Next Article in Journal
The Role of Building Geometry in Urban Heat Islands: Case of Doha, Qatar
Previous Article in Journal
1D Finite Element Modeling of Bond-Slip Behavior and Deflection in Reinforced Concrete Flexural Members
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Performance of BaSnO3-Based Perovskite Solar Cell Designs Under Variable Light Intensities, Temperatures, and Donor and Defect Densities

by
Nouf Alkathran
1,2,*,
Shubhranshu Bhandari
1 and
Tapas K. Mallick
1,3,*
1
Environment and Sustainability Institute, University of Exeter, Penryn Campus, Cornwall TR10 9FE, UK
2
Department of Physics and Astronomy, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
3
Department of Mechanical and Energy Engineering, College of Engineering, Imam Abdulrahman Bin Faisal University, Dammam 34212, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Designs 2025, 9(3), 76; https://doi.org/10.3390/designs9030076
Submission received: 27 April 2025 / Revised: 31 May 2025 / Accepted: 10 June 2025 / Published: 18 June 2025

Abstract

Barium stannate (BaSnO3) has emerged as a promising alternative electron transport material owing to its superior electron mobility, resistance to UV degradation, and energy bandgap tunability, yet BaSnO3-based perovskite solar cells have not reached the efficiency levels of TiO2-based designs. This theoretical study presents a design-driven evaluation of BaSnO3-based perovskite solar cell architectures, incorporating MAPbI3 or FAMAPbI3 perovskite materials, Spiro-OMeTAD, or Cu2O hole transport materials as well as hole-free configurations, under varying light intensity. Using a systematic device modelling approach, we explore the influence of key design variables—such as layer thickness, donor density, and interface defect concentration—of BaSnO3 and operating temperature on the power conversion efficiency (PCE). Among the proposed designs, the FTO/BaSnO3/FAMAPbI3/Cu2O/Au heterostructure exhibits an exceptionally effective arrangement with PCE of 38.2% under concentrated light (10,000 W/m2, or 10 Sun). The structure also demonstrates strong thermal robustness up to 400 K, with a low temperature coefficient of −0.078% K−1. These results underscore the importance of material and structural optimisation in PSC design and highlight the role of high-mobility, thermally stable inorganic transport layers—BaSnO3 as the electron transport material (ETM) and Cu2O as the hole transport material (HTM)—in enabling efficient and stable photovoltaic performance under high irradiance. The study contributes valuable insights into the rational design of high-performance PSCs for emerging solar technologies.

Graphical Abstract

1. Introduction

Among all the existing photovoltaic technologies, perovskite solar cells (PSCs) are the brightest option. Throughout the last ten years, the power conversion efficiency of PSCs has dramatically improved from 3.5% [1] to 25.8% [2,3]. This notable advancement is primarily driven by the materials’ superior optical, electrical, and electronic characteristics. Moreover, the fabrication of PSCs is achievable through a simple and economical solution-processable approach [4,5,6]. The basic structure of PSCs involves an absorber layer (perovskite) sandwiched between two electrodes. Hybrid organic–inorganic halide-based perovskite materials are widely studied for use in PSCs, with lead-based compounds such as Methylammonium lead iodide (MAPbI3) demonstrating superior efficiency [6]. The electron transport layer (ETL) and hole transport layer (HTL) are commonly used as interfacial buffers between the absorber layer and the electrodes. These layers serve an essential function in enabling charge carrier separation and transport to electrodes while also mitigating charge recombination [6,7]. Despite this, the successful fabrication of HTL-free PSCs has been demonstrated, attributed to the inherent dual functionality of perovskite materials, which can serve as both a light absorber and a hole conductor [8,9,10].
Although there have been significant advancements in the field of PSCs, various obstacles still prevent them from being commercially viable, especially in terms of their long-term stability and overall performance [3]. An essential component in enhancing the efficiency of PSCs is the selection of electron transport material. While titanium dioxide (TiO2) remains the predominant ETL material [11], it exhibits certain drawbacks, such as low electron mobility, vulnerability to ultraviolet (UV) radiation-induced degradation, and the necessity for high-temperature processing, all of which can negatively impact the performance of PSCs [12,13,14]. These difficulties, including charge recombination at the interface and hysteresis effects, highlight the need for more effective electron transport materials (ETMs) [15]. Although combined ETL has been adopted to enhance the performance of TiO2-based PSCs, the issue of low electron mobility and poor stability upon UV exposure have not been overcome [16]. Other binary oxides such as tin oxide (SnO2) [17] and ternary metal oxides such as zinc stannate oxide (Zn2SnO4) [15] have been studied as alternative ETMs due to their potential advantages, including better charge transport properties, enhanced stability, compatibility with low temperatures, and scalable manufacturing processes. Consequently, exploring and developing more efficient ETMs is crucial to advancing PSCs’ performance, photostability, and commercial feasibility.
BaSnO3 (BSO) has garnered significant attention due to its promising performance as an n-type semiconductor [18,19]. BSO is an attractive ETM option because it shares similar electronic properties with TiO2, such as bandgap and charge transfer kinetics, while offering improved chemical stability and higher electron mobility [20]. Additionally, the optoelectronic properties of BSO can be readily modified through doping or by adjusting the relative ratios of cations, thereby allowing control over its bandgap energy, work function, and electrical resistivity [12]. Sun et al. introduced BSO as ETL in planar PSCs for the first time. By optimizing the BSO layer’s coverage, roughness, and thickness, BSO-based PSCs demonstrated nearly a 12% increase in power conversion efficiency (PCE) compared to TiO2-based cells. Such findings have been attributed to decreased charge transfer resistance, efficient electron transfer between the perovskite and BSO, and reduced charge recombination [20]. In 2021, A. Rehman [21] conducted a simulation study using SCAPS-1D to explore MAPbI3-based perovskite solar cells, incorporating Cu or Mo as the metal back electrode and BaSnO3 and Cu2O as the ETL and HTL, respectively. Following critical optimisations of device parameters such as layer thickness, the carrier concentration of each layer, the bulk defect density of the photosensitive layer, and the work function of the back electrode, the device, as simulated, reached a notable PCE of 32%. The impact of BSO on the performance of Pb-free PSCs was also investigated using SCAPS-1D software. The simulated device achieved an efficiency of 22.09%; however, by optimizing the thickness and defect density of the perovskite layer and charge transport layers, the device’s PCE improved significantly to 28.21% [22].
Recently, Kohan et al. explored a composite ETL combining c-SnO2 and mp-BSO for PSCs, achieving a PCE of 15.54%, which is comparable to or surpasses previously reported values. Notably, the champion device, even without encapsulation, exhibited significantly enhanced thermal stability and photostability, as well as superior long-term air stability, compared to devices utilizing TiO2 as the ETL [16]. Several investigations have focused on optimizing the efficiency of BSO-based PSCs, with various studies incorporating doping techniques as a key approach. Lanthanum (La)-doped BSO was reported as an efficient ETL for PSCs, yielding a PCE of 21.2%, compared to 19.7% for an mp-TiO2 device, with substantial stability enhancement, retaining about 93% of the initial PCE after 1000 h of full sun exposure [23]. Zinc (Zn)-doped BSO was also optimised to enhance the overall performance of BSO-based PSCs, as it has been demonstrated that Zn dopants contribute to a reduction in charge recombination and enhance charge extraction [24].
Another approach that may improve the efficiency of BSO-based PSCs is enhancing the light intensity of the incident light as seen in concentrated photovoltaics (CPVs), a promising technology that ensures high solar cell efficiency [25,26]. Concentrated photovoltaics can direct sunlight onto a highly efficient solar cell by incorporating different optical elements, thereby magnifying the intensity of the input light. Although PSCs have not been extensively explored for CPV technology, the limited theoretical and experimental studies suggest promising prospects for such third-generation photovoltaic materials, showing good stability and achieving a notable PCE, particularly with formamidinium–cesium-based hybrid perovskites [25,27]. It is observed that upon increasing the incident light intensity, the short-circuit current density (JSC) of the PSC increases linearly and the open-circuit voltage (VOC) increases logarithmically. However, the power conversion efficiency (PCE)’s increasing trend is limited due to the substantial reduction in the fill factor (FF) of the device [28]. The series resistance (RS) of the extraction material, layer interfaces, and transparent conductive electrode (TCE) is the main contributor to the FF drop [25,29]. High light intensity magnifies the production of electron–hole pairs; therefore, high charge mobility is essential in both the ETL and HTL to mitigate charge accumulation at layer interfaces and recombination. Spiro-OMeTAD, first reported by U. Bach and co-workers more than twenty years ago, is the predominant hole transport material employed in PSCs [30]. Despite its prevalence, Spiro-OMeTAD exhibits suboptimal charge mobility [31]. Additionally, the high expense and poor stability of organic HTMs hinder their suitability for the commercialisation of high-performance PSCs, as they are prone to degradation from moisture and air exposure. Consequently, inorganic materials such as CuI, Cu2O, NiO, and CuSCN have been studied as HTMs for their inherent stability, effective hole transport performance, and low-cost processing methods [32,33,34], which could be beneficial for PSCs under high light intensity. Another factor to consider during prolonged exposure to light is the increase in temperature, which can result in the degradation of PSC performance [28]. To improve the performance of BSO-based PSCs under high light intensity, it is therefore essential to select materials that provide thermal stability.
The potential of BSO has not been fully explored, and to overcome this gap, the current manuscript focuses on a simulation-based design and performance analysis of BSO-based MAPbI3 and MAFAPbI3 perovskite solar cells under variable light intensity, aiming to improve device efficiency through rational architecture selection. In this context, various design parameters—including HTM choice, operating temperature, BSO layer thickness, donor concentration, and defect density—are systematically investigated to identify the optimal physical parameters to enhance PCE. This theoretical study offers valuable design insights into the underlying performance mechanisms of BSO-based devices, helping experimental researchers and engineers to optimise structural configurations and material properties and predict suitable operating conditions, such as light intensity levels and temperatures, for use in a real scenario. This enables the selection of appropriate concentrated light systems and thermally stable materials before performing time-consuming and costly experiments. This can significantly speed up the development process of designing high-performance PSCs. In addition, the study highlights the potential of BaSnO3 as a stable and efficient electron transport layer, particularly in emerging architectures such as concentrated photovoltaics and low-cost, hole-free carbon-based solar cells.

2. Methodologies

In this study, BSO with two kinds of perovskite absorber materials (MAPbI3 and MA0.5FA0.5PbI3) is used as the ETL and analysed to evaluate device performance for both conventional planar (n-i-p) PSC structures and hole-transport-material-free carbon-based PSCs (CPSCs), configured as FTO/BSO/perovskite/Cu2O or Spiro-OMeTAD/Au, and FTO/BSO/perovskite/C using the one-dimensional Solar Cell Capacitance Simulator (SCAPS-1D) software version 3.3.11. The SCAPS-1D software can simulate solar cells’ performance by solving Poisson’s Equation (1) and the continuity equations of electron (2) and hole (3) for semiconductors [22,35].
2 Ψ = q ε ( n p + N A N D )
. J n + q n t = + q R
. J n + q p t = q R
Here, Ψ denotes the electrostatic potential, ϵ is the permittivity, R is the recombination rate, q is the charge of the electron, n and p are the concentration of free electrons and holes, NA and ND are donor and acceptor density, and Jn and Jp are the electron and hole current densities, respectively.
Figure 1a,b illustrate the structure of the PSCs and CPSCs that were used in this modelling study. Figure 1c depicts the energy band alignment of the layers tested in this theoretical investigation.
The fundamental material parameters used to model the proposed devices were sourced from a range of numerical and experimental investigations, as outlined in Table 1. Other parameters were assumed to be constant in all materials; for example, the thermal velocities of electrons and holes was set as 107 cm. In addition, the interface defects at the BSO/perovskite and perovskite/HTL interfaces were assumed to be single and neutral, with a total defect density of 1010 cm−2 [36]. At the BSO/perovskite interface, the hole and electron capture cross-sections were set as 10−18 cm2 and 10−19 cm2, respectively, and at the perovskite/HTL interface, they were also assumed to be single and neutral with hole and electron capture cross-sections of 10−19 cm2 and 10−18 cm2, respectively [36]. The work functions of FTO, Au, and C have been considered as 5.1 eV, 4.4 eV [22], and 5.2 eV [37], respectively. The simulations were performed under standard test conditions at solar irradiance of 1000 W/m² within AM 1.5G solar spectrum, and operating temperature of 300 K. In addition, the performance of the devices has been evaluated by testing them under different light intensities ranging from 1000 W/m2 to 10,000 W/m2. Figure 2 illustrates the flowchart outlining the simulation workflow for assessing the performance of different configurations of BaSnO3-based perovskite solar cells under different varying conditions using SCAPS-1D.

3. Results and Discussion

The performances of BSO-based PSCs that employ Cu2O and Spiro-OMeTAD as the HTL and BSO-based CPSCs have been studied. The J-V characteristics of the six examined devices are displayed in Figure 3a. A summary of the J-V parameters can be found in Table 2.
MAFAPbI3 demonstrated superior device performance compared to MAPbI3, leading to the conclusion that it is the more appropriate absorber choice for the BSO electron transport layer. This is obviously due to the conduction band alignment between BSO and MAFAPbI3, which facilitates better electron extraction from the absorber and transportation to the front electrode. These findings can be understood by examining the conduction band offset (CBO) at the ETL/perovskite absorber junction. The CBO is obtained by subtracting the electron affinity of the ETL from that of the perovskite absorber (CBO = χperovskite − χETL) [43]. When the CBO is negative, a “cliff” forms at the junction, enhancing electron transport from the perovskite layer to the ETL owing to the strong electric field, thereby increasing the JSC [44]. However, a more negative CBO lowers the activation energy (Ea) for recombination at the ETL/perovskite junction (defined as Ea = Eg − |CBO|), which, in turn, decreases the VOC and overall cell performance [45]. MAFAPbI3 devices recorded the highest JSC and VOC values, potentially due to well-aligned conduction bands, as the CBO value for these devices is −0.142 eV. Such a value results in a large Ea, which, in turn, minimises defect recombination centres at the junction sites and enhances electron transport from the perovskite absorber to the electron transport layer. In the case of MAPbI3, the CBO of about 0 eV results in a significantly smaller Ea at the junction, which increases charge recombination and hence leads to reduced VOC and PCE.
It is worth noting that applying Cu2O as a stable inorganic HTL has shown promising results among different hole transport materials [21,40,46]. Similarly, this finding can be elucidated by examining the valence band offset (VBO) at the HTL/perovskite junction with respect to bandgap alignment. The VBO, along with the built-in electric field (Vbi), is affected by variations in the valence band position of the HTLs. The VBO can be expressed as ((χHTL + EgHTL) − (χPerovskite + EgPerovskite)) [40,43]. It is reported that a negative VBO value leads to a decrease in Vbi, which, in turn, results in VOC reduction [47]. Therefore, the higher VOC observed in Cu2O-based devices is ascribed to their high charge mobilities and suitable VBO values of 0.142 eV for the MAFAPbI3/Cu2O interface and −0.07 eV for the MAPbI3/Cu2O interface, compared to −0.308 eV and −0.52 eV for the corresponding Spiro-OMeTAD-based devices. The high VBO value in the designed MAFAPbI3/Cu2O device is attributed to the energy barrier for electrons, which blocks electron transport through the interface. Interestingly, the hole-free carbon-based MAFAPbI3 device shows a competitive PCE result of 25.384% due to its high VOC value. Such a high VOC value (1.22 V) can be attributed to the lower recombination centre and more favourable energy band alignment between MAFAPbI3 and the carbon electrode, which provide a more stable interface and minimise losses.
The external quantum efficiency (EQE) curves of the six devices are shown in Figure 3b. Incident photons in the visible range are absorbed slightly more efficiently by devices based on MAPbI3 compared to MAFAPbI3. However, the latter shows better performance in near-infrared regions up to 830 nm due to the lower bandgap.

3.1. The Effect of Varying Light Intensity

The theoretical maximum efficiency of a single-junction solar cell is determined by the bandgap of the perovskite absorber. However, the electrical efficiency can be enhanced by increasing the incident light intensity [48]. A higher light intensity results in more photogenerated carriers and greater quasi-Fermi-level splitting in the absorber material [49]. Therefore, increasing the light intensity of the incident light has a major role in determining the performance indicators of solar cells, such as the JSC, the VOC, the FF, and the PCE, as well as the series and shunt resistances [50]. The photovoltaic performance indicators of the devices under simulation were evaluated using J-V characteristics at light intensities ranging from 1000 W/m2 to 10,000 W/m2, as shown in Figure 4.
Increased light intensity has noticeably influenced JSC in all devices. It is found that JSC increases linearly with increasing light intensity in all devices, although MAFAPbI3-based devices have a slightly higher influence compared to MAPbI3-based devices.
The relation between Jsc and light intensity has been further analysed, as illustrated in Figure 5.
According to the power-law dependency, illumination does not alter recombination behaviour in short-circuit conditions [50].
J s c = P l i g h t α
Here, α represents the recombination coefficient, and P denotes the power of the incident light. From the linear relation, α is calculated for both MAPbI3 and MAFAPbI3 as 0.99. Despite the linear increase in JSC and the logarithmic rise recorded in VOC with light intensity, these changes do not directly lead to an improvement in PCE. The open-circuit voltage of single-junction photovoltaic devices operating under focused illumination V o c is generally determined by the saturation current I 0 [29].
V o c = k T q ln I s c I 0  
where k is the Boltzmann constant, T is the absolute temperature, q is the electron charge, and I s c is the short-circuit current under light concentration.
It is clear from Figure 5 that the fill factor of all configurations experiences a reduction with the increasing light intensity of incoming light. The fill factor value is determined by the following equation [50]:
F F = J m a x V m a x J s c V o c
The initial fill factor and the rate of decline vary significantly among the different configurations. Although MAFAPbI3/Cu2O shows the highest initial FF of 85.3, it experiences a dramatic FF reduction until the 4000 W/m2 level of light intensity, then shows a steady reduction. MAPbI3/Cu2O- and MAPbI3-based carbon devices demonstrate a stable FF and the smallest decline under increasing light intensity. Further, MAPbI3/Cu2O shows a slightly lower rate of decline compared to MAFAPbI3-based cells. Spiro-OMeTAD-based devices initially exhibited a moderate FF, which dropped significantly with increasing light intensity up to 4000 W/m2, after which it declined steadily at higher intensities.. On the other hand, MAFAPbI3 carbon-based devices exhibit the lowest initial FF and the most significant decline, showing the least stable FF. The fill factor reduction phenomenon has been identified due to series resistance in charge transport materials, resistance at layer interfaces, and parasitic resistance in transparent conducting electrodes made from FTO [29]. The increase in charge density upon an increase in light concentration may enhance the probability of charge recombination before being extracted, which requires an absorber material with a high carrier lifetime and highly conductive charge transport materials to optimise charge extraction and collection [28].
The fill factor is the main performance indicator influencing the power conversion efficiency of solar cells, which is the fraction of output power to the input power of light. The maximum power that can be generated by solar cells can be calculated using this formula [29]:
P m a x = I s c   V o c   F F
The different materials and configurations have a considerable effect on the PCE. Specifically, solar cells based on MAFAPbI3 consistently show higher efficiencies compared to MAPbI3-based cells. This finding aligns with Troughton J. et al.’s study [27], which identifies that Cs/FA mixed perovskite is better suited for exposure to high light intensities. This improvement in efficiency is attributed to the reduced presence of trap states relative to MAPbI3, which results in its recombination rate being predominantly an outcome of bimolecular mechanisms [27]. As can be seen in Figure 5, increasing light intensity positively influences the efficiency of Cu2O-based device. When light intensity was increased from 1000 W/m2 to 10,000 W/m2, the efficiency of MAPbI3/Cu2O increased from 25.12% to 26.74% and MAFAPbI3/Cu2O increased from 37.13% to 38.2%; therefore, MAPbI3/Cu2O showed a slightly better response to increasing light intensity. Spiro-OMeTAD-based devices remained steady with the increase in the light intensity level. Despite the lowest efficiency being recorded by MAPbI3/C among the designed devices, it showed a positive response with increasing light intensity. Figure 6 illustrates the response of the different designed devices to the light intensities of 3000 W/m2, 6000 W/m2, and 9000 W/m2.

3.2. The Effect of Operation Temperature

Exposure to high light intensities typically leads to temperature elevation in solar cells. Therefore, evaluating the impact of rising temperatures on the performance of PSCs is crucial for determining their suitability in CPV applications. This simulation has examined the effect of increasing temperature on the functionality of the designed PSCs and CPSCs. The operating temperature was varied from 280 K to 400 K to observe the performance of the proposed devices, while a temperature of 298 K was considered the standard measuring temperature, representing the standard room temperature. The performance of the PSCs and CPSCs was significantly influenced by prolonged exposure to sunlight, which caused notable variations in their operating temperature and consequently affected their output performance. The simulation results are displayed in Figure 6, depicting the normalised values of all PV performance indicators under different operating temperatures.
It is evident from Figure 7 that the normalised values of JSC for all configurations significantly vary with temperature. All configurations show JSC enhancement with elevating temperature except for MAFAPbI3/Cu2O and MAFAPbI3/C. However, the reduction is negligible as when the temperature rises from 280 K to 400 K, Jsc changes from 25.681 mA/cm2 to 25.667 mA/cm2 and from 25.563 mA/cm2 to 25.556 mA/cm2 for MAFAPbI3/Cu2O and MAFAPbI3/C, respectively. VOC shows a general reduction trend with increasing temperature in all configurations. However, MAFAPbI3/Spiro-OMeTAD reaches its maximum VOC value of 1.198 V at 320 K and then shows a reduction. At higher temperatures, the material’s carrier concentration, charge mobility, resistance, and bandgap are significantly affected, resulting in changes to their photovoltaic (PV) parameters [21]. When the temperature increases, JSC increases due to the thermal generation of minority carriers, while VOC decreases due to an elevated saturation current. Temperature impacts the diffusion length of carriers, leading to higher series resistance (Rs), which, in turn, reduces both the FF and the PCE [51], as depicted in Figure 7. The relationship between FF and RS is detailed in Equation (8) [35].
F F S = F F 0 ( 1 R S )
Here, F F 0 indicates the fill factor without the influence of R S .
Interestingly, the FF of all configurations shows a slight increase with temperature to reach its maximum value and then drops, except for in the MAPbI3/Cu2O solar cell, which remains constant until 300 K and then drops with increasing temperature. The PCE of the simulated devices follows the reduction trend observed in the FF. Figure 7 shows how the PCE changes with increasing temperature in all configurations. The most significant finding is that the MAFAPbI3/Spiro-OMeTAD solar cell was the most robust as its FF and PCE continued to increase with temperature and achieved its optimum values of 82.16% and 25.2%, respectively, at 320 K. To evaluate the performance of the designed PSCs and CPSCs more comprehensively, the temperature coefficient (CT) of the PCE of the solar cells ( T C P C E ) was calculated using Equation (9) [21]. Table 3 represents the calculated TCPCE of the different devices stimulated in this study and others obtained from the literature [52].
C T = (   1 η S T C   d η T d T × 100 )
Here, ηSTC is the efficiency in standard test conditions (298 K) and ηT is the efficiency at temperature T.
Compared to a previous simulation study of an optimised BSO/MAPbI3/Cu2O cell, which reported a T C P C E of −0.112% K−1 [21], the simulated MAFAPbI3-based solar cells in our study exhibit significantly lower T C P C E values. Despite the shortage of studies on the T C P C E of perovskite solar cells, it has been proven that they could perform well for PV applications that require high temperatures compared to other solar cells [28,52]. This finding underscores the superior photovoltaic performance stability of BSO/MAFAPbI3-based PSCs at elevated temperatures, demonstrating their potential for enhanced operational reliability in high-temperature environments. In particular, the sample incorporating Spiro-OMeTAD as the hole transport material shows an increase in PCE up to 320 K, and at 400 K, the T C P C E exhibits only a slight reduction of 0.04% K−1, highlighting its advantages for applications in high-temperature conditions.

3.3. The Impact of BSO Thickness, Donor Density, and Defect Density

It is worth noting that the designed devices have not been optimised. The electron transport material has a great influence on the performance of PSCs. The thickness, donor concentration, and defect density of BSO/perovskite interfaces were investigated in this study to determine the optimised BSO parameters. For high-performance PSCs, it is crucial that the ETL exhibits high transmittance and uniform morphology, allowing incoming photons to successfully reach the perovskite with a low recombination rate. Here, the thickness of the BSO ETL was varied from 10 to 100 nm to investigate its effect on device performance parameters, as illustrated in Figure 8.
The JSC of all devices was unaffected by increasing the BSO thickness. Additionally, other performance indicators of the MAPbI3-based devices remained steady and did not vary with BSO thickness. Therefore, it can be inferred that the band offset, intrinsic resistance, and defect density in the BSO layer remain constant as the thickness increases. On the contrary, a greater thickness of BSO negatively influences the VOC, FF, and PCE of MAFAPbI3-based devices. The VOC decreased slightly with increasing BSO thickness, and after a thickness of 30 nm, no reduction in VOC was observed. The decrease in FF and PCE stabilised at BSO thicknesses of 80 nm and 50 nm, respectively. From Figure 8, it can be observed that the PCE of MAFAI3-based devices decreases by approximately 15% as the BSO thickness increases from 10 nm to 60 nm. This reduction can be mainly explained by the longer distance required for photogenerated electrons to travel in thicker ETLs, which promotes carrier recombination and consequently leads to a reduction in overall device efficiency. Therefore, a BSO thickness of 10 nm results in the best performance. However, at such low ETL thickness, it is crucial to maintain uniform full coverage to ensure good contact with the perovskite layer and to avoid direct contact with the FTO substrate.
In laboratory experiments, researchers have adopted several dopants with various concentrations to improve the donor charge density and modify the conduction band alignment of ETLs. In this study, to stimulate this impact, the donor density of BSO was varied from 1014 to 1021 cm−3, as illustrated in Figure 9.
The performance of MAPbI3-based devices was not influenced by an increase in donor density and almost remained constant. However, the performance indicators of MAFAPbI3-based devices were enhanced with rising doping concentrations, except for JSC, which remained steady. From Figure 9, the performance of MAFAPbI3-based solar cells can be optimised and enhanced at a BSO doping density of at least 1020 cm−3. The VOC and PCE of the BSO/MAFAPbI3/Cu2O device showed a direct increase with an increased donor density of BSO. Such an increase could be attributed to the shift in band offsets at the BSO/MAFAPbI3 interface. In addition, using BSO with high doping levels enhances material conductivity, ensuring effective electron transportation [57], and reduces the depletion region in the ETL, creating a “cliff” that blocks minority carriers. This helps separate the photogenerated electron–hole pairs more efficiently and reduces charge carrier recombination [58]. On the contrary, excessive doping will increase non-radiative recombination owing to donor defect formation, which will lead to an increase in the series resistance of the PSCs and, hence, lower PCE [59]. Experimentally, the doping concentration of BSO should be optimised in order to obtain high transparency and high conductivity.
Depending on the synthesis technique, the defect density at the ETL/perovskite absorber junction can alter the device’s performance. Herein, the defect density at the BSO/perovskite junction was varied from 107 cm−2 to 1019 cm−2, while other parameters were not changed. The result of varying the defect densities at BSO/perovskite interfaces on structured devices is illustrated in Figure 10.
It is clear that all configurations were negatively affected by increasing the defect density at the BSO/perovskite junction. The JSC of the MAPbI3-based devices was more negatively impacted compared to the MAFAPbI3-based ones, as they remained constant and then dropped at a defect density of 1015 cm−2. MAFAPbI3-based devices showed a slight decrease in JSC, except for the carbon-based device, which was highly affected at a defect density beyond 1013 cm−2. The VOC of MAPbI3/C and MAFAPbI3/Cu2O was stable and did not account for interface defects.
At defect densities higher than 108 cm−2, MAFAPbI3/Spiro-OMeTAD and MAFAPbI3/C showed a reduction in VOC, while MAPbI3/Spiro-OMeTAD and MAPbI3/C were more stable, only exhibiting a VOC reduction at defect densities higher than 1014 cm−2. The devices’ performance dependency on the interface defects at the BSO/perovskite junction has been translated into the devices’ overall PCEs.
The Cu2O-based MAFAPbI3 cell is the most stable and reliable device, which can withstand increasing interface defects, as well as carbon-based MAPbI3. The PCE of the devices with Spiro-OMeTAD was significantly reduced when subjected to a higher defect interface density. To enhance the performance, the interface defect density at the BSO/MAPbI3 interface must be limited to less than 1014 cm−2, and it should be below 1010 cm−2 for MAFAPbI3/Spiro-OMeTAD and MAFAPbI3/C devices.
The comparison of BSO-based simulated devices with varying parameters as the main contributor is highlighted in Table 4.

4. Conclusions and Future Work Suggestions

In this study, the performance of perovskite solar cells employing BaSnO3 as the electron transport layer was evaluated using MAPbI3- and MAFAPbI3-based perovskite absorbers. Various hole transport materials and hole-free (carbon-based) configurations were implemented and analysed under varying light intensities, temperatures, donor concentrations, and defect densities using SCAPS, a one-dimensional solar cell simulator. In normal conditions at a light intensity of 1000 W/m2, MAFAPbI3 devices demonstrated greater performance compared to the traditional perovskite absorber MAPbI3. As light intensity increased, devices using Cu2O showed improved performance. MAPbI3 with a carbon-based electrode showed a promising trend with increasing light intensity, recording the smallest fill factor drop among all configurations. Such a result demonstrates that conduction band alignment, high-mobility charge transport materials, and stable thermal perovskite absorbers are essential in highly efficient and stable devices for concentrated PSC applications. All devices were negatively impacted by elevating temperature, with Voc exhibiting an initial drop and FF decreasing with temperature beyond 300 K, except for Spiro-OMeTAD devices, which were the most robust. Decreasing the thickness and increasing the donor concentration of BSO had a substantial effect in enhancing the performance of MAFAPbI3-based devices. The defect density at the BSO/perovskite interface also influenced devices performance; however, Cu2O-based devices exhibited the least sensitivity to interface defects. FTO/BSO/MAFAPbI3/Cu2O/Au was the optimal device, showing an outstanding efficiency and stability compared to the other cells, yielding a PCE of 38.2% with a JSC = 343.3 mA/cm2, VOC = 1.39 V, and FF = 80.17% under 10,000 W/m2 light intensity. Moreover, it showed a good temperature coefficient of PCE of −0.078% K−1, ensuring efficient and stable performance under high exposure to light and heat up to 400 K. At a normal light intensity (1000 W/m2), this suggested device recorded a PCE of 29% at 10 nm BSO thickness, and its PCE improved from 27% to 33% when increasing the donor concentration from 1014 cm−3 to 1021 cm−3. The optimal device showed a slight drop in PCE from ~28% to 27% when elevating the interface defect density from 107 cm−2 to 1019 cm−2. For high-temperature PSC applications, Spiro-OMeTAD and BSO could be suggested as effective charge transport materials for mixed-cation MAFAPbI3, as it exhibited a slow PCE degradation rate of −0.04% K−1.
This simulation provides a valuable framework for researchers investigating the development of concentrated perovskite solar cells by highlighting the potential of adopting BSO, high-mobility hole transport materials, and mixed-cation perovskite absorber materials. Furthermore, it opens new avenues for enhancing the efficiency of low-cost, carbon-based perovskite solar cells by integrating BSO as an electron transport layer in concentrated photovoltaic (CPV) applications.
In future work, several strategies could be pursued to further improve the performance of BSO-based devices. These include employing deposition techniques that enable the formation of thin, uniform films such as spin-coating and spray pyrolysis, as well as using doping methods to tailor the conduction band level of BSO, increase donor concentration, and enhance electron transport properties. Further research includes using BSO as an efficient electron transport material for perovskite solar cells illuminated with varying light intensities.

Author Contributions

N.A.: article preparation, data analysis, and conceptualisation; S.B.: conceptualisation and direction; T.K.M.: conceptualisation and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Government of the Kingdom of Saudi Arabia through a PhD scholarship awarded to Nouf Alkathran.

Data Availability Statement

The data presented in this study are available upon reasonable request from the corresponding authors.

Acknowledgments

The authors gratefully acknowledge Marc Burgelman, University of Gent, Belgium, for providing the SCAPS 1-D simulation software.

Conflicts of Interest

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

Abbreviations

AbbreviationDefinition
PSCsPerovskite solar cells
ETLElectron transport layer
HTLHole transport layer
MAPbI3Methylammonium lead iodide
MAFAPbI3Methylammonium formamidinium lead iodide
ETMsElectron transport materials
PCEPower conversion efficiency
CPVConcentrated photovoltaic
JSCShort-circuit current density
VOCOpen-circuit voltage
FFFill factor
RSSeries resistance
BSOBaSnO3 barium stannate oxide
CPSCsCarbon-based perovskite solar cells
EQEExternal quantum efficiency
CBOConduction band offset
VBOValence band offset
T C P C E Temperature coefficient (CT) of the PCE of solar cells
CTTemperature coefficient
ηSTCEfficiency at standard test conditions
ηTEfficiency at temperature T
EaActivation energy for recombination
EgEnergy bandgap
χ Electron affinity
NDDonor concentration density
NtDefect density
ΨElectrostatic potential
ϵPermittivity
RRecombination rate
qCharge of the electron
nConcentration of free electrons
pConcentration of free holes
NAAcceptor concentration density
JnElectron density
JpHoles density
αRecombination coefficient
PlightPower of the incident light
kBoltzmann constant
TAbsolute temperature
I 0 Saturation current
I s c * Short-circuit current under light concentration
V o c * Open-circuit voltage under light concentration
P m a x * Maximum power under light concentration
J m a x Maximum current density
V m a x Maximum voltage
F F 0 Fill factor without R S influence

References

  1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, Z.; Liu, T.; Yu, L.; He, T.; Liu, P.; Li, M.; Yuan, M. Efficient and Stable FA-Rich Perovskite Photovoltaics: From Material Properties to Device Optimization; John Wiley and Sons Inc.: Hoboken, NJ, USA, 2022. [Google Scholar] [CrossRef]
  3. Wang, R.; Huang, T.; Xue, J.; Tong, J.; Zhu, K.; Yang, Y. Prospects for Metal Halide Perovskite-Based Tandem Solar Cells. Nat. Photonics 2021, 15, 411–425. [Google Scholar] [CrossRef]
  4. Bhandari, S.; Roy, A.; Mallick, T.K.; Sundaram, S. Impact of different light induced effect on organic hole-transporting layer in perovskite solar cells. Mater. Lett. 2020, 268, 127568. [Google Scholar] [CrossRef]
  5. Bhandari, S.; Valsalakumar, S.; Chanchangi, Y.; Selvaraj, P.; Mallick, T.K. Effect of novel graphitic carbon/NiO hole transporting electrode on the photovoltaic and optical performance of semi-transparent perovskite solar cells. RSC Adv. 2023, 13, 7380–7384. [Google Scholar] [CrossRef]
  6. Bello, S.; Urwick, A.; Bastianini, F.; Nedoma, A.J.; Dunbar, A. An introduction to perovskites for solar cells and their characterisation. Energy Rep. 2022, 8, 89–106. [Google Scholar] [CrossRef]
  7. Tonui, P.; Oseni, S.O.; Sharma, G.; Yan, Q.; Mola, G.T. Perovskites Photovoltaic Solar Cells: An Overview of Current Status; Elsevier Ltd.: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  8. Hussain, I.; Tran, H.P.; Jaksik, J.; Moore, J.; Islam, N.; Uddin, M.J. Functional materials, device architecture, and flexibility of perovskite solar cell. Emergent Mater. 2018, 1, 133–154. [Google Scholar] [CrossRef]
  9. Hui, W.; Zhang, B.; Wang, B.; Gu, L.; Bao, Y.; Kang, X.; Su, Z.; Li, D.; Huang, W.; Song, L.; et al. Stable Electron-Transport-Layer-Free Perovskite Solar Cells with over 22% Power Conversion Efficiency. Nano Lett. 2023, 23, 2195–2202. [Google Scholar] [CrossRef]
  10. Li, S.; Li, Y.; Sun, X.; Li, Y.; Deng, F.; Tao, X. Hole transport layer-free carbon-based perovskite solar cells with high-efficiency up to 17.49% in air: From-bottom-to-top perovskite interface modification. Chem. Eng. J. 2023, 455, 140727. [Google Scholar] [CrossRef]
  11. Lu, H.; Jiang, Y.; Zhong, J.; Zhao, R.; Liang, T.; Li, H.; Zhao, J.; Fang, S.; Ji, C.; Li, D.; et al. Fabricating an optimal rutile TiO2 electron transport layer by delicately tuning TiCl4 precursor solution for high performance perovskite solar cells. Nano Energy 2020, 68, 104336. [Google Scholar] [CrossRef]
  12. Myung, C.W.; Lee, G.; Kim, K.S. La-doped BaSnO3 electron transport layer for perovskite solar cells. J. Mater. Chem. A Mater. 2018, 6, 23071–23077. [Google Scholar] [CrossRef]
  13. Leijtens, T.; Eperon, G.E.; Pathak, S.; Abate, A.; Lee, M.M.; Snaith, H.J. Overcoming ultraviolet light instability of sensitized TiO2 with meso-superstructured organometal tri-halide perovskite solar cells. Nat. Commun. 2013, 4, 3885. [Google Scholar] [CrossRef] [PubMed]
  14. Raj, A.; Sharma, S.; Singh, K.; Laref, A.; Anshul, A.; Kumar, M.; Kumar, A. Effect of doping engineering in TiO2 electron transport layer on photovoltaic performance of perovskite solar cells. Mater. Lett. 2022, 313, 131692. [Google Scholar] [CrossRef]
  15. Wu, W.Q.; Chen, D.; Li, F.; Cheng, Y.B.; Caruso, R.A. Solution-processed Zn2SnO4 electron transporting layer for efficient planar perovskite solar cells. Mater. Today Energy 2018, 7, 260–266. [Google Scholar] [CrossRef]
  16. Kohan, M.; Mahmoudi, T.; Wang, Y.; Im, Y.H.; Hahn, Y.B. SnO2/BaSnO3 electron transport materials for stable and efficient perovskite solar cells. Appl. Surf. Sci. 2023, 613, 156068. [Google Scholar] [CrossRef]
  17. Guo, Q.; Dong, J.; Lin, J.; Huang, M.; Lan, Z.; Huang, Y.; Wei, Y.; Liu, X.; Yang, Y.; Jia, J.; et al. High-Performance and Hysteresis-Free Perovskite Solar Cells Based on Rare-Earth-Doped SnO2 Mesoporous Scaffold. Research 2019, 2019, 4049793. [Google Scholar] [CrossRef]
  18. Shin, S.S.; Park, J.H.; Kim, J.S.; Suk, J.H.; Lee, K.D.; Hong, K.S.; Kim, D.W.; Kim, J.Y.; Cho, I.S. Improved quantum efficiency of highly efficient perovskite BaSnO3-based dye-sensitized solar cells. ACS Nano 2013, 7, 1027–1035. [Google Scholar] [CrossRef]
  19. Shin, S.S.; Han, B.S.; Kim, J.S.; Hwang, D.; Oh, L.S.; Suk, J.H.; Kim, J.Y.; Kim, D.H.; Hong, K.S.; Kim, D.W. Controlled interfacial electron dynamics in highly efficient Zn2SnO4-based dye-sensitized solar cells. ChemSusChem 2014, 7, 501–509. [Google Scholar] [CrossRef]
  20. Sun, C.; Fang, B.; Yang, J.; Chen, Y.; Guan, L.; Liu, H.; Li, H.; Guo, Y.; Duan, H. Ternary oxide BaSnO3 nanoparticles as an efficient electron-transporting layer for planar perovskite solar cells. J. Alloys Compd. 2017, 722, 196–206. [Google Scholar] [CrossRef]
  21. Rahman, A. Design and Simulation of High-performance Planar npp+ Heterojunction CH3NH3PbI3 Based Perovskite Solar Cells Using BaSnO3 ETM and Cu2O HTM. Res. Sq. 2021. [Google Scholar] [CrossRef]
  22. Tara, A.; Bharti, V.; Dixit, H.; Sharma, S.; Gupta, R. Performance evaluation of all-inorganic cesium-based perovskite solar cell with BaSnO3 as ETL. J. Nanoparticle Res. 2023, 25, 9. [Google Scholar] [CrossRef]
  23. Shin, S.S.; Gao, Z.R.; Lin, Q.-F.; Liu, C.; Gao, F.; Lin, C.; Zhang, S.; Deng, H.; Mayoral, A.; Fan, W.; et al. Colloidally Prepared La-Doped BaSnO3 Electrodes for Efficient, Photostable Perovskite Solar Cells. 2017. Available online: https://www.science.org (accessed on 12 December 2024).
  24. Dileep, K.R.; Veerappan, G.; Rao, T.N.; Mallick, S.; Ashina, A.; Rajbhar, M.K.; Ramasamy, E. A facile co-precipitation method for synthesis of Zn doped BaSnO3 nanoparticles for photovoltaic application. Mater. Chem. Phys. 2021, 258, 123939. [Google Scholar] [CrossRef]
  25. Wang, Z.; Klug, M.T.; Wenger, B.; Herz, L.M.; Johnston, M.B.; Lin, Q.; Snaith, H.J.; Christoforo, M.G.; Lin, Y.-H.; Klug, M.T.; et al. High irradiance performance of metal halide perovskites for concentrator photovoltaics. Nat. Energy 2018, 3, 855–861. [Google Scholar] [CrossRef]
  26. Zhou, Y.; Gray-Weale, A. A numerical model for charge transport and energy conversion of perovskite solar cells. Phys. Chem. Chem. Phys. 2016, 18, 4476–4486. [Google Scholar] [CrossRef] [PubMed]
  27. Troughton, J.; Gasparini, N.; Baran, D. Cs0.15FA0.85PbI3 perovskite solar cells for concentrator photovoltaic applications. J. Mater. Chem. A Mater. 2018, 6, 21913–21917. [Google Scholar] [CrossRef]
  28. Sadhukhan, P.; Nazeeruddin, M.K.; Sengupta, P.; Sundaram, S.; Roy, A.; Das, S.; Mallick, T.K. The Emergence of Concentrator Photovoltaics for Perovskite Solar Cells. Appl. Phys. Rev. 2021, 8, 041324. [Google Scholar] [CrossRef]
  29. Baig, H.; Kanda, H.; Asiri, A.M.; Nazeeruddin, M.K.; Mallick, T. Increasing efficiency of perovskite solar cells using low concentrating photovoltaic systems. Sustain. Energy Fuels 2020, 4, 528–537. [Google Scholar] [CrossRef]
  30. Bach, U.; Spreitzer, H.; Lupo, D.; Comte, P.; Weissörtel, F.; Grätzel, M.; Salbeck, J.; Moser, J.E. Solid-State Dye-Sensitized Mesoporous TiO2 Solar Cells with High Photon-to-Electron Conversion Efficiencies. 1998. Available online: www.nature.com (accessed on 12 December 2024).
  31. Nguyen, W.H.; Bailie, C.D.; Unger, E.L.; McGehee, M.D. Enhancing the hole-conductivity of spiro-OMeTAD without oxygen or lithium salts by using spiro(TFSI)2 in perovskite and dye-sensitized solar cells. J. Am. Chem. Soc. 2014, 136, 10996–11001. [Google Scholar] [CrossRef]
  32. Bhandari, S.; Mallick, T.K.; Sundaram, S. Enlightening the temperature coefficient of triple mesoscopic CH3NH3PbI3−xClx/NiO and double mesoscopic CsFAMAPbI3−xBrx/CuSCN carbon perovskite solar cells. J. Phys. Energy 2023, 5, 2. [Google Scholar] [CrossRef]
  33. Christians, J.A.; Fung, R.C.M.; Kamat, P.V. An inorganic hole conductor for Organo-lead halide perovskite solar cells. improved hole conductivity with copper iodide. J. Am. Chem. Soc. 2014, 136, 758–764. [Google Scholar] [CrossRef]
  34. Ito, S.; Tanaka, S.; Vahlman, H.; Nishino, H.; Manabe, K.; Lund, P. Carbon-double-bond-free printed solar cells from TiO2/CH3NH3PbI3/CuSCN/Au: Structural control and photoaging effects. ChemPhysChem 2014, 15, 1194–1200. [Google Scholar] [CrossRef]
  35. Srivastava, V.; Chauhan, R.K.; Lohia, P. Highly efficient cesium-based halide perovskite solar cell using SCAPS-1D software: Theoretical study. J. Opt. 2023, 52, 1218–1225. [Google Scholar] [CrossRef]
  36. Son, C.; Son, H.; Jeong, B.S. Enhanced Conversion Efficiency in MAPbI3 Perovskite Solar Cells through Parameters Optimization via SCAPS-1D Simulation. Appl. Sci. 2024, 14, 2390. [Google Scholar] [CrossRef]
  37. Valsalakumar, S.; Bhandari, S.; Mallick, T.K.; Hinshelwood, J.; Sundaram, S. Experimental Validation of Optimized Solar Cell Capacitance Simulation for Rheology-Modulated Carbon-Based Hole Transport Layer-Free Perovskite Solar Cell. Adv. Energy Sustain. Res. 2024, 5, 244. [Google Scholar] [CrossRef]
  38. Abena, A.M.N.; Ngoupo, A.T.; Abega, F.X.A.; Ndjaka, J.M.B. Numerical investigation of solar cells based on hybrid organic cation perovskite with inorganic HTL via SCAPS-1D. Chin. J. Phys. 2022, 76, 94–109. [Google Scholar] [CrossRef]
  39. Tan, K.; Lin, P.; Wang, G.; Liu, Y.; Xu, Z.; Lin, Y. Controllable design of solid-state perovskite solar cells by SCAPS device simulation. Solid State Electron. 2016, 126, 75–80. [Google Scholar] [CrossRef]
  40. Hossain, M.I.; Alharbi, F.H.; Tabet, N. Copper oxide as inorganic hole transport material for lead halide perovskite based solar cells. Solar Energy 2015, 120, 370–380. [Google Scholar] [CrossRef]
  41. Abdelaziz, S.; Zekry, A.; Shaker, A.; Abouelatta, M. Investigating the performance of formamidinium tin-based perovskite solar cell by SCAPS device simulation. Opt. Mater. 2020, 101, 109738. [Google Scholar] [CrossRef]
  42. Abena, A.M.N.; Ngoupo, A.T.; Ndjaka, J.M.B. Computational analysis of mixed cation mixed halide-based perovskite solar cell using SCAPS-1D software. Heliyon 2022, 8, e11428. [Google Scholar] [CrossRef]
  43. Raza, E.; Ahmad, Z.; Riaz, K.; Aziz, F.; Bhadra, J.; Al-Thani, N.J.; Asif, M.; Ahmed, A. Numerical simulation analysis towards the effect of charge transport layers electrical properties on cesium based ternary cation perovskite solar cells performance. Solar Energy 2021, 225, 842–850. [Google Scholar] [CrossRef]
  44. Minemoto, T.; Murata, M. Theoretical analysis on effect of band offsets in perovskite solar cells. Sol. Energy Mater. Sol. Cells 2015, 133, 8–14. [Google Scholar] [CrossRef]
  45. Ahmed, A.; Riaz, K.; Mehmood, H.; Tauqeer, T.; Ahmad, Z. Performance optimization of CH3NH3Pb(I1-xBrx)3 based perovskite solar cells by comparing different ETL materials through conduction band offset engineering. Opt. Mater. 2020, 105, 109897. [Google Scholar] [CrossRef]
  46. Xu, L.; Imenabadi, R.M.; Vandenberghe, W.G.; Hsu, J.W.P. Minimizing performance degradation induced by interfacial recombination in perovskite solar cells through tailoring of the transport layer electronic properties. APL Mater. 2018, 6, 5021138. [Google Scholar] [CrossRef]
  47. Cheng, N.; Li, W.; Sun, S.; Zhao, Z.; Xiao, Z.; Sun, Z.; Zi, W.; Fang, L. A simulation study of valence band offset engineering at the perovskite/Cu2ZnSn(Se1-xSx)4 interface for enhanced performance. Mater. Sci. Semicond. Process. 2019, 90, 59–64. [Google Scholar] [CrossRef]
  48. Sharma, S.; Jain, K.K.; Sharma, A. Solar Cells: In Research and Applications—A Review. Mater. Sci. Appl. 2015, 6, 1145–1155. [Google Scholar] [CrossRef]
  49. El Himer, S.; Salvestrini, J.P.; El Himer, S.; Ahaitouf, A.; El Ayane, S. Photovoltaic Concentration: Research and Development. Energies 2020, 13, 5721. [Google Scholar] [CrossRef]
  50. Zerfaoui, H.; Dib, D.; Kadem, B. The Simulated Effects of Different Light Intensities on the SiC-Based Solar Cells. Silicon 2019, 11, 1917–1923. [Google Scholar] [CrossRef]
  51. Ahmed, S.; Jannat, F.; Khan, M.A.K.; Alim, M.A. Numerical development of eco-friendly Cs2TiBr6 based perovskite solar cell with all-inorganic charge transport materials via SCAPS-1D. Optik 2021, 225, 165765. [Google Scholar] [CrossRef]
  52. Dong, Z.; Zhu, L.; Liu, H.; Jiang, X.; Wang, H.; Li, W.; Chen, H. High-Temperature Perovskite Solar Cells. Solar RRL 2021, 5, 370. [Google Scholar] [CrossRef]
  53. Ghosh, B.K.; Hasanuzzman, M.; Saad, I.; Mohamad, K.A.; Hossain, M.K. Photovoltaic technologies photo-thermal challenges: Thin active layer solar cells significance. Optik 2023, 274, 170567. [Google Scholar] [CrossRef]
  54. Bhandari, S.; Roy, A.; Ghosh, A.; Mallick, T.K.; Sundaram, S. Perceiving the temperature coefficients of carbon-based perovskite solar cells. Sustain. Energy Fuels 2020, 4, 6283–6298. [Google Scholar] [CrossRef]
  55. Moot, T.; Parilla, P.A.; Johnston, S.W.; Morales, D.; Gould, I.E.; Wolf, E.J.; McAndrews, G.; Patel, J.B.; McGehee, M.D.; Luther, J.M.; et al. Temperature Coefficients of Perovskite Photovoltaics for Energy Yield Calculations. ACS Energy Lett. 2021, 6, 2038–2047. [Google Scholar] [CrossRef] [PubMed]
  56. Tress, W.; Carlsen, B.; Alharbi, E.A.; Agarwalla, A.; Graetzel, M.; Domanski, K.; Hagfeldt, A. Performance of perovskite solar cells under simulated temperature-illumination real-world operating conditions. Nat. Energy 2019, 4, 568–574. [Google Scholar] [CrossRef]
  57. Guo, Y.; Xue, Y.; Geng, C.; Li, C.; Li, X.; Niu, Y. Structural, Electronic, and Optical Characterizations of the Interface between CH3NH3PbI3 and BaSnO3 Perovskite: A First-Principles Study. J. Phys. Chem. C 2019, 123, 16075–16082. [Google Scholar] [CrossRef]
  58. Liang, C.; Xing, G. Doping Electron Transporting Layer: An Effective Method to Enhance JSC of All-Inorganic Perovskite Solar Cells. Energy Environ. Mater. 2021, 4, 500–501. [Google Scholar] [CrossRef]
  59. Alla, M.; Boubker, F.; Choudhary, E.; Bimli, S.; Rouchdi, M.; Samtham, M.; Kasaudhan, A.; Manjunath, V. Towards lead-free all-inorganic perovskite solar cell with theoretical efficiency approaching 23%. Mater. Technol. 2022, 37, 2963–2969. [Google Scholar] [CrossRef]
  60. Verma, U.K.; Singh, A. International Journal of Convergence in Healthcare Effect of Intensity on Perovskite Solar Cells Parameters by SCAPS-1D Simulation. Available online: www.ijcih.com (accessed on 20 December 2024).
  61. Li, H.; Huang, Y.; Zhu, M.; Yan, P.; Sheng, C. Analyzing Efficiency of Perovskite Solar Cells Under High Illumination Intensities by SCAPS Device Simulation. Nanomaterials 2025, 15, 286. [Google Scholar] [CrossRef]
  62. Shivesh, K.; Singh, S.V.; Kushwaha, A.K.; Kumar, M.; Alam, I. Investigating the theoretical performance of Cs2TiBr6-based perovskite solar cell with La-doped BaSnO3 and CuSbS2 as the charge transport layers. Int. J. Energy Res. 2021. [Google Scholar]
  63. Alghafis, A.; Sobayel, K. Performance evaluation of inorganic Cs3Bi2I9-based perovskite solar cells with BaSnO3 charge transport layer. Phys. Scr. 2024, 99, 7233. [Google Scholar] [CrossRef]
Figure 1. The device structure of (a) PSC and (b) CPSC, and (c) the diagram depicting the energy level alignment.
Figure 1. The device structure of (a) PSC and (b) CPSC, and (c) the diagram depicting the energy level alignment.
Designs 09 00076 g001
Figure 2. Flowchart of the simulation methodology used in SCAPS software.
Figure 2. Flowchart of the simulation methodology used in SCAPS software.
Designs 09 00076 g002
Figure 3. (a) J-V characteristics of PSCs with different transport layers and (b) their corresponding EQE curves.
Figure 3. (a) J-V characteristics of PSCs with different transport layers and (b) their corresponding EQE curves.
Designs 09 00076 g003
Figure 4. The J-V characteristics of the simulated devices.
Figure 4. The J-V characteristics of the simulated devices.
Designs 09 00076 g004
Figure 5. Influence of light intensity variation on the performance of perovskite solar cells.
Figure 5. Influence of light intensity variation on the performance of perovskite solar cells.
Designs 09 00076 g005
Figure 6. J-V curves of the simulated devices at different light intensities.
Figure 6. J-V curves of the simulated devices at different light intensities.
Designs 09 00076 g006
Figure 7. Normalised values of PV performance indicators of the simulated devices under different operation temperatures.
Figure 7. Normalised values of PV performance indicators of the simulated devices under different operation temperatures.
Designs 09 00076 g007
Figure 8. Dependence of device performance parameters on thickness of BSO.
Figure 8. Dependence of device performance parameters on thickness of BSO.
Designs 09 00076 g008
Figure 9. BSO donor concentration impact on the performance parameters of devices.
Figure 9. BSO donor concentration impact on the performance parameters of devices.
Designs 09 00076 g009
Figure 10. Defect density impact at BSO/perovskite interfaces.
Figure 10. Defect density impact at BSO/perovskite interfaces.
Designs 09 00076 g010
Table 1. Input parameters of each layer.
Table 1. Input parameters of each layer.
ParametersFTO
[38]
BSO
[22]
MAFAPbI3
[39]
MAPbI3
[36]
Cu2O
[40]
Spiro-OMeTAD [41]
Thickness (nm)50050 (vary)450600200200
Bandgap, Eg (eV)3.53.161.481.552.172.88
Electron affinity, χ (eV)43.903.758 3.93.202.05
Relative permittivity,  ϵr9176.5327.113
CB effective density of states, NC (cm−3)2.2 × 10171.2 × 10192.2 × 10152.8 × 10202.02 × 10172.2 × 1018
VB effective density of states, NV (cm−3)2.2 × 10161.8 × 10192.2 × 10173.9 × 10201.1 × 10191.8 × 1019
Electron mobility, μn (cm2/Vs)202002 11.82002.0 × 10−4
Hole mobility, μp (cm2/Vs)10252 11.8802.0 × 10−4
Donor density, ND (cm−3)1.0 × 1019
[22]
1.0 × 10171091.0 × 101300
Acceptor density, NA (cm−3)00.0 1091.0 × 10131 × 10182.0 × 1019
Defect density, Nt (cm−3)1 × 10141.0 × 10151.0 × 1013 [42]3.0 × 10141.0 × 1015 [36]1.0 × 1015
Table 2. Performance indicators of BSO-based devices.
Table 2. Performance indicators of BSO-based devices.
Device ConfigurationVOC (V)JSC (mA/cm2)FF (%)PCE (%)
FTO/BSO/MAPbI3/Cu2O/Au0.93899724.68617981.661718.66
FTO/BSO/MAPbI3/Spiro-OMeTAD/Au0.90834924.31952782.05518.1264
FTO/BSO/MAPbI3/C0.82797124.291724283.769216.8484
FTO/BSO/MAFAPbI3/Cu2O/Au1.2541625.67772785.914127.6678
FTO/BSO/MAFAPbI3/Spiro-OMeTAD/Au1.1969825.6388682.061725.184
FTO/BSO/MAFAPbI3/C1.22169625.56146481.278825.382
Table 3. The temperature coefficients of the simulated solar cells.
Table 3. The temperature coefficients of the simulated solar cells.
Solar CellsTCPCE (%)/Temperature UnitRef.
c-Si-based module−0.45[53]
a-Si-based module−0.13[53]
CdTe-based module−0.21[53]
CIGS-based module−0.36[53]
Organic+0.4[52]
DSSC−0.79[52]
TiO2/CsPbI2Br C-PSC−0.23/°C at 200 °C, in reverse measurement [52]
TiO2/CsPbI2Br C-PSC 0.035   at   200   ° C, in forward measurement[52]
BSO/MAPbI3/Cu2O, −0.112 at 400 K, simulation[21]
BSO/MAPbI3/Cu2O−0.341 at 400 K, simulationThis work
BSO/MAPbI3/Spiro-OMeTAD−0.277 at 400 K, simulationThis work
BSO/MAPbI3/C−0.292 at 400 K, simulationThis work
BSO/MAFAPbI3/Cu2O−0.078 at 400 K, simulationThis work
BSO/MAFAPbI3/Spiro-OMeTAD+0.0285 at 320 K, −0.04 at 400 K, simulationThis work
BSO/MAFAPbI3/C−0.066 at 400 K, simulationThis work
TiO2/m-Al2O3/MAPbI3/C+2.5 × 10−2 (5 °C ≤ T ≤ 25 °C) and −1.8 × 10−2 (25 °C ≤ T ≤ 75 °C)[54]
PTAA/FA0.75Cs0.22MA0.03Pb(I0.82Br0.15Cl0.03)3−0.11 at 80 °C[55]
NiOx/FA0.79MA0.16Cs0.05Pb (I0.83,Br0.17)3−0.08 at 80 °C[55]
TiO2/FAMACsPb(I,Br)3/Spiro-OMeTAD /AuA single  T C P C E is not found at 50 °C[56]
Table 4. Optimal PCE of simulated devices under specific conditions compared to previous studies.
Table 4. Optimal PCE of simulated devices under specific conditions compared to previous studies.
ParametersResultReferences
Light intensity19.02% PCE (at 1000 Wm−2)[60]
38% PCE (at 10,000 Wm−2)This work
21% PCE (at 1000 Wm−2)[61]
32.04% PCE (at 1000 Wm−2)[21]
BSO thickness22.21% PCE (at 10 nm)[22]
~27% PCE (at 30 nm)[59]
31.24% PCE (at 100 nm)[21]
29% PCE (at 10 nm)This work
BSO donor concentration29.13% PCE (at 1021 cm−3)[62]
33% PCE (at 1021 cm−3)This work
31.24% PCE (at >1020 cm−3)[21]
~28% PCE (at 1020 cm−3)[59]
BSO/perovskite interface defect density18.59% PCE (at 1015 cm−2)[60]
27.6% PCE (at 1014 cm−2)This work
22.10% PCE (at 1014 cm−2)[22]
20.9% PCE (at 1010 cm−2)[63]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alkathran, N.; Bhandari, S.; Mallick, T.K. Theoretical Performance of BaSnO3-Based Perovskite Solar Cell Designs Under Variable Light Intensities, Temperatures, and Donor and Defect Densities. Designs 2025, 9, 76. https://doi.org/10.3390/designs9030076

AMA Style

Alkathran N, Bhandari S, Mallick TK. Theoretical Performance of BaSnO3-Based Perovskite Solar Cell Designs Under Variable Light Intensities, Temperatures, and Donor and Defect Densities. Designs. 2025; 9(3):76. https://doi.org/10.3390/designs9030076

Chicago/Turabian Style

Alkathran, Nouf, Shubhranshu Bhandari, and Tapas K. Mallick. 2025. "Theoretical Performance of BaSnO3-Based Perovskite Solar Cell Designs Under Variable Light Intensities, Temperatures, and Donor and Defect Densities" Designs 9, no. 3: 76. https://doi.org/10.3390/designs9030076

APA Style

Alkathran, N., Bhandari, S., & Mallick, T. K. (2025). Theoretical Performance of BaSnO3-Based Perovskite Solar Cell Designs Under Variable Light Intensities, Temperatures, and Donor and Defect Densities. Designs, 9(3), 76. https://doi.org/10.3390/designs9030076

Article Metrics

Back to TopTop