases from around 81 to 64% for an ETL structure based on WS2. As a result, the PCE reduces from approximately 27% to 5%, leading it the greatest among these three ETLs. To get the best performance, maintaining a defect density at 1010 cm−2 is crucial, which has been identified as the appropriate level for further investigation.


image file: d5ra05366h-f17.tif
Fig. 17 Influence of interface defects between (a) ETL/Eu2NiMnO6 and (b) HTL/Eu2NiMnO6 on the VOC, JSC, FF, and PCE parameters.

Eqn (12) defines the limit of interface recombination for the open-circuit voltage (VOC).103

 
image file: d5ra05366h-t7.tif(12)
In the above formula, A defines the ideality factor, c is the effective barrier height, and St specifies the recombination velocity at the interface.

Hole transport layers (HTLs) greatly influence photovoltaic (PV) device performance, with defect density serving as a crucial parameter. High defect concentrations (Nt) can restrict charge transfer, increase recombination events, and reduce the device's mechanical stability. In addition, inconsistent defects may alter the optical characteristics of the HTL, affecting absorption and uniformity.104 Ensuring low and uniform defect densities in HTLs is vital for achieving high efficiency and long-term durability. Fig. 17(b) presents the effect of interface defect density (Nt) on VOC​, FF, JSC​, and PCE for defect values between 1010-1018 cm−2. The Fig. 17 (b) shows that the HTL(CFTS)/Eu2NiMnO6 solar cell reaches its highest PCE of ∼25.2% at Nt = 1010 cm−2, but efficiency decreases as defect density increases due to enhanced recombination losses. VOC​ drops steadily from ∼0.74 V to ∼0.28 V with higher Nt, reflecting stronger non-radiative recombination. JSC​ remains nearly constant (∼45.6 mA cm−2) up to 1016 cm−2 and then decreases at higher defect densities, while FF stays stable around 78% at low Nt but declines significantly beyond 1014 cm−2.

3.11. Impedance effects on various optimized devices

The impedance, or Nyquist plot, of a solar cell allows for qualitative analysis of resistive losses, capacitance, and recombination issues.105 The Nyquist plot shown in Fig. 18 provides a comprehensive understanding of the impedance characteristics of PSCs employing different ETL materials. The geometrical capacitance of the SC is depicted on the Y-axis, which indicates the buildup of carriers at the interface layers. Resistance arising from recombination is shown on the X-axis. It is apparent from the graph that the diameter of the semicircle varies for each ETL-based device. The WS2 ETL-based structure's enlarged semicircle indicates a higher system impedance around 2200 ohm cm2. The impedance of the WS2 ETL-based structure is much greater than that of the other ETL-based designs. The C60 ETL-based device had the smallest impedance, observing around 270 ohm cm2. High-frequency measurements of resistance indicate the material's recombination resistance. The capacitance at these frequencies reflects the value of geometric capacitance, indicating that charge accumulates at the interfaces.106 Considering hysteresis and ionic mobility, the low-frequency response is more suspicious.107 A complete analysis of the impedance properties for PSCs is provided by the Nyquist plot. It elucidates the influence of various materials on ETL in terms of capacitance, resistive losses, and recombination rates. This comprehension is critical to the optimal and steady operation of solar cell devices.
image file: d5ra05366h-f18.tif
Fig. 18 Comparison of Nyquist plots for Eu2NiMnO6 absorbers with different ETL materials (WS2, C60, PCBM).

3.12. SCAPS-1D results compared to previous work

Table 4 compares photovoltaic parameters of previous solar cells with the same absorber to optimized ENMO-based PSCs. The previously published structure is FTO/TiO2/ENMO/CuI/Au, with the efficiency of these theoretical results, is 9.41%62 and 12.63%.108 The calculated PCE for the solar structures presented for ITO/WS2/ENMO/CFTS/Au, ITO/C60/ENMO/CFTS/Au and ITO/PCBM/ENMO/CFTS/Au is 26.45, 23.43, and 23.74%, which exceeds the theoretical results reported in earlier studies. The main reason for the difference is the careful selection of ETL and HTL contributed to higher JSC values in our devices. We investigated absorber characteristics, such as thickness, which vary from those in previous theoretical studies of device structures. Furthermore, the combinations of ETL and HTL we explored differ from those studied in earlier theoretical research. Additionally, the optical properties vary between different absorbers, leading to differences in solar energy absorption. The ENMO-based PSC design exhibits FF values comparable to those found in earlier research.
Table 4 Theoretical analysis of the ENMO absorber layer
Optimized devices VOC (V) JSC (mA cm−2) FF (%) PCE (%) Ref.
FTO/TiO2/ENMO/CuI/Au 0.772 16.43 74.16 9.41 62
FTO/TiO2/ENMO/CuI/Au 0.78 21.5 75.33 12.63 108
ITO/WS2/ENMO/CFTS/Au 0.720 45.2872 81.02 26.45 This study
ITO/C60/ENMO/CFTS/Au 0.703 44.551 74.82 23.43 This study
ITO/PCBM/ENMO/CFTS/Au 0.683 45.16 76.94 23.74 This study


4 Conclusion

The primary purpose of this research is to explore the double perovskite Eu2NiMnO6's (ENMO) ability to develop useful photovoltaic applications utilizing the SCAPS-1D tool findings. It is discovered that the three solar configurations-ITO/WS2/ENMO/CFTS/Au, ITO/C60/ENMO/CFTS/Au, and ITO/PCBM/ENMO/CFTS/Au-are the best SC configurations in terms of PV characteristics. The WS2-based structure exhibited the highest performance with a VOC of 0.7208 V, a JSC of 45.78 mA cm−2, an FF of 81.02%, and a PCE of 26.45%. This study examined the effects of various PV parameters while varying absorber thickness (0.4 to 1.4 µm). From 0.4–0.8, µm the curve is inclined, and the best efficiency we got for the absorber thickness is 0.8 µm, for all the configurations, and after the rate of efficiency is almost constant. Then HTL is observed at (0.1 to 0.5 µm), where we notice that for all device configurations with a variation of HTL thickness, efficiency is nearly constant. The impact of variations in acceptor density (NA) was also investigated in this study. Acceptor density from 7 × 1014 cm−3 to 7 × 1018 cm−3 revealed insights into the performance variations. Since shunt resistance enhances VOC, FF, PCE, and JSC have a steady, negligible influence, and series resistance decreases PCE, FF, JSC, and VOC almost constantly. There is a considerable effect of temperature for all three configurations. The devices linked with PCBM and C60 ETLs exhibited the maximum rates of generation and recombination at (0.8–0.9) µm and (0.9–1) µm, respectively. Compared to other ETL-associated devices, the WS2 ETL device's appropriate band alignment resulted in superior JV and QE characteristics. These results have great significance for researchers exploring double perovskite-based PSCs since they allow for the creation of ideal SC configurations before the manufacturing and testing of these devices.

Author contributions

Md. Abu Bakkar Siddique: Investigation, methodology, data curation, conceptualization, writing original manuscript; Nazmul Shahadath: Investigation, methodology, data curation, review-editing; Md. Tarekuzzaman: Formal analysis, software, conceptualization, review-editing; Md. Raihan Kabir: Formal analysis, methodology, data curation, review-editing; Sohail Ahmad: Data curation, validation, Formal analysis; Rashel Mohammad Khokan: Formal analysis, validation, supervision, review-editing. Md. Rasheduzzaman: Formal analysis, validation, review-editing; S. M. G Mostafa: Formal analysis, validation, review-editing; Mohammad Jalal Uddin Formal analysis, validation, review-editing; Md. Zahid Hasan: Formal analysis, validation, supervision, review-editing.

Conflicts of interest

There is no conflict to declare.

Abbreviations

PSCPerovskite solar cell
PCEPower conversion efficiency
JSCShort-circuit current density
VOCOpen-circuit voltage
FFFill factor
ITOIndium tin oxide
PVPhotovoltaic
JVCurrent density–voltage
WS2Tungsten disulfide
C60Buckminsterfullerene
EAElectron affinity
εrDielectric permittivity (relative)
NCCB effective density of states
NDShallow uniform donor density
NtDefect density
PCBMPhenyl-C61-butyric acid methyl ester
CBTSCopper barium tin sulfide
MSMott–Schottky
QEQuantum efficiency
WFWork function
Fn/pFermi level of the electron/hole
AuGold
EC/EVEnergy level of the conduction/valence band
AlAluminium
SCSolar cell
NVVB effective density of states
µnElectron mobility
µhHole mobility
NAShallow uniform acceptor density
CFTSCopper ferrite tin sulfide

Data availability

Data will be made available on request. The SCAPS-1D program was graciously provided by Dr M. Burgelman of the University of Gent in Belgium, for which the authors extend their sincere gratitude.

Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05366h.

Acknowledgements

The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through a Large Research Project under grant number RGP2/614/46.

References

  1. M. Asif and T. Muneer, Renew. Sustain. Energy Rev., 2007, 11, 1388–1413 CrossRef.
  2. M. H. K. Rubel, S. K. Mitro, M. K. Hossain, K. M. Hossain, M. M. Rahaman, J. Hossain, B. K. Mondal, A. Akter, Md. F. Rahman, I. Ahmed and A. K. M. A. Islam, Mater. Today Commun., 2022, 33, 104302 CrossRef CAS.
  3. S. Saurabh, M. K. Hossain, S. Singh, S. K. Agnihotri and D. P. Samajdar, RSC Adv., 2023, 13, 9878–9891 RSC.
  4. H. Al-Dmour, R. H. Alzard, H. Alblooshi, K. Alhosani, S. AlMadhoob and N. Saleh, Front. Chem., 2019, 7, 561 CrossRef CAS PubMed.
  5. H. Al-Dmour and D. M. Taylor, Thin Solid Films, 2011, 519, 8135–8138 CrossRef CAS.
  6. N. Shahadath, Md. A. B. Siddique, Md. Tarekuzzaman, M. H. Ishraq, Md. F. Rahman, A. M. Arfi, Md. Rasheduzzaman, S. M. G. Mostafa and Md. Z. Hasan, New J. Chem., 2024, 48, 14336–14353 RSC.
  7. E. Kabir, P. Kumar, S. Kumar, A. A. Adelodun and K.-H. Kim, Renew. Sustain. Energy Rev., 2018, 82, 894–900 CrossRef.
  8. E. Oublal, M. Al-Hattab, A. Ait Abdelkadir and M. Sahal, Sol. Energy, 2022, 246, 57–65 CrossRef CAS.
  9. H.-H. Huang, Y.-C. Shih, L. Wang and K.-F. Lin, Energy Environ. Sci., 2019, 12, 1265–1273 RSC.
  10. Md. A. Bakkar Siddique, Md. S. Parves, Md. Tarekuzzaman, Md. R. Kabir, M. S. M. Al-Saleem, J. Y. Al-Humaidi, Md. Rasheduzzaman, M. M. Hossen, M. M. Rahman and Md. Z. Hasan, Langmuir, 2025, 41, 19797–19820 CrossRef CAS PubMed.
  11. X. Wang, T. Zhang, Y. Lou and Y. Zhao, Mater. Chem. Front., 2019, 3, 365–375 RSC.
  12. A. Miyata, A. Mitioglu, P. Plochocka, O. Portugall, J. T. Wang, S. D. Stranks, H. J. Snaith and R. J. Nicholas, Nat. Phys., 2015, 11(7), 582–587 Search PubMed.
  13. J. Huang, Y. Yuan, Y. Shao and Y. Yan, Nat. Rev. Mater., 2017, 2(7), 1–9 Search PubMed.
  14. M. A. Madanat, A. A. Al-Tabbakh, M. Alsa’eed, H. Al-Dmour and M. S. Mousa, Ultramicroscopy, 2022, 234, 113479 CrossRef CAS PubMed.
  15. J. J. Yoo, G. Seo, M. R. Chua, T. G. Park, Y. Lu, F. Rotermund, Y. K. Kim, C. S. Moon, N. J. Jeon, J. P. Correa-Baena and V. Bulović, Nature, 2021, 590(7847), 587–593 CrossRef CAS PubMed.
  16. J. Jeong, M. Kim, J. Seo, H. Lu, P. Ahlawat, A. Mishra, Y. Yang, M. A. Hope, F. T. Eickemeyer, M. Kim, Y. J. Yoon, I. W. Choi, B. P. Darwich, S. J. Choi, Y. Jo, J. H. Lee, B. Walker, S. M. Zakeeruddin, L. Emsley, U. Rothlisberger, A. Hagfeldt, D. S. Kim, M. Grätzel and J. Y. Kim, Nature, 2021, 592, 381–385 CrossRef CAS PubMed.
  17. T. J. Macdonald, M. Batmunkh, C.-T. Lin, J. Kim, D. D. Tune, F. Ambroz, X. Li, S. Xu, C. Sol, I. Papakonstantinou, M. A. McLachlan, I. P. Parkin, J. G. Shapter and J. R. Durrant, Small Methods, 2019, 3, 1900164 CrossRef CAS.
  18. M. Saliba, J.-P. Correa-Baena, C. M. Wolff, M. Stolterfoht, N. Phung, S. Albrecht, D. Neher and A. Abate, Chem. Mater., 2018, 30, 4193–4201 CrossRef CAS.
  19. A. M. Alanazi, H. Al Dmour, S. A. Popoola, H. Oudghiri Hassani, S. Rakass, R. Al-Faze and F. Kooli, Inorganics, 2023, 11, 423 CrossRef CAS.
  20. M. R. Kabir, M. H. Ishraq, M. Tarekuzzaman, N. I. Nahid, J. K. Modak, S. Ahmad, A. M. Arfi, M. Rasheduzzaman and M. Z. Hasan, Eng. Res. Express, 2025, 7, 015319 CrossRef.
  21. T. I. Alanazi, O. S. Game, J. A. Smith, R. C. Kilbride, C. Greenland, R. Jayaprakash, K. Georgiou, N. J. Terrill and D. G. Lidzey, RSC Adv., 2020, 10, 40341–40350 RSC.
  22. S. Abdelaziz, A. Zekry, A. Shaker and M. Abouelatta, Opt. Mater., 2022, 123, 111893 CrossRef CAS.
  23. Z. T. M. Noori, A. M. Naji, O. A. Nief, M. K. A. Mohammed, N. M. Ahmed and S. Singh, Int. J. Energy Res., 2022, 46, 17285–17294 CrossRef CAS.
  24. S. Al-Taweel, S. Al-Trawneh, H. Al-Dmour, O. Al-Gzawat, W. Alhalasah and M. Mousa, Heliyon, 2023, 9, e21039 CrossRef CAS PubMed.
  25. Md. R. Kabir, N. Shahadath, Md. Tarekuzzaman, Md. A. B. Siddique, O. Alsalmi, Md. Rasheduzzaman, M. A. Qader, M. M. Hossen and Md. Z. Hasan, RSC Adv., 2025, 15, 17906–17932 RSC.
  26. T. He, Y. Jiang, X. Xing and M. Yuan, Adv. Mater., 2020, 32(26), 1903937 CrossRef CAS PubMed.
  27. A. N. Singh, S. Kajal, J. Kim, A. Jana, J. Y. Kim and K. S. Kim, Adv. Energy Mater., 2020, 10, 2070129 CrossRef CAS.
  28. J. Xu, R. Saklatvala, S. Mittal, S. Deshmukh and A. Procopio, Adv. Sci., 2020, 7, 1903394 CrossRef CAS PubMed.
  29. S. Kashyap, J. Madan, M. K. A. Mohammed, M. Khalid Hossain, S. Ponnusamy and R. Pandey, Mater. Lett., 2023, 339, 134096 CrossRef CAS.
  30. M. K. A. Mohammed, A. K. Al-Mousoi, S. Singh, A. Kumar, M. K. Hossain, S. Q. Salih, P. Sasikumar, R. Pandey, A. A. Yadav and Z. M. Yaseen, Opt. Mater., 2023, 138, 113702 CrossRef.
  31. R. Pandey, S. Bhattarai, K. Sharma, J. Madan, A. K. Al-Mousoi, M. K. Mohammed and M. K. Hossain, ACS Appl. Electron. Mater., 2023, 5(10), 5303–5315 CrossRef CAS.
  32. M. K. Hossain, A. A. Arnab, D. P. Samajdar, M. H. K. Rubel, M. M. Hossain, Md. R. Islam, R. C. Das, H. Bencherif, Md. F. Rahman, J. Madan, R. Pandey, S. Bhattarai, M. Amami and D. K. Dwivedi, Energy Fuels, 2023, 37, 13377–13396 CrossRef CAS.
  33. H. J. Snaith, J. Phys. Chem. Lett., 2013, 4, 3623–3630 CrossRef CAS.
  34. M. Faizan, G. Murtaza, S. H. Khan, A. Khan, A. Mehmood, R. Khenata and S. Hussain, Bull. Mater. Sci., 2016, 39, 1419–1425 CrossRef CAS.
  35. A. Schmitz, L. L. Schaberg, S. Sirotinskaya, M. Pantaler, D. C. Lupascu, N. Benson and G. Bacher, ACS Energy Lett., 2020, 5, 559–565 CrossRef CAS.
  36. Md. Sariful Sheikh, D. Ghosh, A. Dutta, S. Bhattacharyya and T. P. Sinha, Mater. Sci. Eng. B, 2017, 226, 10–17 CrossRef CAS.
  37. M. K. Hossain, G. F. I. Toki, A. Kuddus, M. H. K. Rubel, M. M. Hossain, H. Bencherif, Md. F. Rahman, Md. R. Islam and M. Mushtaq, Sci. Rep., 2023, 13, 2521 CrossRef CAS PubMed.
  38. Y. Raoui, H. Ez-Zahraouy, N. Tahiri, O. El Bounagui, S. Ahmad and S. Kazim, Sol. Energy, 2019, 193, 948–955 CrossRef CAS.
  39. N. Singh, A. Agarwal and M. Agarwal, Opt. Mater., 2021, 114, 110964 CrossRef CAS.
  40. K. Sekar, L. Marasamy, S. Mayarambakam, H. Hawashin, M. Nour and J. Bouclé, RSC Adv., 2023, 13, 25483–25496 RSC.
  41. M. Elhamel, Z. Hebboul, M. E. Naidjate, A. Draoui, A. Benghia, M. abdelilah Fadla, M. B. Kanoun and S. Goumri-Said, J. Solid State Chem., 2023, 323, 124022 CrossRef CAS.
  42. H. M. Christen and G. Eres, J. Phys. Condens. Matter, 2008, 20, 264005 CrossRef CAS PubMed.
  43. S. M. AL-Shomar, Mater. Res. Express, 2021, 8, 026402 CrossRef CAS.
  44. A. K. Al-Mousoi, M. K. Mohammed, A. Kumar, R. Pandey, J. Madan, D. Dastan, M. K. Hossain, P. Sakthivel and Z. M. Yaseen, Phys. Chem. Chem. Phys., 2023, 25(24), 16459–16468 RSC.
  45. M. F. Rahman, N. Mahmud, I. Alam, M. H. Ali, M. M. Moon, A. Kuddus, G. F. Toki, M. H. Rubel, M. A. Al Asad and M. K. Hossain, AIP Adv., 2023, 13(4), 045309 CrossRef CAS.
  46. Md. Hasan Ali, A. T. M. Saiful Islam, M. D. Haque, Md. Ferdous Rahman, M. Khalid Hossain, N. Sultana and A. Z. M. Touhidul Islam, Mater. Today Commun., 2023, 34, 105387 CrossRef CAS.
  47. A. Isha, A. Kowsar, A. Kuddus, M. K. Hossain, M. H. Ali, M. D. Haque and M. F. Rahman, Heliyon, 2023, 9, e15716 CrossRef CAS PubMed.
  48. M. H. Ali, M. A. Al Mamun, M. D. Haque, M. F. Rahman, M. K. Hossain and A. Z. Md Touhidul Islam, ACS Omega, 2023, 8(7), 7017–7029 CrossRef CAS PubMed.
  49. Md. F. Rahman, Md. M. Alam Moon, M. K. Hossain, Md. H. Ali, Md. D. Haque, A. Kuddus, J. Hossain and A. B. Md. Ismail, Heliyon, 2022, 8, e12034 CrossRef CAS PubMed.
  50. D. Dastan, M. K. Mohammed, A. K. Al-Mousoi, A. Kumar, S. Q. Salih, P. S. JosephNg, D. S. Ahmed, R. Pandey, Z. M. Yaseen and M. K. Hossain, Sci. Rep., 2023, 13(1), 9076 CrossRef CAS PubMed.
  51. M. Minbashi, A. Ghobadi, M. H. Ehsani, H. Rezagholipour Dizaji and N. Memarian, Sol. Energy, 2018, 176, 520–525 CrossRef CAS.
  52. S. M. Sze, Y. Li and K. K. Ng, Physics of Semiconductor Devices, John wiley & sons, 2021 Search PubMed.
  53. M. K. Hossain, M. H. Rubel, G. I. Toki, I. Alam, M. F. Rahman and H. Bencherif, ACS Omega, 2022, 7(47), 43210–43230 CrossRef CAS PubMed.
  54. S. Bhattarai and T. D. Das, Sol. Energy, 2021, 217, 200–207 CrossRef CAS.
  55. M. S. Sheikh, A. P. Sakhya, A. Dutta and T. P. Sinha, Comput. Mater. Sci., 2019, 161, 293–299 CrossRef CAS.
  56. D. C. Kakarla, K. M. Jyothinagaram, A. K. Das and V. Adyam, J. Am. Ceram. Soc., 2014, 97, 2858–2866 CrossRef CAS.
  57. M. Nasir, S. Kumar, N. Patra, D. Bhattacharya, S. N. Jha, D. R. Basaula, S. Bhatt, M. Khan, S. W. Liu, S. Biring and S. Sen, arXiv, 2018,  DOI:10.48550/ARXIV.1808.00170.
  58. C. Chen, X. Lu, X. Hu, G. Liang and G. Fang, J. Mater. Chem. C, 2024, 12, 16–28 RSC.
  59. H. Dixit, N. K. Bansal, S. Porwal, D. Kumar and T. Singh, Optik, 2023, 295, 171474 CrossRef CAS.
  60. A. Talukdar, P. Debnath, J. Sarkar and S. Chatterjee, Indian J. Phys., 2024, 98, 3913–3929 CrossRef CAS.
  61. M. A. Nalianya, C. Awino, H. Barasa, V. Odari, F. Gaitho, B. Omogo and M. Mageto, Optik, 2021, 248, 168060 CrossRef CAS.
  62. M. Kumar, A. Raj, A. Kumar and A. Anshul, Mater. Today Commun., 2021, 26, 101851 CrossRef CAS.
  63. M. H. Ishraq, Md. R. Kabir, Md. Tarekuzzaman, Md. F. Rahman, Md. Rasheduzzaman and Md. Z. Hasan, Adv. Theory Simul., 2024, 2400360 CrossRef CAS.
  64. C. Walkons, R. Murshed and S. Bansal, Sol. RRL, 2020, 4(10), 2000299 CrossRef CAS.
  65. T. Minemoto and M. Murata, Sol. Energy Mater. Sol. Cells, 2015, 133, 8–14 CrossRef CAS.
  66. W. Tress, Adv. Energy Mater., 2017, 7, 1602358 CrossRef.
  67. Y.-N. Zhang, B. Li, L. Fu, Y. Zou, Q. Li and L.-W. Yin, Sol. Energy Mater. Sol. Cells, 2019, 194, 168–176 CrossRef CAS.
  68. F. Xu, X. Yang, T. Huang, Z. Li, Y. Ji and R. Zhu, Nano Energy, 2024, 130, 110171 CrossRef CAS.
  69. M. A. Green, A. Ho-Baillie and H. J. Snaith, Nat. Photonics, 2014, 8, 506–514 CrossRef CAS.
  70. Y. Zhao and K. Zhu, J. Phys. Chem. Lett., 2013, 4, 2880–2884 CrossRef CAS.
  71. J.-F. Liao, W.-Q. Wu, Y. Jiang, J.-X. Zhong, L. Wang and D.-B. Kuang, Chem. Soc. Rev., 2020, 49, 354–381 RSC.
  72. H. Wang, H. Li, W. Cai, P. Zhang, S. Cao, Z. Chen and Z. Zang, Nanoscale, 2020, 12, 14369–14404 RSC.
  73. M. Khalid Hossain, G. F. Ishraque Toki, I. Alam, R. Pandey, D. P. Samajdar, M. Ferdous Rahman, M. Rasidul Islam, M. H. K. Rubel, H. Bencherif, J. Madan and M. K. A. Mohammed, New J. Chem., 2023, 47, 4801–4817 RSC.
  74. Z. Iqbal, F. Zu, A. Musiienko, E. Gutierrez-Partida, H. Köbler, T. W. Gries, G. V. Sannino, L. Canil, N. Koch, M. Stolterfoht, D. Neher, M. Pavone, A. B. Muñoz-García, A. Abate and Q. Wang, ACS Energy Lett., 2023, 8, 4304–4314 CrossRef CAS PubMed.
  75. J. Siekmann, A. Kulkarni, S. Akel, B. Klingebiel, M. Saliba, U. Rau and T. Kirchartz, Adv. Energy Mater., 2023, 13, 2300448 CrossRef CAS.
  76. E. I. Emon, A. M. Islam, M. K. Sobayel, S. Islam, M. Akhtaruzzaman, N. Amin, A. Ahmed and M. J. Rashid, Heliyon, 2023, 9, e14438 CrossRef CAS PubMed.
  77. M. A. Green, Solid-State Electron., 1981, 24, 788–789 CrossRef CAS.
  78. M. T. Dang, L. Hirsch and G. Wantz, Adv. Mater., 2011, 23, 3597–3602 CrossRef CAS PubMed.
  79. R. R. Scherberger, H. Kaess and S. Brückner, Arzneimittelforschung, 1975, 25, 1460–1463 CAS.
  80. I. Alam, R. Mollick and M. A. Ashraf, Phys. B, 2021, 618, 413187 CrossRef CAS.
  81. A. S, N. Abraham and V. Suresh Babu, Mater. Today: Proc., 2021, 43, 3432–3437 CAS.
  82. J. Xu, O. Voznyy, R. Comin, X. Gong, G. Walters, M. Liu, P. Kanjanaboos, X. Lan and E. H. Sargent, Adv. Mater., 2016, 28, 2807–2815 CrossRef CAS PubMed.
  83. O. Saidani, S. Goumri-Said, A. Yousfi, G. S. Sahoo and M. B. Kanoun, RSC Adv., 2025, 15, 7342–7353 RSC.
  84. N. Marinova, W. Tress, R. Humphry-Baker, M. I. Dar, V. Bojinov, S. M. Zakeeruddin, M. K. Nazeeruddin and M. Grätzel, ACS Nano, 2015, 9, 4200–4209 CrossRef CAS PubMed.
  85. A. Owolabi, M. Onimisi, J. Ukwenya, A. Bature, U. Ushiekpan and U. Ushiekpan, Am. J. Phys. Appl., 2020, 8, 8–18 Search PubMed.
  86. M. K. Hossain, D. P. Samajdar, R. C. Das, A. A. Arnab, Md. F. Rahman, M. H. K. Rubel, Md. R. Islam, H. Bencherif, R. Pandey, J. Madan and M. K. A. Mohammed, Energy Fuels, 2023, 37, 3957–3979 CrossRef CAS.
  87. L. Lin, L. Jiang, P. Li, B. Fan and Y. Qiu, J. Phys. Chem. Solids, 2019, 124, 205–211 CrossRef CAS.
  88. M. K. Hossain, G. F. Ishraque Toki, D. P. Samajdar, M. H. K. Rubel, M. Mushtaq, Md. R. Islam, Md. F. Rahman, S. Bhattarai, H. Bencherif, M. K. A. Mohammed, R. Pandey and J. Madan, Energy Fuels, 2023, 37, 7380–7400 CrossRef CAS.
  89. M. N. H. Riyad, A. Sunny, M. M. Khatun, S. Rahman and S. R. A. Ahmed, Eng. Rep., 2023, 5, e12600 CrossRef CAS.
  90. A. Sunny, S. Rahman, M. M. Khatun and S. R. A. Ahmed, AIP Adv., 2021, 11, 065102 CrossRef CAS.
  91. A. K. Thakur, G. Wantz, G. Garcia-Belmonte, J. Bisquert and L. Hirsch, Sol. Energy Mater. Sol. Cells, 2011, 95, 2131–2135 CrossRef CAS.
  92. F. Behrouznejad, S. Shahbazi, N. Taghavinia, H. P. Wu and E. W. Diau, J. Mater. Chem. A, 2016, 4(35), 13488–13498 RSC.
  93. S. R. Raga, E. M. Barea and F. Fabregat-Santiago, J. Phys. Chem. Lett., 2012, 3, 1629–1634 CrossRef CAS PubMed.
  94. M. Burgelman, K. Decock, A. Niemegeers, J. Verschraegen and S. Degrave, SCAPS manual, University of Ghent, Ghent, Belgium, 2016 Search PubMed.
  95. M. K. Hossain, A. A. Arnab, R. C. Das, K. M. Hossain, M. H. Rubel, M. F. Rahman, H. Bencherif, M. E. Emetere, M. K. Mohammed and R. Pandey, RSC Adv., 2022, 12(54), 35002–35025 RSC.
  96. V. M. Caselli and T. J. Savenije, J. Phys. Chem. Lett., 2022, 13, 7523–7531 CrossRef CAS PubMed.
  97. N. K. Elumalai, L. M. Yin, V. Chellappan, Z. Jie, Z. Peining and S. Ramakrishna, Phys. Status Solidi A, 2012, 209, 1592–1597 CrossRef CAS.
  98. L. Wang, M. Tahir, H. Chen and J. B. Sambur, Nano Lett., 2019, 19, 9084–9094 CrossRef CAS PubMed.
  99. M. Malik, M. Kashif, F. B. Sumona and M. U. N. Tariq, J. Opt., 2025, 27, 015901 CrossRef CAS.
  100. F. Izadi, A. Ghobadi, A. Gharaati, M. Minbashi and A. Hajjiah, Optik, 2021, 227, 166061 CrossRef CAS.
  101. Z. Zhao, Y. Liu, D. Wang, C. Ling, Q. Chang, J. Li, Y. Zhao and H. Jin, Sol. Energy Mater. Sol. Cells, 2020, 209, 110443 CrossRef CAS.
  102. C. Ding, Y. Zhang, F. Liu, Y. Kitabatake, S. Hayase, T. Toyoda, K. Yoshino, T. Minemoto, K. Katayama and Q. Shen, Nano Energy, 2018, 53, 17–26 CrossRef CAS.
  103. G. G. Malliaras, J. R. Salem, P. J. Brock and C. Scott, Phys. Rev. B:Condens. Matter Mater. Phys., 1998, 58, R13411–R13414 CrossRef CAS.
  104. M. Reza, A. Ghosh, M. Reza, S. N. Wornob and S. Sultana, Energy Nexus, 2025, 18, 100417 CrossRef CAS.
  105. S. Bhattarai, R. Pandey, J. Madan, S. Tayeng, P. K. Kalita, M. Z. Ansari, L. B. Farhat, M. Amami and M. K. Hossain, RSC Adv., 2023, 13(38), 26851–26860 RSC.
  106. E. von Hauff and D. Klotz, J. Mater. Chem. C, 2022, 10, 742–761 RSC.
  107. A. Guerrero, J. Bisquert and G. Garcia-Belmonte, Chem. Rev., 2021, 121, 14430–14484 CrossRef CAS PubMed.
  108. A. Raj, A. Anshul and M. Kumar, Mater. Today: Proc., 2022, 67, 748–751 CAS.

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