Next Article in Journal
Single and Bunch Soliton Generation in Optical Fiber Lasers Using Bismuth Selenide Topological Insulator Saturable Absorber
Next Article in Special Issue
Ultra-Low Loading of Gold on Nickel Foam for Nitrogen Electrochemistry
Previous Article in Journal
High-Performance Layered CaV4O9-MXene Composite Cathodes for Aqueous Zinc Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Device Simulation of Highly Stable and 29% Efficient FA0.75MA0.25Sn0.95Ge0.05I3-Based Perovskite Solar Cell

College of Engineering and Technology, American University of the Middle East, Egaila 54200, Kuwait
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(9), 1537; https://doi.org/10.3390/nano13091537
Submission received: 10 April 2023 / Revised: 27 April 2023 / Accepted: 29 April 2023 / Published: 3 May 2023

Abstract

:
A new type of perovskite solar cell based on mixed tin and germanium has the potential to achieve good power conversion efficiency and extreme air stability. However, improving its efficiency is crucial for practical application in solar cells. This paper presents a quantitative analysis of lead-free FA0.75MA0.25Sn0.95Ge0.05I3 using a solar cell capacitance simulator to optimize its structure. Various electron transport layer materials were thoroughly investigated to enhance efficiency. The study considered the impact of energy level alignment between the absorber and electron transport layer interface, thickness and doping concentration of the electron transport layer, thickness and defect density of the absorber, and the rear metal work function. The optimized structures included poly (3,4-ethylenedioxythiophene)polystyrene sulfonate (PEDOT:PSS) as the hole transport layer and either zinc oxide (ZnO) or zinc magnesium oxide (Zn0.7Mg0.3O) as the electron transport layer. The power conversion efficiency obtained was 29%, which was over three times higher than the initial structure. Performing numerical simulations on FA0.75MA0.25Sn0.95Ge0.05I3 can significantly enhance the likelihood of its commercialization. The optimized values resulting from the conducted parametric study are as follows: a short-circuit current density of 30.13 mA·cm−2), an open-circuit voltage of 1.08 V, a fill factor of 86.56%, and a power conversion efficiency of 28.31% for the intended solar cell.

1. Introduction

Renewable energy has become a crucial aspect of global energy production due to the increasing demand for clean and sustainable sources of energy. Research into solar cell technology is highly appealing and holds great promise, since sunlight is an unlimited and free resource that is both fundamentally renewable and environmentally friendly in contrast to finite fossil fuels [1,2]. In this context, the use of metal halide perovskites (MHP) in photovoltaics has seen tremendous growth in the past 10 years, with the recent power conversion efficiency (PCE) reaching above  25 % [3,4,5]. Despite this progress, the challenge lies in the fact that all the current MHP used to achieve high PCE contain lead, making it doubtful that this technology will be adopted on a large scale, especially in Europe, where strict regulations against the use of lead in electronics have been put in place [6].
As a result, there has been a growing interest in recent years to replace the lead component (Pb) in the perovskite formula  A P b X 3 with less toxic elements such as tin (Sn), bismuth (Bi), antimony (Sb), copper (Cu), or germanium (Ge) for both photovoltaic applications and crystal formation [7]. Of these alternatives, Sn-based perovskites [8,9] have been the most extensively studied and have demonstrated impressive  P C E of up to  13 % [10]. Sn-based perovskites have the advantage of smaller optical bandgaps [11,12,13] and greater charge mobility [14] compared to their Pb-based counterparts, making them ideal for single junction solar cells and all-perovskite tandem solar cells. Additionally, Sn is a naturally abundant element that does not present any environmental or health hazards. Perovskite solar cells (PSCs) based on Sn are widely recognized for their instability when exposed to the air due to the tendency of Sn to oxidize from a +2 to +4 state, which results in the creation of oxygen vacancies that can act as traps [2,15,16].
Another potential replacement for lead is Ge, a group 14 elements such as Sn and Pb. Germanium is a strong candidate for perovskite solar cells, as it has a higher electronegativity and more covalent character than lead [13]. Despite numerous theoretical studies suggesting the potential of germanium halide perovskites for solar cell applications [17,18,19,20], they have only rarely been studied experimentally due to their mercurial character in a +2 oxidation state [21]. To date, the  P C E of Ge-based PSC is still below 5% due to factors such as a smaller ionic radius, limited solubility in polar solvents, and a relatively wide bandgap (>1.6 eV) [18,22,23]. Although lead-free perovskites have demonstrated good efficiencies, there remains a persistent need to improve their stability and effectiveness [24,25,26,27], as their power conversion efficiency (PCE) values still fall considerably below the Shockley–Queisser efficiency limit of 33.7% [28] for a single junction. It is important to note that this limit has been exceeded by using nanoscale metallization in perovskite solar cells [28,29,30,31]. However, although metallization represents progress in perovskite technology, it is unlikely to significantly change the market situation for these cells. The primary challenge for perovskite cells is their poor durability and rapid degradation in the presence of oxygen and atmospheric water. To address the issue of oxidation in PSC, various techniques have been explored [16,32,33,34,35], including changing the electronic structure of the perovskite material, the encapsulation of PSC, incorporating hydrogen bonding, and applying a hydrophobic layer, among others [32,36].
One of the techniques for enhancing the performance of PSC involves mixing cations in tin–germanium Sn:Ge-based PSC. This has shown positive results in the literature [17,37,38,39]. By changing the Sn:Ge ratio, researchers have been able to achieve a lower bandgap and improved stability [40]. According to a study by Ito et al. [38], the efficiency of pure Sn-based perovskites improved from  3.31 % to  4.48 % (and even further to  6.90 % after 72 h) when  5 % germanium was doped into the material. The measurement was taken in the air without encapsulation. The addition of germanium is believed to have increased the stability of the perovskite structure while decreasing the trap density. This trend was also observed by Ng et al. [39], as they recorded the highest  P C E ( 7.9 % ) of Sn:Ge-based PSC to date. The efficiency of the Sn:Ge-based PSC is significantly lower than that of the Pb-based version due to low  V o c and  J s c . This is likely caused by poor absorption at the UV range, as noted in prior research [41,42,43], and issues with the energy level alignment at the interface electron transport layer (ETL)/perovskite [44,45].
Further improvement in the  P C E is still required, as the experimental outcomes have yet to reach the  25 %   P C E recorded by lead-based perovskite solar cells [44]. Studying the properties of the materials used in solar cells and controlling them through reliable simulation software can result in the creation of solar cells that are both highly efficient and cost-effective. PSC simulation is an interesting and straightforward process that can be carried out using various trustworthy programs such as SCAPS, AFORS-HET, Sentarus, and Silvaco [43,46,47,48,49].
In this contribution, we aim to enhance the efficiency of  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based solar cells by utilizing SCAPS software, developed by Gent University [50]. This particular perovskite has demonstrated a good PCE of 7.9% and impressive air stability in previous experiments [38,39]. This study presents a simple yet comprehensive simulation of the  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based PSC with a conventional (n-i-p) planar structure, which has not been previously conducted. The simulation mainly focuses on the use of metal oxide transport layers, particularly  Z n O and  Z n 0.7 M g 0.3 O . These materials have suitable electronic energies, high transparency, and uniform substrate coverage, making them excellent candidates for an ETL in the low-cost and large-scale production of lead-free PSC [51,52].
To attain maximum efficiency, an optimization process is carried out. First, a range of ETL materials are evaluated, and the most suitable ones are chosen. Then, their thicknesses and doping concentrations are optimized. Following this, the thickness of the perovskite absorber layer and its defect density are optimized. The effect of the rear metal work function on the photovoltaic performance of the device is analyzed subsequently. Finally, the results of the optimized structure are presented, demonstrating an improvement in efficiency of around  29 % .

2. Materials and Methods

The design and performance analysis of a solar cell were conducted using the SCAPS-1D software program. This numerical simulation tool was developed by researchers in the Department of Electronics and Information Systems (ELIS) at the University of Gent in Belgium [50]. The simulation program solves Poisson’s equation and the continuity equation for free electrons and holes in the conduction and valence bands. It enables the computation and observation of various electrical properties and parameters, such as the current density–voltage characteristics ( J V curve), the energy band structure of the heterojunction, quantum efficiency ( Q E ), open circuit voltage ( V o c ), short circuit current (Jsc), current density, power conversion efficiency ( P C E ), and fill factor ( F F ), among others. All simulations were conducted at a temperature of  300   K under the standard illumination of  1000   W / m 2 and an air mass of AM 1.5 G. The absorber layer was sandwiched between the hole transport layer (HTL) and ETL layers.
Figure 1 illustrates the proposed PSC structure with  f l u o r i n e d o p e d t i n o x i d e ( F T O ) / E T L / F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 / P E D O T : P S S / G o l d ( A u ) . The solar cell has a conventional structure (n-i-p), meaning that light enters the cell from the ETL side, with FTO acting as the front contact and Au as the back contact. For the HTL,  P E D O T : P S S is used in every structure. However, instead of using organic ETLs fullerene ( C 60 ) and [6,6]-phenyl-C(61)-butyric acid methyl ester ( P C B M ), as in the experimental work, this study investigated two ideal ETLs,  Z n O and  Z n 0.7 M g 0.3 O , as well as the conventional and extensively studied ETL titanium dioxide ( T i O 2 ). The photovoltaic performance of the cell was compared using three different metal oxides ( T i O 2 Z n O , and  Z n 0.7 M g 0.3 O ), alternately used as the ETL layer with the two organic ETLs  C 60 and  P C B M .
The energy level diagram of the perovskite with two organic ETLs and three inorganic metal oxide ETLs, along with other layers, is depicted in Figure 2.
Table 1, Table 2 and Table 3 provide a summary of the device and material parameters that were taken from theories, experiments, and the literature. The parameters listed in the table were considered while creating the initial setup for the simulation process. Various properties, such as the thickness and doping concentration of the ETL, the thickness and defect density of the absorber layer, and the rear metal work function, were adjusted to achieve the best possible outcome and to examine their impact on the device’s performance.
Aside from the earlier study that involved altering the ETL materials, various parameters, including the thickness and doping concentration of the ideal ETL layers and absorber, the defect density of the perovskite layer, and the rear metal work function of the cells, were adjusted to evaluate their effects on the device’s performance. The goal was to achieve the most effective cell structure through these modifications.

3. Results and Discussion

In this section, the study’s findings are presented, which started by examining the impact of different ETLs on the solar cell performance. After identifying the best structures based on this analysis, the research then explored various factors. These included optimizing the thickness and doping concentration of the ideal ETL, refining the absorber thickness, assessing the effect of the absorber layer’s defect density, and investigating how the solar cell’s performance was affected by the rear metal work function.

3.1. Impact of ETL Material on Solar Cell Performance

In planar PSC, the interface between the ETL and perovskite absorber layer plays a vital role in determining their overall performance [45]. To ensure high-quality ETL, several properties must be considered [44]. Firstly, the ETL should possess a suitable lowest unoccupied molecular orbital (LUMO) energy level that matches the conduction band energy of perovskite materials. Secondly, it should have high electron mobility and photochemical stability under solar irradiation. Lastly, it should be optically transparent to ensure maximum light absorption by the perovskite layer in the n-i-p PSC. Therefore, various ETLs, including,  C 60 P C B M T i O 2 Z n O , and  Z n 0.7 M g 0.3 O , are being examined to investigate how device performances differ with the uses of different ETLs. The electrical and optical parameters of the ETLs are listed in Table 2. Figure 3 depicts both the impact of the ETL material on the current density–voltage characteristics and its effect on the  P C E .
Table 4 lists the solar cell performance metrics, as well as the conduction band offset (CBO) values, for the simulated devices with five different ETLs. CBO refers to the difference in electron affinity between the absorber and the ETL (Equation (1)).
C B O = χ a b s o r b e r χ E T L
Figure 3 clearly shows that PSC with the organic ETL materials  C 60 and  P C B M have low performance, yielding  P C E values below  12 % . In contrast, all the structures with inorganic ETL materials produce  P C E values above  14 % . Among the structures simulated, those incorporating  Z n O ETL and  Z n 0.7 M g 0.3 O ETL materials are the most efficient, achieving  P C E values of  20.88 % and  21.55 % , respectively.
Devices incorporating organic ETLs and  T i O 2 ETL exhibit significantly lower PCEs than PSCs utilizing  Z n O ETL and  Z n M g O ETL, primarily due to their lower  V o c . The low  V o c is likely attributable to the band alignment present within their structure. Table 4 reveals that, as the CBO becomes increasingly negative, the  V o c value decreases correspondingly. PSCs incorporating  Z n O ETL or  Z n 0.7 M g 0.3 O ETL, on the other hand, exhibit CBO values close to zero or even positive, which accounts for their comparatively higher  V o c values. When the conduction band minimum (CBM) of the ETL is located below that of the absorber, it results in a negative CBO, and a cliff-like structure forms at the heterojunction ETL/absorber. In solar cells, this cliff structure is detrimental, since it promotes the accumulation of electrons and holes near the interface following charge separation, leading to greater charge recombination via the interface’s deep-level defects, which results in lower  V o c .
Alongside the band alignment within the structure of the PSC, another factor that could contribute to the difference in behavior between the devices is the ETL’s bandgap. A suitable ETL material must possess optical transparency to ensure maximum light absorption by the perovskite layer in the n-i-p PSC, which leads to the generation of more electrons and ultimately results in a higher  J s c achieved by the cell. This claim is supported by Figure 3 and Table 4. PSCs with organic ETLs, which have bandgap values of  1.7   e V and  2   e V , respectively, demonstrate the lowest values of  J s c , specifically  21.97   m A · c m 2 and  24.64   m A · c m 2 . Conversely, PSCs with inorganic metal oxide ETLs score the highest values of  J s c , above  27   m A · c m 2 , as they have bandgap values above  2.8   e V .
Figure 4 endorses this observation, illustrating the quantum efficiency of PSCs with various ETLs. PSCs with organic ETLs have a notably low quantum efficiency, particularly for wavelengths below  700   n m , attributable to their below  2   e V bandgap values. On the other hand, PSCs with inorganic metal oxide ETLs exhibit the highest quantum efficiency, since their bandgap values exceed  2.8   e V .
The results presented in this section demonstrate that  Z n O and  Z n 0.7 M g 0.3 O outperform the other tested ETLs. Consequently, the study investigates how the thickness and doping concentration of these ETL materials affect the performances of solar cell devices.

3.1.1. Impact of ETL Thickness

In the previous analysis, we compared the ETL materials while keeping the layer thickness constant at  50   n m . However, in this section, we study the impact of varying the thickness of the ETL on the solar cell’s performance. We varied the ETL thickness from  50   n m to  200   n m and analyzed the photovoltaic performances. Figure 5 shows the results as a function of the ETL thickness. The  V o c and  P C E of both devices remain constant and independent of an ETL thickness up to around  160   n m . However, they decrease significantly as the thickness increases further. The increase in thickness causes electrons to travel a longer distance to reach the top electrode, resulting in a higher likelihood of electron recombination with minority carriers (holes). This, in turn, causes the  V o c to decline sharply. Additionally, the cell with a  Z n O ETL shows a noticeable decrease in  J s c , likely due to a decrease in light transmittance through the  Z n O layer.
This decrease in light transmittance is confirmed by Figure 6, which shows the quantum efficiency of the cells as a function of the ETL thickness. The QE and the  J s c of the cell with a  Z n _ 0.7 M g _ 0.3 O ETL remain unchanged due to the wide bandgap of  Z n 0.7 M g 0.3 O ( 4.1   e V ).
It is clear that the photovoltaic parameters deteriorate as the ETL thickness increases, resulting in a decrease in PCE for both ETLs. This is caused by inefficient charge carrier transport to the electrodes, an increase in series resistance that reduces the fill factor FF, and a higher probability of recombination as the ETL thickness increases. Therefore, for the remainder of this study, a thickness of 50 nm is adopted. Any thinner layer may not fully cover the perovskite layer, causing direct contact between FTO and perovskite, which leads to carrier recombination and reduced hole-blocking efficiency [63].

3.1.2. Impact of ETL Doping Concentration

In addition to identifying the most suitable ETL materials, namely  Z n O and  Z n 0.7 M g 0.3 O , and optimizing their thickness at 50 nm, it is important to consider the impact of the doping concentration  N D on the photovoltaic parameters of PSCs. In the previous sections, a fixed doping concentration of  N D = 1 × 10 17   c m 3 was used for all materials tested. However, this section presents a study on the effect of varying the  N D from  1 × 10 15   c m 3 to  1 × 10 20   c m 3 for the current density–voltage characteristics and power conversion efficiency (PCE), as shown in Figure 7.
The results shown in Figure 7 clearly demonstrate that increasing the doping concentration in the ETL significantly enhances the  V o c and  F F of the cells, resulting in a higher PCE. Notably, the  J s c of both devices remains almost constant, as it has already reached a high value of approximately  28   m A · c m 2 , and the effect of doping on  J s c is negligible.
Both devices achieved a peak  P C E of  22 % at doping concentrations of  1 × 10 19   c m 3 and  1 × 10 20   c m 3 , with little difference between them. The improvement in  V o c and  F F can be attributed to the effect of doping on the energy level alignment between the ETL and the perovskite layer, which enhances the charge transport properties and reduces the recombination losses.
Although increasing the doping concentration of the ETL improves the  V o c F F , and overall efficiency of the PSC, the optimal doping concentration has been determined to be  1 × 10 19   c m 3 due to practical manufacturing challenges. Higher doping concentrations are difficult to achieve practically and could potentially create deep Coulomb traps, which may adversely affect carrier mobility [64].

3.2. Effect of the Perovskite Layer on the Solar Cell Performance

In addition to the crucial role of ETL materials in improving the PSC performance, the absorber layer also has a significant impact on the efficiency of the solar cell. This section will examine how the absorber material  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 affects the solar cell performance, with a specific focus on the thickness and defect density of this absorber.

3.2.1. Impact of Absorber Thickness

The thickness of the absorber layer in perovskite solar cells can significantly impact the device’s performance, as it determines the amount of light absorption and the efficiency of the conversion process. It is crucial to maintain an optimal thickness range, because if the absorber layer is too thin, it may not absorb sufficient light to generate enough current. Conversely, if the absorber layer is too thick, the charge carriers generated by the absorbed light may struggle to travel through the material and reach the electrodes, leading to lower device efficiency. The previous analyses were performed using a  400   n m thick  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 . In this section, the impact of the absorber thickness on the solar cell’s performance was studied by varying the thickness from  200   n m to  1500   n m . The results obtained for the photovoltaic outputs are shown in Figure 8, while Figure 9 displays the effect of the absorber thickness on the quantum efficiency with respect to the wavelengths of the light.
By observing Figure 8 and Figure 9, it is evident that the behavior of all photovoltaic parameters and the quantum efficiency as a function the of absorber thickness is similar for both devices, with  Z n O ETL and  Z n 0.7 M g 0.3 O ETL. The  P C E of both devices steadily increases with the increasing thickness until it reaches a maximum value of  22.9 % at  600   n m , beyond which it gradually decreases. This can be attributed to the opposing trends of the  V o c and the  J s c , which have the most significant impact on the  P C E .
Firstly, a considerable increase in  J s c was observed in both devices by increasing the thickness of the absorber layer, but it reached saturation at  700   n m . Beyond this point, the effect of the absorber layer thickness became insignificant. This substantial enhancement of  J s c is attributed to the generation of additional electron–hole pairs in the perovskite, which occurred as a result of increased light absorption resulting from the thicker absorber layer. The increase in  J s c is supported by the higher QE of the device at larger thicknesses, as shown in Figure 9.
In contrast, the open-circuit voltage  V o c decreases as the absorber layer thickness increases. While a thicker layer allows for more photons to be absorbed and more electron–hole pairs to be generated, it also leads to a high density of defects that act as recombination centers. As a result, the lifetime of electron–hole pairs is reduced, and more pairs recombine before reaching the electrodes, causing a decline in the  V o c . In addition, the series resistance of the device increases with thicker absorber layers, further lowering the  V o c and the FF. Consequently, increasing the absorber layer thickness beyond  600   n m results in diminishing returns and decreases the overall efficiency of the solar cell.

3.2.2. Impact of Absorber Defect Density  N t

Although adjusting the thickness of the absorber has improved the efficiency of solar cells, further enhancements in the solar cell performance can be achieved by considering the defect density of the perovskite layer as an additional influential parameter.
The initial defect density  N t of the absorber was set at  1 × 10 16   c m 3 , which is the same as the value obtained in the experiment conducted by Ng et al. [39] on  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 . Recent experimental studies by Zheng et al. [65] and Chen et al. [66] showed that the defect density in perovskite can be as low as  1 × 10 11   c m 3 and  1 × 10 12   c m 3 , respectively. In our simulation study, we varied the defect density between  1 × 10 12   c m 3 and  1 × 10 16   c m 3 and plotted the changes in the photovoltaic properties with  N t for devices using  Z n O ETL and  Z n 0.7 M g 0.3 O ETL in Figure 10.
The  J s c of both devices remain constant, while  V o c and  F F are greatly increased when the defect density in perovskite is reduced, resulting in a significant enhancement of the  P C E . When the defect density reaches a low level of  1 × 10 13   c m 3 , both cells show a significant improvement in performance. The cells with  Z n O ETL and  Z n 0.7 M g 0.3 O ETL exhibit a  J s c of  30.11   m A · c m 2 and  30.11   m A · c m 2 V o c of  1.087   V and  1.087   V , FF of  86.46 % and  86.74 % , and  P C E of  28.31 % and  28.26 % , respectively. However, further reducing the  N t from  1 × 10 13   c m 3 to  1 × 10 12   c m 3 only leads to a slight improvement in the cell performance. Hence, an absorber defect density  N t = 1 × 10 13   c m 3 is adopted for the rest of this study.
The impact of perovskite defect density on the device performance can be explained by the Shockley–Read–Hall (SRH) recombination model [64,67]. To gain a better understanding of this effect, the relationship between the SRH recombination rate and depth from the surface for various defect densities was examined. Figure 11 illustrates these findings. The results indicate that recombination becomes more significant in the light-absorbing layer due to the higher defect density resulting from the low film quality. The defects in the absorber layer act as recombination centers for the electron–hole pairs generated by incident photons, which reduces the efficiency of the device. The defect density in the absorber layer can be influenced by a variety of factors, such as the synthesis process, the perovskite composition, and the deposition method. Therefore, greater efforts should be made to improve the fabrication technique of solar cells.

3.3. Impact of Rear Electrode Work Function

The work function of the rear metal in a PSC plays a critical role in determining the energy alignment at the interface between the HTL and the rear contact electrode, which affects the built-in potential  V b i and the charge carrier extraction. In this study, the rear electrode in the PSC was initially made of gold (Au), which is a common choice for metal back contact. However, other materials such as aluminum (Al), silver (Ag), chromium (Cr), nickel (Ni), palladium (Pd), and platinum (Pt), with different work functions ranging from 4.2 to 5.7 eV, are also used in PSCs and optoelectronic devices. Table 5 presents the work functions of several chosen metals in the field [68].
Our study aimed to examine how the rear metal work function affects the photovoltaic properties of the devices. The results are depicted in Figure 12.
As the work function of the anode increases, both the  V o c and  F F of the solar cell increase, resulting in greater efficiency, until they reach a maximum and plateau at 5 eV and above. This is because a decrease in the metal work function reduces the built-in electric field in the absorber layer, causing a poor collection of photo-generated carriers and resulting in lower  V o c and  F F .
In cases where the anode’s work function is lower than that of PEDOT:PSS ( 5.0   e V ) ( 5.0   e V ) [69], with metals such as Ag, Cu, and Au, a rectifying Schottky barrier contact is created at the anode–PEDOT interface. This contact acts as an obstacle to the movement of holes to the anode, thereby decreasing the  V o c F F , and  P C E , as illustrated in Figure 12. On the other hand, when using an anode made of Au, Ni, Pd, or Pt, which have a higher work function than PEDOT, an ohmic contact is established at the anode–PEDOT:PSS interface. This enables efficient hole transport across the interface, resulting in higher  V o c F F , and  P C E values in the PSC. Therefore, selecting one of these anodes is crucial when manufacturing the solar cell device.

4. Conclusions

Despite conducting extensive experimental studies on  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based PSCs, the highest achieved  P C E remained below  8 % , which falls short of the desired benchmark for commercial applications. Our work utilized SCAPS-1D software to simulate a conventional (n-i-p) structure and systematically compared the effectiveness of various ETL materials. Specifically, we tested two organic ETLs and three inorganic metal oxide ETLs while keeping the other layers unchanged. Our findings showed that selecting appropriate ETL materials could significantly increase the  P C E of the cell to  21 % . The  Z n O and  Z n 0.7 M g 0.3 O ETLs were found to be the most effective ETL materials due to their excellent band alignments with the absorber and wide bandgaps. The efficiency of the solar cell was further improved by increasing the doping concentration of the ETL and the absorber thickness, reducing the absorber defect density, and selecting gold or any metal with a work function greater than  5.1   e V . These enhancements led to an unprecedented  P C E of almost  29 % . It is crucial to be cautious while interpreting these findings, because they may not accurately represent the experimental efforts, which were only able to achieve a maximum PCE of less than 8%. Future research should focus on refining the device fabrication techniques, as our novel results could provide a feasible approach to develop cost-effective, highly efficient, and stable  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based PSCs.

Author Contributions

Conceptualization, H.S.; methodology, H.S.; software, H.S.; validation, H.S.; formal analysis, H.S.; investigation, H.S.; resources, H.S.; data curation, H.S.; writing—original draft preparation, H.S.; writing—review and editing, H.S. and Z.A.B.; visualization, H.S. and Z.A.B.; supervision, H.S.; and project administration, H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author would like to acknowledge Marc Burgelman (University of Gent) for providing the simulation software SCAPS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Z.; Guo, Y.; Li, H.; Zhou, Y.; Zuo, X.; Yu, Y.; Pan, C.; Strzalka, J.; Nam, C.-Y.; Rafailovich, M.H. Roles of Interfacial Tension in Regulating Internal Organization of Low Bandgap Polymer Bulk Heterojunction Solar Cells by Polymer Additives. Adv. Mater. Interfaces 2018, 5, 1800435. [Google Scholar] [CrossRef]
  2. Noel, N.K.; Stranks, S.D.; Abate, A.; Wehrenfennig, C.; Guarnera, S.; Haghighirad, A.-A.; Sadhanala, A.; Eperon, G.E.; Pathak, S.K.; Johnston, M.B.; et al. Lead-Free Organic–Inorganic Tin Halide Perovskites for Photovoltaic Applications. Energy Environ. Sci. 2014, 7, 3061–3068. [Google Scholar] [CrossRef]
  3. Kim, J.Y.; Lee, J.-W.; Jung, H.S.; Shin, H.; Park, N.-G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918. [Google Scholar] [CrossRef] [PubMed]
  4. Liang, F.-C.; Jhuang, F.-C.; Fang, Y.-H.; Benas, J.-S.; Chen, W.-C.; Yan, Z.-L.; Lin, W.-C.; Su, C.-J.; Sato, Y.; Chiba, T.; et al. Synergistic Effect of Cation Composition Engineering of Hybrid Cs1−xFAxPbBr3 Nanocrystals for Self-Healing Electronics Application. Adv. Mater. 2023, 35, 2207617. [Google Scholar] [CrossRef] [PubMed]
  5. Ma, C.; Park, N.-G. A Realistic Methodology for 30% Efficient Perovskite Solar Cells. Chem 2020, 6, 1254–1264. [Google Scholar] [CrossRef]
  6. Abate, A. Perovskite Solar Cells Go Lead Free. Joule 2017, 1, 659–664. [Google Scholar] [CrossRef]
  7. Mahajan, P.; Datt, R.; Chung Tsoi, W.; Gupta, V.; Tomar, A.; Arya, S. Recent Progress, Fabrication Challenges and Stability Issues of Lead-Free Tin-Based Perovskite Thin Films in the Field of Photovoltaics. Coord. Chem. Rev. 2021, 429, 213633. [Google Scholar] [CrossRef]
  8. Sabbah, H.; Arayro, J.; Mezher, R. Numerical Simulation and Optimization of Highly Stable and Efficient Lead-Free Perovskite FA1−xCsxSnI3-Based Solar Cells Using SCAPS. Materials 2022, 15, 4761. [Google Scholar] [CrossRef]
  9. Xu, K. Development of Tin-Based Perovskite Materials for Solar Cell Applications: A Minireview. Instrum. Sci. Technol. 2021, 49, 91–105. [Google Scholar] [CrossRef]
  10. Cao, J.; Yan, F. Recent Progress in Tin-Based Perovskite Solar Cells. Energy Environ. Sci. 2021, 14, 1286–1325. [Google Scholar] [CrossRef]
  11. Ke, W.; Kanatzidis, M.G. Prospects for Low-Toxicity Lead-Free Perovskite Solar Cells. Nat. Commun. 2019, 10, 965. [Google Scholar] [CrossRef]
  12. Stoumpos, C.C.; Malliakas, C.D.; Kanatzidis, M.G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019–9038. [Google Scholar] [CrossRef] [PubMed]
  13. Marshall, K.P.; Walker, M.; Walton, R.I.; Hatton, R.A. Enhanced Stability and Efficiency in Hole-Transport-Layer-Free CsSnI3 Perovskite Photovoltaics. Nat. Energy 2016, 1, 16178. [Google Scholar] [CrossRef]
  14. Herz, L.M. Charge-Carrier Mobilities in Metal Halide Perovskites: Fundamental Mechanisms and Limits. ACS Energy Lett. 2017, 2, 1539–1548. [Google Scholar] [CrossRef]
  15. Rajendra Kumar, G.; Kim, H.-J.; Karupannan, S.; Prabakar, K. Interplay between Iodide and Tin Vacancies in CsSnI3 Perovskite Solar Cells. J. Phys. Chem. C 2017, 121, 16447–16453. [Google Scholar] [CrossRef]
  16. Liao, Y.; Liu, H.; Zhou, W.; Yang, D.; Shang, Y.; Shi, Z.; Li, B.; Jiang, X.; Zhang, L.; Quan, L.N.; et al. Highly Oriented Low-Dimensional Tin Halide Perovskites with Enhanced Stability and Photovoltaic Performance. J. Am. Chem. Soc. 2017, 139, 6693–6699. [Google Scholar] [CrossRef]
  17. Ju, M.-G.; Dai, J.; Ma, L.; Zeng, X.C. Lead-Free Mixed Tin and Germanium Perovskites for Photovoltaic Application. J. Am. Chem. Soc. 2017, 139, 8038–8043. [Google Scholar] [CrossRef]
  18. Stoumpos, C.C.; Frazer, L.; Clark, D.J.; Kim, Y.S.; Rhim, S.H.; Freeman, A.J.; Ketterson, J.B.; Jang, J.I.; Kanatzidis, M.G. Hybrid Germanium Iodide Perovskite Semiconductors: Active Lone Pairs, Structural Distortions, Direct and Indirect Energy Gaps, and Strong Nonlinear Optical Properties. J. Am. Chem. Soc. 2015, 137, 6804–6819. [Google Scholar] [CrossRef] [PubMed]
  19. Ming, W.; Shi, H.; Du, M.-H. Large Dielectric Constant, High Acceptor Density, and Deep Electron Traps in Perovskite Solar Cell Material CsGeI3. J. Mater. Chem. A 2016, 4, 13852–13858. [Google Scholar] [CrossRef]
  20. Sun, P.-P.; Li, Q.-S.; Yang, L.-N.; Li, Z.-S. Theoretical Insights into a Potential Lead-Free Hybrid Perovskite: Substituting Pb2+ with Ge2+. Nanoscale 2016, 8, 1503–1512. [Google Scholar] [CrossRef]
  21. Ansari, M.I.H.; Qurashi, A.; Nazeeruddin, M.K. Frontiers, Opportunities, and Challenges in Perovskite Solar Cells: A Critical Review. J. Photochem. Photobiol. C Photochem. Rev. 2018, 35, 1–24. [Google Scholar] [CrossRef]
  22. Gao, P.; Bin Mohd Yusoff, A.R.; Nazeeruddin, M.K. Dimensionality Engineering of Hybrid Halide Perovskite Light Absorbers. Nat. Commun. 2018, 9, 5028. [Google Scholar] [CrossRef]
  23. Krishnamoorthy, T.; Ding, H.; Yan, C.; Leong, W.L.; Baikie, T.; Zhang, Z.; Sherburne, M.; Li, S.; Asta, M.; Mathews, N.; et al. Lead-Free Germanium Iodide Perovskite Materials for Photovoltaic Applications. J. Mater. Chem. A 2015, 3, 23829–23832. [Google Scholar] [CrossRef]
  24. Mei, A.; Li, X.; Liu, L.; Ku, Z.; Liu, T.; Rong, Y.; Xu, M.; Hu, M.; Chen, J.; Yang, Y.; et al. A Hole-Conductor–Free, Fully Printable Mesoscopic Perovskite Solar Cell with High Stability. Science 2014, 345, 295–298. [Google Scholar] [CrossRef]
  25. Arya, S.; Mahajan, P.; Gupta, R.; Srivastava, R.; Tailor, N.K.; Satapathi, S.; Sumathi, R.R.; Datt, R.; Gupta, V. A Comprehensive Review on Synthesis and Applications of Single Crystal Perovskite Halides. Prog. Solid State Chem. 2020, 60, 100286. [Google Scholar] [CrossRef]
  26. Grätzel, M. The Light and Shade of Perovskite Solar Cells. Nat. Mater 2014, 13, 838–842. [Google Scholar] [CrossRef] [PubMed]
  27. Boix, P.P.; Agarwala, S.; Koh, T.M.; Mathews, N.; Mhaisalkar, S.G. Perovskite Solar Cells: Beyond Methylammonium Lead Iodide. J. Phys. Chem. Lett. 2015, 6, 898–907. [Google Scholar] [CrossRef]
  28. Jacak, J.E.; Jacak, W.A. Routes for Metallization of Perovskite Solar Cells. Materials 2022, 15, 2254. [Google Scholar] [CrossRef] [PubMed]
  29. Wu, R.; Yang, B.; Zhang, C.; Huang, Y.; Cui, Y.; Liu, P.; Zhou, C.; Hao, Y.; Gao, Y.; Yang, J. Prominent Efficiency Enhancement in Perovskite Solar Cells Employing Silica-Coated Gold Nanorods. J. Phys. Chem. C 2016, 120, 6996–7004. [Google Scholar] [CrossRef]
  30. Laska, M.; Krzemińska, Z.; Kluczyk-Korch, K.; Schaadt, D.; Popko, E.; Jacak, W.A.; Jacak, J.E. Metallization of Solar Cells, Exciton Channel of Plasmon Photovoltaic Effect in Perovskite Cells. Nano Energy 2020, 75, 104751. [Google Scholar] [CrossRef]
  31. Yao, K.; Zhong, H.; Liu, Z.; Xiong, M.; Leng, S.; Zhang, J.; Xu, Y.; Wang, W.; Zhou, L.; Huang, H.; et al. Plasmonic Metal Nanoparticles with Core–Bishell Structure for High-Performance Organic and Perovskite Solar Cells. ACS Nano 2019, 13, 5397–5409. [Google Scholar] [CrossRef]
  32. Shao, S.; Liu, J.; Portale, G.; Fang, H.-H.; Blake, G.R.; ten Brink, G.H.; Koster, L.J.A.; Loi, M.A. Highly Reproducible Sn-Based Hybrid Perovskite Solar Cells with 9% Efficiency. Adv. Energy Mater. 2018, 8, 1702019. [Google Scholar] [CrossRef]
  33. Lee, S.J.; Shin, S.S.; Kim, Y.C.; Kim, D.; Ahn, T.K.; Noh, J.H.; Seo, J.; Seok, S.I. Fabrication of Efficient Formamidinium Tin Iodide Perovskite Solar Cells through SnF2–Pyrazine Complex. J. Am. Chem. Soc. 2016, 138, 3974–3977. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, K.; Wu, P.; Yang, W.; Su, R.; Luo, D.; Yang, X.; Tu, Y.; Zhu, R.; Gong, Q. Low-Dimensional Perovskite Interlayer for Highly Efficient Lead-Free Formamidinium Tin Iodide Perovskite Solar Cells. Nano Energy 2018, 49, 411–418. [Google Scholar] [CrossRef]
  35. Ke, W.; Stoumpos, C.C.; Zhu, M.; Mao, L.; Spanopoulos, I.; Liu, J.; Kontsevoi, O.Y.; Chen, M.; Sarma, D.; Zhang, Y.; et al. Enhanced Photovoltaic Performance and Stability with a New Type of Hollow 3D Perovskite {en}FASnI3. Sci. Adv. 2017, 3, e1701293. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, F.; Ma, J.; Xie, F.; Li, L.; Chen, J.; Fan, J.; Zhao, N. Organic Cation-Dependent Degradation Mechanism of Organotin Halide Perovskites. Adv. Funct. Mater. 2016, 26, 3417–3423. [Google Scholar] [CrossRef]
  37. Sabbah, H. Numerical Simulation of 30% Efficient Lead-Free Perovskite CsSnGeI3-Based Solar Cells. Materials 2022, 15, 3229. [Google Scholar] [CrossRef]
  38. Ito, N.; Kamarudin, M.A.; Hirotani, D.; Zhang, Y.; Shen, Q.; Ogomi, Y.; Iikubo, S.; Minemoto, T.; Yoshino, K.; Hayase, S. Mixed Sn–Ge Perovskite for Enhanced Perovskite Solar Cell Performance in Air. J. Phys. Chem. Lett. 2018, 9, 1682–1688. [Google Scholar] [CrossRef]
  39. Ng, C.H.; Nishimura, K.; Ito, N.; Hamada, K.; Hirotani, D.; Wang, Z.; Yang, F.; Likubo, S.; Shen, Q.; Yoshino, K.; et al. Role of GeI2 and SnF2 Additives for SnGe Perovskite Solar Cells. Nano Energy 2019, 58, 130–137. [Google Scholar] [CrossRef]
  40. Chen, M.; Ju, M.-G.; Garces, H.F.; Carl, A.D.; Ono, L.K.; Hawash, Z.; Zhang, Y.; Shen, T.; Qi, Y.; Grimm, R.L.; et al. Highly Stable and Efficient All-Inorganic Lead-Free Perovskite Solar Cells with Native-Oxide Passivation. Nat. Commun. 2019, 10, 16. [Google Scholar] [CrossRef]
  41. Baig, F.; Khattak, Y.H.; Marí, B.; Beg, S.; Ahmed, A.; Khan, K. Efficiency Enhancement of CH3NH3SnI3 Solar Cells by Device Modeling. J. Electron. Mater. 2018, 47, 5275–5282. [Google Scholar] [CrossRef]
  42. Du, H.-J.; Wang, W.-C.; Zhu, J.-Z. Device Simulation of Lead-Free CH3NH3SnI3 Perovskite Solar Cells with High Efficiency. Chin. Phys. B 2016, 25, 108802. [Google Scholar] [CrossRef]
  43. Sabbah, H.; Arayro, J.; Mezher, R. Simulation and Investigation of 26% Efficient and Robust Inverted Planar Perovskite Solar Cells Based on GA0.2FA0.78SnI3-1%EDAI2 Films. Nanomaterials 2022, 12, 3885. [Google Scholar] [CrossRef]
  44. Pan, H.; Zhao, X.; Gong, X.; Li, H.; Ladi, N.H.; Zhang, X.L.; Huang, W.; Ahmad, S.; Ding, L.; Shen, Y.; et al. Advances in Design Engineering and Merits of Electron Transporting Layers in Perovskite Solar Cells. Mater. Horiz. 2020, 7, 2276–2291. [Google Scholar] [CrossRef]
  45. Ding, C.; Zhang, Y.; Liu, F.; Kitabatake, Y.; Hayase, S.; Toyoda, T.; Yoshino, K.; Minemoto, T.; Katayama, K.; Shen, Q. Effect of the Conduction Band Offset on Interfacial Recombination Behavior of the Planar Perovskite Solar Cells. Nano Energy 2018, 53, 17–26. [Google Scholar] [CrossRef]
  46. Minemoto, T.; Kawano, Y.; Nishimura, T.; Chantana, J. Numerical Reproduction of a Perovskite Solar Cell by Device Simulation Considering Band Gap Grading. Opt. Mater. 2019, 92, 60–66. [Google Scholar] [CrossRef]
  47. Gagandeep; Singh, M.; Kumar, R. Simulation of Perovskite Solar Cell with Graphene as Hole Transporting Material. AIP Conf. Proc. 2019, 2115, 030548. [Google Scholar] [CrossRef]
  48. Nandal, V.; Nair, P.R. Predictive Modeling of Ion Migration Induced Degradation in Perovskite Solar Cells. ACS Nano 2017, 11, 11505–11512. [Google Scholar] [CrossRef] [PubMed]
  49. Du, Y.; Cai, H.; Wen, H.; Wu, Y.; Li, Z.; Xu, J.; Huang, L.; Ni, J.; Li, J.; Zhang, J. Revealing the Unfavorable Role of Superfluous CH3NH3PbI3 Grain Boundary Traps in Perovskite Solar Cells on Carrier Collection. RSC Adv. 2016, 6, 83264–83272. [Google Scholar] [CrossRef]
  50. Burgelman, M.; Nollet, P.; Degrave, S. Modelling Polycrystalline Semiconductor Solar Cells. Thin Solid Film. 2000, 361–362, 527–532. [Google Scholar] [CrossRef]
  51. Gamal, N.; Sedky, S.H.; Shaker, A.; Fedawy, M. Design of Lead-Free Perovskite Solar Cell Using Zn1−xMgxO as ETL: SCAPS Device Simulation. Optik 2021, 242, 167306. [Google Scholar] [CrossRef]
  52. Rahman, M.S.; Miah, S.; Marma, M.S.W.; Sabrina, T. Simulation Based Investigation of Inverted Planar Perovskite Solar Cell with All Metal Oxide Inorganic Transport Layers. In Proceedings of the 2019 International Conference on Electrical, Computer and Communication Engineering (ECCE), Cox’sBazar, Bangladesh, 7–9 February 2019; pp. 1–6. [Google Scholar]
  53. Dang, Y.; Zhou, Y.; Liu, X.; Ju, D.; Xia, S.; Xia, H.; Tao, X. Formation of Hybrid Perovskite Tin Iodide Single Crystals by Top-Seeded Solution Growth. Angew. Chem. Int. Ed. 2016, 55, 3447–3450. [Google Scholar] [CrossRef] [PubMed]
  54. Jokar, E.; Chien, C.-H.; Tsai, C.-M.; Fathi, A.; Diau, E.W.-G. Robust Tin-Based Perovskite Solar Cells with Hybrid Organic Cations to Attain Efficiency Approaching 10%. Adv. Mater. 2019, 31, 1804835. [Google Scholar] [CrossRef]
  55. Liu, X.; Yan, K.; Tan, D.; Liang, X.; Zhang, H.; Huang, W. Solvent Engineering Improves Efficiency of Lead-Free Tin-Based Hybrid Perovskite Solar Cells beyond 9%. ACS Energy Lett. 2018, 3, 2701–2707. [Google Scholar] [CrossRef]
  56. Minemoto, T.; Kawano, Y.; Nishimura, T.; Shen, Q.; Yoshino, K.; Iikubo, S.; Hayase, S.; Chantana, J. Theoretical Analysis of Band Alignment at Back Junction in Sn–Ge Perovskite Solar Cells with Inverted p-i-n Structure. Sol. Energy Mater. Sol. Cells 2020, 206, 110268. [Google Scholar] [CrossRef]
  57. Singh, N.; Agarwal, A.; Agarwal, M. Numerical Simulation of Highly Efficient Lead-Free All-Perovskite Tandem Solar Cell. Sol. Energy 2020, 208, 399–410. [Google Scholar] [CrossRef]
  58. Stamate, M.D. On the Dielectric Properties of Dc Magnetron TiO2 Thin Films. Appl. Surf. Sci. 2003, 218, 318–323. [Google Scholar] [CrossRef]
  59. Doroody, C.; Rahman, K.S.; Rosly, H.N.; Harif, M.N.; Haque, F.; Tiong, S.K.; Amin, N. Impact of High Resistivity Transparent (HRT) Layer in Cadmium Telluride Solar Cells from Numerical Simulation. J. Renew. Sustain. Energy 2020, 12, 023702. [Google Scholar] [CrossRef]
  60. Hoye, R.L.Z.; Ehrler, B.; Böhm, M.L.; Muñoz-Rojas, D.; Altamimi, R.M.; Alyamani, A.Y.; Vaynzof, Y.; Sadhanala, A.; Ercolano, G.; Greenham, N.C.; et al. Improved Open- Circuit Voltage in ZnO–PbSe Quantum Dot Solar Cells by Understanding and Reducing Losses Arising from the ZnO Conduction Band Tail. Adv. Energy Mater. 2014, 4, 1301544. [Google Scholar] [CrossRef] [PubMed]
  61. Chantana, J.; Kato, T.; Sugimoto, H.; Minemoto, T. Structures of Cu(In,Ga)(S,Se)2 Solar Cells for Minimizing Open-circuit Voltage Deficit: Investigation of Carrier Recombination Rates. Prog. Photovolt. Res. Appl. 2019, 27, 630–639. [Google Scholar] [CrossRef]
  62. He, X.; Wu, L.; Hao, X.; Zhang, J.; Li, C.; Wang, W.; Feng, L.; Du, Z. The Band Structures of Zn1−xMgxO(In) and the Simulation of CdTe Solar Cells with a Zn1−xMgxO(In) Window Layer by SCAPS. Energies 2019, 12, 291. [Google Scholar] [CrossRef]
  63. Onyeagba, C.R.; Islam, M.; Yarlagadda, P.K.D.V.; Tesfamichael, T. Investigating the Properties of Tin-Oxide Thin Film Developed by Sputtering Process for Perovskite Solar Cells. Mater. Renew. Sustain. Energy 2023, 12, 31–37. [Google Scholar] [CrossRef]
  64. 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]
  65. Zheng, X.; Chen, B.; Dai, J.; Fang, Y.; Bai, Y.; Lin, Y.; Wei, H.; Zeng, X.C.; Huang, J. Defect Passivation in Hybrid Perovskite Solar Cells Using Quaternary Ammonium Halide Anions and Cations. Nat. Energy 2017, 2, 17102. [Google Scholar] [CrossRef]
  66. Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhang, W.; Yang, X.; Chen, H.; Bi, E.; Ashraful, I.; Grätzel, M.; et al. Efficient and Stable Large-Area Perovskite Solar Cells with Inorganic Charge Extraction Layers. Science 2015, 350, 944–948. [Google Scholar] [CrossRef] [PubMed]
  67. Haider, S.Z.; Anwar, H.; Wang, M. A Comprehensive Device Modelling of Perovskite Solar Cell with Inorganic Copper Iodide as Hole Transport Material. Semicond. Sci. Technol. 2018, 33, 035001. [Google Scholar] [CrossRef]
  68. Michaelson, H.B. The Work Function of the Elements and Its Periodicity. J. Appl. Phys. 1977, 48, 4729–4733. [Google Scholar] [CrossRef]
  69. Mengistie, D.A.; Ibrahem, M.A.; Wang, P.-C.; Chu, C.-W. Highly Conductive PEDOT:PSS Treated with Formic Acid for ITO-Free Polymer Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 2292–2299. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a PSC based on mixed Sn:Ge  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 .
Figure 1. Schematic diagram of a PSC based on mixed Sn:Ge  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 .
Nanomaterials 13 01537 g001
Figure 2. Band alignment between the ETL materials and  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 perovskite.
Figure 2. Band alignment between the ETL materials and  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 perovskite.
Nanomaterials 13 01537 g002
Figure 3. Impact of the ETL material on (a) the current density–voltage characteristics and (b) the PCE of the PSC.
Figure 3. Impact of the ETL material on (a) the current density–voltage characteristics and (b) the PCE of the PSC.
Nanomaterials 13 01537 g003
Figure 4. Quantum efficiency for the simulated devices with different ETL materials.
Figure 4. Quantum efficiency for the simulated devices with different ETL materials.
Nanomaterials 13 01537 g004
Figure 5. Variations of (a P C E , (b V o c , (c J s c , and (d F F as a function of the ETL thickness for the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Figure 5. Variations of (a P C E , (b V o c , (c J s c , and (d F F as a function of the ETL thickness for the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Nanomaterials 13 01537 g005
Figure 6. Effect of ETL thickness on the quantum efficiency of the PSC with two different ETL materials: (a Z n O and (b)   Z n 0.7 M g 0.3 O .
Figure 6. Effect of ETL thickness on the quantum efficiency of the PSC with two different ETL materials: (a Z n O and (b)   Z n 0.7 M g 0.3 O .
Nanomaterials 13 01537 g006
Figure 7. Effect of the doping concentration of the ETL  Z n O on (a) the current density–voltage characteristics of the PSC and (b) the PCE, and the effect of the doping concentration of ETLZn0.7Mg0.3O on (c) the current density–voltage characteristics of the PSC and (d) the PCE.
Figure 7. Effect of the doping concentration of the ETL  Z n O on (a) the current density–voltage characteristics of the PSC and (b) the PCE, and the effect of the doping concentration of ETLZn0.7Mg0.3O on (c) the current density–voltage characteristics of the PSC and (d) the PCE.
Nanomaterials 13 01537 g007
Figure 8. Impact of the absorber thickness on (a P C E , (b V o c , (c J s c , and (d F F for PSCs with the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Figure 8. Impact of the absorber thickness on (a P C E , (b V o c , (c J s c , and (d F F for PSCs with the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Nanomaterials 13 01537 g008
Figure 9. Impact of the absorber thickness on the quantum efficiency of PSCs with (a Z n O ETL and (b Z n 0.7 M g 0.3 O ETL.
Figure 9. Impact of the absorber thickness on the quantum efficiency of PSCs with (a Z n O ETL and (b Z n 0.7 M g 0.3 O ETL.
Nanomaterials 13 01537 g009
Figure 10. Effect of the absorber defect density on the current density–voltage characteristics and the PCE of the PSC with  Z n O ETL (a,b) and  Z n 0.7 M g 0.3 O ETL (c,d), respectively.
Figure 10. Effect of the absorber defect density on the current density–voltage characteristics and the PCE of the PSC with  Z n O ETL (a,b) and  Z n 0.7 M g 0.3 O ETL (c,d), respectively.
Nanomaterials 13 01537 g010
Figure 11. Effect of the absorber layer defect density on the recombination rate with the depth from the surface of the PSC with (a Z n O ETL and ( b Z n 0.7 M g 0.3 O ETL.
Figure 11. Effect of the absorber layer defect density on the recombination rate with the depth from the surface of the PSC with (a Z n O ETL and ( b Z n 0.7 M g 0.3 O ETL.
Nanomaterials 13 01537 g011
Figure 12. Impact of the work function of the anode on the (a P C E , (b V o c , (c J s c , and (d F F for PSCs with the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Figure 12. Impact of the work function of the anode on the (a P C E , (b V o c , (c J s c , and (d F F for PSCs with the two ETL materials  Z n O and  Z n 0.7 M g 0.3 O .
Nanomaterials 13 01537 g012
Table 1. Electrical and optical properties employed in the simulation of  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based PSC.
Table 1. Electrical and optical properties employed in the simulation of  F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 -based PSC.
Parameters   C 60 (ETL) [53,54,55,56]   F A 0.75 M A 0.25 S n 0.95 G e 0.05 I 3 (Aborsober) [38,39,57] P E D O T : P S S (HTL)
[53,54,55,56]
Thickness ( µ m )   0.05   0.4   0.05
Bandgap   E g (eV)   1.7   1.4   1.8
Electron Affinity  χ (eV)   4.50   3.670   3.40
Dielectric permittivity   18   8.2   18
CB effective density
of states  ( c m 3 )
  2.2 × 10 18   2.2 × 10 18   2.2 × 10 18
VB effective density
of states  ( c m 3 )
  1.8 × 10 19   1.8 × 10 19   1.8 × 10 19
Electron mobility
( c m 2 / V . s )
  8 × 10 2   2   4.5 × 10 2
Hole mobility
( c m 2 / V . s )
  8 × 10 2   2   4.5 × 10 2
Donor Concentration
N D   ( c m 3 )
  1 × 10 17   1 × 10 13   1 × 10 7
Acceptor concentration  N A   ( c m 3 )   0   0   1 × 10 18
Table 2. Electrical and optical properties of different ETL materials.
Table 2. Electrical and optical properties of different ETL materials.
Parameters P C B M
[53,54,55,56]
T i O 2
[42,58]
Z n O
[59,60]
Z n 0.7 M g 0.3 O
[56,61,62]
Thickness ( µ m )0.050.050.050.05
Bandgap  E g (eV)2   3.26 3.3   4
Electron Affinity  χ (eV)4   4   3.7   3.65
Dielectric permittivity3.9   32   9   8
CB effective density
of states  ( c m 3 )
  2.5 × 10 21   2.2 × 10 18   2.2 × 10 18   2.2 × 10 18
VB effective density
of states  ( c m 3 )
  2.5 × 10 21   1.8 × 10 19   1.8 × 10 19   1.8 × 10 19
Electron mobility
( c m 2 / V . s )
  2.5 × 10 2   20   100   100
Hole mobility
( c m 2 / V . s )
  2.5 × 10 2   10   25   25
Donor Concentration
N D   ( c m 3 )
  1 × 10 17   1 × 10 17   1 × 10 17   1 × 10 17
Acceptor concentration  N A   ( c m 3 ) 0   0   0   0
Table 3. Defect density values inside the layers and at the interface of the cell.
Table 3. Defect density values inside the layers and at the interface of the cell.
ParametersETLHTLPerovskiteHTL/PerovskitePerovskite/ETL
Defect TypeNeutralNeutralNeutralNeutralNeutral
Capture cross-section
for electrons  σ n   ( c m 2 )
1 × 10 15 1 × 10 15 1 × 10 15 1 × 10 18 1 × 10 15
Capture cross-section
for hole  σ p   ( c m 2 )
1 × 10 15 1 × 10 15 1 × 10 15 1 × 10 16 1 × 10 15
Energetic distributionSingleSingleGaussianSingleSingle
Energy level with respect to  E v (above  E v ) (eV)0.60.6500.60.60.6
Characteristic energy (eV)0.10.10.10.10.1
Total density  N t   ( c m 3 )   1 × 10 15   1 × 10 15   1 × 10 16   1 × 10 12   1 × 10 12
Table 4. Conduction band offset for the ETL materials and their photovoltaic properties.
Table 4. Conduction band offset for the ETL materials and their photovoltaic properties.
ETLCBO/eV P C E / % V o c / V J s c / m A · c m 3   F F / %
  C 60   0.83   8.47   0.51   21.97   75.43
  P C B M   0.33   11.64   0.60   24.64   78.09
  T i O 2   0.33   14.43   0.65   27.61   80.02
  Z n O   0.02   20.88   0.95   27.75   78.61
  Z n 0.7 M g 0.3 O   0.02   21.55   0.94   27.86   82.33
Table 5. Conduction band offset for the ETL materials and their photovoltaic properties.
Table 5. Conduction band offset for the ETL materials and their photovoltaic properties.
MetalAlAgCrAuNiPdPt
work function
M / e V
  4.125   4.26   4.4   5.1   5.15   5.3   5.15
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

Sabbah, H.; Baki, Z.A. Device Simulation of Highly Stable and 29% Efficient FA0.75MA0.25Sn0.95Ge0.05I3-Based Perovskite Solar Cell. Nanomaterials 2023, 13, 1537. https://doi.org/10.3390/nano13091537

AMA Style

Sabbah H, Baki ZA. Device Simulation of Highly Stable and 29% Efficient FA0.75MA0.25Sn0.95Ge0.05I3-Based Perovskite Solar Cell. Nanomaterials. 2023; 13(9):1537. https://doi.org/10.3390/nano13091537

Chicago/Turabian Style

Sabbah, Hussein, and Zaher Abdel Baki. 2023. "Device Simulation of Highly Stable and 29% Efficient FA0.75MA0.25Sn0.95Ge0.05I3-Based Perovskite Solar Cell" Nanomaterials 13, no. 9: 1537. https://doi.org/10.3390/nano13091537

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop