Next Article in Journal
WAXS and SAXS Investigation of Collagen-Rich Diet Effect on Multiscale Arrangement of Type I Collagen in Tilapia Skin Fed in Aquaponics Plant
Previous Article in Journal
Crystallographic Evidence of η1-Coordination of Bulky Aminopyridine in Halide-Containing Iron (II) Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Silico Investigation of the Impact of Hole-Transport Layers on the Performance of CH3NH3SnI3 Perovskite Photovoltaic Cells

by
Zhansaya Omarova
1,
Darkhan Yerezhep
1,*,
Abdurakhman Aldiyarov
1 and
Nurlan Tokmoldin
2
1
Faculty of Physics and Technology, Al Farabi Kazakh National University, 71 Al-Farabi Ave., Almaty 050040, Kazakhstan
2
Optoelectronics of Disordered Semiconductors, Institute of Physics and Astronomy, University of Potsdam, Karl-Liebknecht-Straße 24–25, 14476 Potsdam-Golm, Germany
*
Author to whom correspondence should be addressed.
Submission received: 22 April 2022 / Revised: 9 May 2022 / Accepted: 12 May 2022 / Published: 14 May 2022
(This article belongs to the Section Materials for Energy Applications)

Abstract

:
Perovskite solar cells represent one of the recent success stories in photovoltaics. The device efficiency has been steadily increasing over the past years, but further work is needed to enhance the performance, for example, through the reduction of defects to prevent carrier recombination. SCAPS-1D simulations were performed to assess efficiency limits and identify approaches to decrease the impact of defects, through the selection of an optimal hole-transport material and a hole-collecting electrode. Particular attention was given to evaluation of the influence of bulk defects within light-absorbing CH3NH3SnI3 layers. In addition, the study demonstrates the influence of interface defects at the TiO2/CH3NH3SnI3 (IL1) and CH3NH3SnI3/HTL (IL2) interfaces across the similar range of defect densities. Finally, the optimal device architecture TiO2/CH3NH3SnI3/Cu2O is proposed for the given absorber layer using the readily available Cu2O hole-transporting material with PCE = 27.95%, FF = 84.05%, VOC = 1.02 V and JSC = 32.60 mA/cm2, providing optimal performance and enhanced resistance to defects.

1. Introduction

Photovoltaic cells based on crystalline silicon have proven themselves at the industrial scale as a viable alternative energy source due to their high performance, material abundance and proven microelectronic technology [1,2,3]. At the same time, in spite of the excellent power conversion efficiencies [4,5,6], the technology still requires further improvement through reduction in costs, facilitation of the manufacturing steps and minimization of the environmental pollution associated with the production process [7,8,9]. In recent years, perovskite solar cells (PSC) based on organometallic lead halides have emerged as strong competitors to silicon in the photovoltaic market [10,11,12], as well as an efficient technology to complement the silicon photovoltaic devices in tandem architecture [13,14,15]. These materials boast high efficiency (about 25%) at a fraction of the silicon device thickness, as well as ease of fabrication, making them highly promising for future photovoltaic applications [16,17,18,19,20,21,22]. One of the main downsides of these devices is that the best efficiencies have so far been demonstrated by organometallic lead halides which present a serious danger to the environment due to the toxicity of lead [23,24,25]. Thus, replacing elemental lead with environmentally friendly alternatives in the perovskite lattice is a pressing problem in photovoltaics [26,27].
A significant amount of research is being dedicated to developing lead-free PSC, with the tin halide perovskite being one of the most promising alternatives [28,29,30,31]. Tin is widely distributed in nature and has similar electronic properties to lead, since it is a member of the same group in the periodic table [32,33]. In addition, perovskites based on tin halide have excellent light absorbing properties and high carrier mobilities [34,35,36]. Additionally, tin-based perovskites provide a high theoretical PCE due to a smaller band gap than the equivalent lead-based perovskites [37].
A PSC in most cases consists of an absorbing layer and two transport layers—an electron transport layer (ETL) and a hole transport layer (HTL), which perform the function of collecting charge carriers [38]. The device structure is completed with a transparent conducting electrode on top of a glass substrate, on the one side, and a metal electrode (such as Au, Ag, Al, etc.), on the other. To be efficient, the transport layers must possess certain properties: above all, high transparency and a high selective charge transport. As such, determination of the optimal properties of an ETL and an HTL plays an important role in maximizing the PSC performance and maintaining the device stability [39].
So far, a variety of transport layers have been tested, ranging from metal-oxides and fullerenes to organic materials and self-assembled monolayers [19,40,41,42,43]. For example, TiO2 has been a popular choice for an ETL in the n-i-p device configuration, due to its chemical stability, good electron transport, efficient hole-blocking at the interface and environmental friendliness [44,45,46]. On the other side, a significant effort has also been placed on the development and optimization of HTLs. Efficient HTL materials should possess the following properties [47,48,49,50]:
  • high carrier mobility to increase the fill factor (FF);
  • a wide optical band gap and high transparency to minimize optical losses;
  • high resistance to water, light and heat;
  • low cost of materials and production;
  • environmental friendliness.
The hole transport layers must be chosen thoughtfully to prevent charge recombination at the interfacial layer boundaries and ensure high device performance. At the same time, it is very important that their fabrication is compatible with the low-cost deposition, solution-processability and flexibility, similar to the other layers within the perovskite cell structure. We thus selected three HTL layers, which are commonly used in perovskite-based solar cells: Spiro-OMeTAD [51,52,53,54], PEDOT:PSS [55,56,57] and Cu2O [58,59,60]. It must be noted that whilst the former two materials can be deposited in a straightforward fashion from, respectively, organic or aqueous solvents, the current approach to solution deposition of Cu2O is more complicated, initially involving the deposition of CuI following its further chemical conversion to Cu2O [61]. We nevertheless consider this metal-oxide HTL in our study due to its high performance and versatility.
Further improvement in the device power conversion efficiency (PCE) requires a detailed understanding of the PSC working mechanism, based not only on experimental research, which may end up being expensive and time-consuming, but also on device simulation [62,63,64,65,66,67,68,69,70]. A limited number of studies focusing on the simulation of lead-free PSCs, as well as on the optimization of their charge-transport layers, has been conducted so far [71,72,73,74,75]. In this article, a cell architecture has been selected using CH3NH3SnI3 as an absorbing layer with three different HTLs, namely 2,2′,7,7′-Tetrakis-(N,N-di-p-methoxyphenylamine) 9,9′-spirobifluorene (Spiro-OMETAD), poly(3,4-ethylenedioxythiophene)-poly(styrene sulfonate) (PEDOT:PSS) and cuprous oxide (Cu2O). Moreover, we consider the applicability of various metal back contacts, such as Al, Ti, Cr, Ag, Ni, Cu, C, Au and Pt. In the presented numerical study, various PSC configurations are defined and evaluated using the SCAPS 1D simulator.

2. Method

The crystal structure of the tin-based perovskite CH3NH3SnI3 employed in this study as the absorber layer is shown in Figure 1. The simulated photovoltaic cell configurations are as follows (Figure 2):
Structure 1. TiO2/CH3NH3SnI3/Spiro-OMeTAD;
Structure 2. TiO2/CH3NH3SnI3/PEDOT:PSS;
Structure 3. TiO2/CH3NH3SnI3/Cu2O.
For all simulations of the photovoltaic cell under illumination the standard photovoltaic radiation spectrum AM 1.5G was used (1000 W/m2, T = 300 K). The CH3NH3SnI3 layer thickness remained fixed and equal to 500 nm.
Numerical simulation has proved to be an important tool for understanding the physical properties and design of a variety of solar cells based on crystalline, polycrystalline and amorphous materials [76,77,78]. Among numerical analysis instruments, SCAPS-1D (ELIS, University of Ghent, Belgium) has recently become popular, as it has proven its effectiveness in simulating a variety of research systems [25,42,62,63,64,65,76,79]. The SCAPS-1D software uses a combination of mathematical equations, including the Poisson equation, continuity equations, total charge transfer equations, whose detailed description can be found elsewhere [80,81].
The parameters of the simulated photovoltaic cell are presented in Table 1 and are based on the published literature [53,82,83,84,85].

3. Results

3.1. Influence of Bulk Defect Density on JSC, VOC, FF, PCE in the Absorber Layer

We start with the study of sensitivity of the performance of PSCs having different HTLs towards the presence of bulk defects within the perovskite layer. In the manufacture of perovskite films, various defects may arise, which have a detrimental impact upon the device performance [86]. Depending on their location, defects are characterized as being deep or shallow. Shallow defects do not usually have a significant impact, whereas deep defects have a more detrimental influence on the device performance, as they are located close to the center of the band gap and trap both types of carriers (holes and electrons), thereby allowing sufficient time for recombination [87,88]. Thus, approximately 1010–1013 cm−3 can be identified as shallow defects, and deep defects starting from the defect density of 1014 cm−3 and up to about 1016 cm−3 [89,90]. Deep defects are located at the center of the band gap, i.e., at approximately 0.65 eV to 0.76 eV above the edge of the valence band, depending on the material [91]. To study the impact of bulk defects on the perovskite efficiency, the energy level position was chosen to be located at 0.6 eV above the valence band with the concentration range between 1010 cm−3 and 1017 cm−3.
Bulk defects are introduced to explain the influence of internal defects on the properties of semiconductors, including fixing the Fermi energy on the surface of semiconductor materials, stabilization of the Fermi energy, and formation of Schottky barriers [92]. Figure 3 shows the impact of the bulk defect density on the parameters of the CH3NH3SnI3 based solar cells with different HTLs in the defect concentration range from 1010 cm−3 to 1017 cm−3. As seen in Figure 3a–c, the device parameters (JSC, VOC and FF) remain unchanged at defect concentrations below 1013 cm−3, followed by a sharp decrease at higher defect concentrations. The increase to 1017 cm−3 results in a rather similar reduction in VOC from 1.0 V to 0.7 V (Figure 3b) and JSC from 32 mA/cm2 to below 25 mA/cm2 for all of the three device structures (Figure 3a). The largest difference is observed for FF, with Structure 1 showing the lower initial value of 86%, compared to 88% for the other structures, and a stronger decrease to 37%, compared to 46% for Structure 3. As a result, the increase in bulk defect concentration from 1010 cm−3 to 1017 cm−3 leads to a reduction in PCE from around 29% to 6%, 7% and 8%, respectively, for Structures 1, 2 and 3 (Figure 3d). The bulk defects act as recombination centers for charge carriers, which is responsible for the loss in the open-circuit voltage. At the same time, the respective charge trapping on these centers may also lead to a reduction in effective carrier mobility, causing a drop ©n JSC and FF.

3.2. Influence of the Density on JSC, VOC, FF, PCE of Interfacial Defects

Interfacial recombination is known to be one of the major factors affecting the performance of PSCs [93,94,95]. To this end, we performed a study on the influence of the density of interfacial defects at both interfaces of the perovskite layer, TiO2/CH3NH3SnI3 (IL1) and CH3NH3SnI3/HTL (IL2), on the efficiency of devices with different HTLs (Figure 4). The total density (Ni) of interfacial defects varied in the range from 1010 cm2 to 1017 cm2. Within this range, VOC drops insignificantly for all the device structures (Figure 4b). Surprisingly, the device performance remains unchanged in a wide range of interface defect concentrations up to 1015 cm2. Further increase to 1017 cm2 results in a sharp decrease in the performance for all the device structures. It must be noted that the FF of Structure 3 remains relatively stable for the studied range of interface defect concentrations. For all the structures, there are no significant changes in FF up to Ni of 1016 cm-2; above this density there is a slight decrease in FF from 81% to 78%, from 83% to 82% and from 84% to 83%, for Structures 1, 2 and 3, respectively (see Figure 4c).
Next, we simulated the impact of Ni in the range from 1010 to 1017 cm2 at the ETL/perovskite and perovskite/HTL interfaces. Nt was chosen to be 1014 cm−3 throughout the entire numerical experiment. Figure 5b,d shows a decrease in VOC and PCE with increasing Ni at the ETL/perovskite interface for the devices with different HTLs. When the density of the defect states at the interface reaches 1017 cm2, we observe a faster decrease in the efficiency of the Spiro-OMeTAD based PSC (from 27% to 19%). For the PEDOT:PSS and Cu2O layers, the decrease is less significant (from 28% to 22% and 28% to 24%, respectively). JSC and FF, however, do not show a noticeable decrease for the selected range of defect concentrations (Figure 5a,c). It should also be noted that in the Ni range from 1011 cm2 to 1016 cm2, a rise in FF is observed for all the HTLs. For example, for the Cu2O and PEDOT:PSS layers, the highest FF value of 86% is observed for Ni in the range of 1014–1015 cm−2. It may be concluded that the allowable concentration of interfacial defects Ni for the TiO2/CH3NH3SnI3 (IL1) interface is 1014 cm2, since the efficiency of the photovoltaic cell deteriorates greatly beyond this level. At the defect density of 1014 cm−2, the optimal Cu2O HTL demonstrates the following photovoltaic characteristics: JSC = 32.5 mA/cm2, VOC = 0.8 V, FF = 83%, PCE = 24%.
Separately, we studied the influence of the density of defect states at the HTL/perovskite interface (Figure 6). Surprisingly, the influence of the density of interfacial defects at the IL2 interface is more pronounced than at the IL1 interface. Upon a change in the HTL, the change in VOC with an increase in the density of defects has a similar character, as in the case of the IL1 interface (Figure 6b). When Ni reaches 1017 cm−2, VOC drops from 1.0 V to 0.7 V for the Spiro-OMeTAD layer, from 1.0 V to 0.8 V for the PEDOT:PSS layer, and from 1.0 V up to 0.9 V for the Cu2O layer. The density of the defect states at the HTL/perovskite interface had no significant impact on JSC up to 1015 cm−2 (Figure 6a). Above this concentration, JSC decreases sharply for all the studied device structures. Figure 6c shows the dependence of FF with increasing Ni. Similar to Figure 5c, an increase in Ni to 1015 cm−2 entails an increase in the FF value. For the Spiro-OMeTAD HTL the peak FF value is 84% at Ni of 1013 cm−2, for PEDOT:PSS—86% at Ni of 1014 cm−2, and for Cu2O—86% at a Ni of 1015 cm−2. PCE decreases with increasing Ni for all the structures with different HTLs, as depicted in Figure 6d. Thus, for the Spiro-OMeTAD layer, PCE decreases from 27% to 18%, for PEDOT:PSS—from 28% to 21%, and for Cu2O—from 28% to 23% with an increase in the density of defects from 1010 cm−2 to 1017 cm−2.
The allowable limit of Ni achieved prior to a rapid performance decrease for Structure 1 is 1013 cm−2, for Structure 2—1014 cm−2 and for Structure 3—1015 cm−2. The difference in the limits of resistance towards defects, depending on the HTL, is associated with various degrees of recombination of charge carriers at the interface. A high limit of resistance towards defects at the TiO2/CH3NH3SnI3 interface indicates good matching of the conduction band levels for the adjacent materials. It is thus concluded that an increase in the density of bulk defects within the active perovskite layer (CH3NH3SnI3) affects the device performance more strongly than the increase in the number of interfacial defects, regardless of the choice of an HTL material. These results provide a quantitative understanding of the threshold defect density values for different perovskite cell structures. To increase the overall cell efficiency, the recombination loss at the interfaces must be reduced [96].

3.3. Influence of the Metal Contact on the Device Performance

Selection of an optimal top metal contact is an important route towards increasing the economic feasibility of a PSC. The main parameter to be varied in this respect is the metal contact work function, which characterizes the amount of energy required to extract an electron from the surface of a metal contact [97]. In this work, various metals with work functions ranging from 4.28 eV (Al) to 5.65 eV (Pt) have been studied (Table 2) [68,97,98,99]. The work function value is directly related to the height of the energy barrier at the metal/charge-transport layer interface affecting the ohmic nature of the contact and the resulting solar cell efficiency (Figure 7).
Based on the simulation results, the highest performance with PCE of ~28% is demonstrated by metals with work functions ranging from 5.00 eV to 5.65 eV. For example, using a conventional Au contact (5.10 eV), the following efficiencies are achieved for various HTLs: 26.9% (Spiro-OMeTAD), 27.8% (PEDOT:PSS) and 27.9% (Cu2O). Figure 8 shows the J-V characteristic curves for the devices featuring different HTLs. The Spiro-OMeTAD layer shows the least efficient performance, suggesting that its replacement with PEDOT:PSS or Cu2O is justified.

3.4. Influence of Temperature on the Device Performance

One of the main hindrances in the widespread commercialization of perovskite solar cells is their long-term stability. Most perovskites do not undergo any phase transitions, which may affect their performance, over a wide temperature range. This suggests that the main cause of thermal degradation in PSCs is not a phase transition, but the decomposition of the perovskite material. The perovskite components are known to be connected by relatively weak ionic bonds, hydrogen bonds and van der Waals forces [100]. Weak interconnections, under the influence of the atmosphere, heat and light radiation, inevitably lead to the material destruction [101]. Herein, we focus on the study of the device thermal stability in the absence of chemical degradation in the temperature range of 290–400 K. Figure 9 shows the impact of temperature on the performance of a PSC for various HTLs. A gradual decrease in all parameters with increasing temperature can be observed for all three device structures. On average, the temperature increase from 290 K to 400 K results in a perovskite PCE decrease of 5% in the absolute value. Cu2O shows the highest performance in the whole range, demonstrating its high promise as an HTL. It can also be seen that when comparing organic HTLs, PEDOT:PSS shows higher efficiency than Spiro-OMeTAD at low temperatures (290–360 K). The general pattern of decrease in PCE with increasing temperature is consistent with earlier work [68,83,102]. Although the impact of defects is not considered directly in this study, we assume that the relatively weak perovskite crystal lattice may provoke the formation of defects; therefore, ionic defects in the lattice itself can be activated under the thermal impact. Accordingly, the accumulation of ionic defects may cause degradation of the crystalline structure of the perovskite film and the transport layers, which can significantly affect the solar cell stability [100].

3.5. Comparison of J-V Characteristics for Different HTLs

To determine the optimal HTL among those studied, the J-V curves were simulated for various combinations of the TiO2/CH3NH3SnI3/HTL/Au structure (Figure 10).
The Cu2O based PSC shows the most optimal J-V characteristic curve. This is indicated by the lower series resistance and higher shunt resistance demonstrated by Structure 3. As a result, Cu2O is shown to be a promising inorganic HTL for PSCs, whose relatively low cost and stability may also signal its high promise and economic feasibility in other applications.

3.6. Influence of the Thickness of the Light-Absorbing Layer on the Device Performance

Finally, an attempt was made to evaluate the impact of the absorbing layer thickness on the PSC performance. The perovskite thickness was varied in the range of 300–1300 nm. Figure 11 shows the PSC J-V characteristics for various thicknesses of the CH3NH3SnI3 absorbing layer.
The simulation results are summarized in Figure 12, with the blue line indicating photovoltaic characteristics in the presence of interfacial defects (1010 cm−2), and the red line—in their absence. According to Figure 12, interfacial defects do not result in any significant deterioration, but rather a small improvement. Figure 12a suggests that, as the thickness is increased in the range of 300–700 nm, a noticeable growth in JSC is observed from 28.97 mA/cm2 to 34.01 mA/cm2 due to a higher optical absorption; upon the addition of interfacial defects this value rises by a further 0.10 mA/cm2. Increasing the thickness from 700 nm to 1300 nm leads to a further small rise in JSC by 1.14 mA/cm2.
As shown in Figure 12b, VOC decreases slightly as the thickness of the absorber layer increases throughout the studied range. It should be noted that the decrease in VOC is directly related to the increase in JSC, which directly affects carrier recombination. Figure 12c shows the variation in FF as a function of the absorber layer thickness. The monotonous decrease in FF for an increasingly thick absorber layer is explained by an increase in series resistance. Figure 12d shows the resulting PCE performance, affected by an increase in JSC and reductions in VOC and FF. In the range of 300–700 nm PCE shows a steady increase by more than 10%. However, at higher thicknesses up to 1300 nm, a moderate loss of PCE of about 2% is observed. Thus, 700 nm indicates the optimal device thickness under given conditions.

4. Conclusions

The simulations focus on the influence of bulk defects within the absorbing layer, interfacial defects, temperature and thickness of the absorbing layer on the device performance. These factors were studied for three different HTLs, including Spiro-OMeTAD, PEDOT:PSS and Cu2O, in order to determine the optimal transport layer, least sensitive towards performance deterioration. It has been found that the PSC works best at the room temperature of 300 K, in the absence of ionic contribution. For bulk defect densities larger than 1015 cm−3, a decrease in the performance was observed for all the device structures. The tolerance limit for the density of interfacial defects at the TiO2/CH3NH3SnI3 (IL1) and CH3NH3SnI3/HTL (IL2) interfaces was 1014 cm−3 and 1010 cm−3. Among the considered structures, the TiO2/CH3NH3SnI3/Cu2O structure shows the best defect resistance and the highest performance. The most suitable metallic contact was determined to be Pt, which provides high efficiency of 26.96%, 27.80% and 27.95% for structures with HTLs of Spiro-OMeTAD, PEDOT:PSS and Cu2O, respectively. As a result of the study, the chosen optimal device structure was identified: TiO2/CH3NH3SnI3/Cu2O with JSC = 32.60 mA/cm2, VOC = 1.02 V, FF = 84.05% and PCE = 27.95%, which provide the high PCE and higher resistance towards both bulk and interfacial defects. This efficiency value is below the Shockley–Quisser limit of around 33% for the bandgap of the given perovskite absorber. It is, however, significantly higher than the current record experimental value of 14.6% for a tin perovskite solar cell [103]. Whilst the Shockley–Quisser limit assumes a number of simplifications, such as the absence of optical losses, non-radiative recombination and perfect carrier mobility, these factors can, to some extent, be accounted for in in-silico investigations. At the same time, the latter produce an “optimistic” performance limit for the given material properties with further improvements required regarding a realistic description of bulk, interface and surface recombination, the role of defects on carrier mobility and lifetime, as well as the impact of mobile ions on the device performance and stability. We therefore hope that the results of this work will help to reveal further approaches to achieve higher efficiency in tin-based perovskite solar cells.

Author Contributions

Conceptualization, Z.O., D.Y., A.A. and N.T.; methodology, Z.O., D.Y. and N.T.; software, Z.O.; formal analysis, Z.O., D.Y. and N.T.; investigation, Z.O., D.Y.; data curation, Z.O., D.Y. and N.T.; writing—original draft preparation, Z.O. and D.Y.; writing—review and editing, Z.O., D.Y., A.A. and N.T.; visualization, D.Y.; supervision, N.T.; project administration, A.A. and N.T.; funding acquisition, A.A. All authors have read and agreed to the published version of the manuscript.

Funding

These studies have been carried out with the financial support of the Ministry of Education and Science of the Republic of Kazakhstan under grant AP08855738.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available upon a reasonable request to the corresponding author.

Acknowledgments

These studies have been carried out with the financial support of the Ministry of Education and Science of the Republic of Kazakhstan under grant AP08855738. The authors acknowledge the provision of SCAPS-1D software by Marc Burgelman. The authors acknowledge the graphic drawings to doctoral student Al Farabi Kazakh National University Golikov Oleg.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Green, M.A.; Hishikawa, Y.; Dunlop, E.D.; Levi, D.H.; Hohl-Ebinger, J.; Ho-Baillie, A.W.Y. Solar cell efficiency tables (version 52). Prog. Photovolt. Res. Appl. 2018, 26, 427–436. [Google Scholar] [CrossRef]
  2. Yoshikawa, K.; Kawasaki, H.; Yoshida, W.; Irie, T.; Konishi, K.; Nakano, K.; Uto, T.; Adachi, D.; Kanematsu, M.; Uzu, H.; et al. Silicon heterojunction solar cell with interdigitated back contacts for a photoconversion efficiency over 26%. Nat. Energy 2017, 2, 17032. [Google Scholar] [CrossRef]
  3. Elsmani, M.I.; Fatima, N.; Jallorina, M.P.A.; Sepeai, S.; Su’ait, M.S.; Ahmad Ludin, N.; Mat Teridi, M.A.; Sopian, K.; Ibrahim, M.A. Recent issues and configuration factors in perovskite-silicon tandem solar cells towards large scaling production. Nanomaterials 2021, 11, 3186. [Google Scholar] [CrossRef] [PubMed]
  4. Alvarez, H.S.; Silva, A.R.; Espinola, L.C.J.; Vaz, A.R.; Diniz, J.A. NH 4 OH-B silicon texturing of periodic V-groove channels, upright, and inverted pyramids structures. IEEE J. Photovolt. 2021, 11, 570–574. [Google Scholar] [CrossRef]
  5. Yuan, S.; Cui, Y.; Zhuang, Y.; Chen, P.; Hu, Y.; Yang, B.; Yu, Y.; Ren, Y.; Wang, W.; Chen, W.; et al. Passivated emitter and rear cell silicon solar cells with a front polysilicon passivating contacted selective emitter. Phys. Status Solidi—Rapid Res. Lett. 2021, 15, 2100057. [Google Scholar] [CrossRef]
  6. Richter, A.; Müller, R.; Benick, J.; Feldmann, F.; Steinhauser, B.; Reichel, C.; Fell, A.; Bivour, M.; Hermle, M.; Glunz, S.W. Design rules for high-efficiency both-sides-contacted silicon solar cells with balanced charge carrier transport and recombination losses. Nat. Energy 2021, 6, 429–438. [Google Scholar] [CrossRef]
  7. Owusu, P.A.; Asumadu-Sarkodie, S. A review of renewable energy sources, sustainability issues and climate change mitigation. Cogent Eng. 2016, 3, 1167990. [Google Scholar] [CrossRef]
  8. Ramírez-Márquez, C.; Villicaña-García, E.; Cansino-Loeza, B.; Segovia-Hernández, J.G.; Ponce-Ortega, J.M. Inherent occupational health hazards in the production of solar grade silicon. Process Saf. Environ. Prot. 2020, 142, 285–294. [Google Scholar] [CrossRef]
  9. Ramírez-Márquez, C.; Contreras-Zarazúa, G.; Martín, M.; Segovia-Hernández, J.G. Safety, economic, and environmental optimization applied to three processes for the production of solar-grade silicon. ACS Sustain. Chem. Eng. 2019, 7, 5355–5366. [Google Scholar] [CrossRef]
  10. Momblona, C.; Gil-Escrig, L.; Bandiello, E.; Hutter, E.M.; Sessolo, M.; Lederer, K.; Blochwitz-Nimoth, J.; Bolink, H.J. Efficient vacuum deposited p-i-n and n-i-p perovskite solar cells employing doped charge transport layers. Energy Environ. Sci. 2016, 9, 3456–3463. [Google Scholar] [CrossRef]
  11. Zhao, X.; Sun, Y.; Liu, S.; Chen, G.; Chen, P.; Wang, J.; Cao, W.; Wang, C. Humidity sensitivity behavior of CH3NH3PbI3 perovskite. Nanomaterials 2022, 12, 523. [Google Scholar] [CrossRef] [PubMed]
  12. Triolo, C.; De Giorgi, M.L.; Lorusso, A.; Cretì, A.; Santangelo, S.; Lomascolo, M.; Anni, M.; Mazzeo, M.; Patané, S. Light emission properties of thermally evaporated CH3NH3PbBr3 perovskite from nano- to macro-scale: Role of free and localized excitons. Nanomaterials 2022, 12, 211. [Google Scholar] [CrossRef] [PubMed]
  13. Nayak, P.K.; Mahesh, S.; Snaith, H.J.; Cahen, D. Photovoltaic solar cell technologies: Analysing the state of the art. Nat. Rev. Mater. 2019, 4, 269–285. [Google Scholar] [CrossRef]
  14. Zhou, D.; Zhou, T.; Tian, Y.; Zhu, X.; Tu, Y. Perovskite-based solar cells: Materials, methods, and future perspectives. J. Nanomater. 2018, 2018, 8148072. [Google Scholar] [CrossRef]
  15. Hwang, T.; Lee, B.; Kim, J.; Lee, S.; Gil, B.; Yun, A.J.; Park, B. From nanostructural evolution to dynamic interplay of constituents: Perspectives for perovskite solar cells. Adv. Mater. 2018, 30, 1704208. [Google Scholar] [CrossRef] [PubMed]
  16. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.-K.; Moon, C.S.; Jeon, N.J.; Correa-Baena, J.-P.; et al. Efficient perovskite solar cells via improved carrier management. Nature 2021, 590, 587–593. [Google Scholar] [CrossRef]
  17. Li, N.; Niu, X.; Li, L.; Wang, H.; Huang, Z.; Zhang, Y.; Chen, Y.; Zhang, X.; Zhu, C.; Zai, H.; et al. Liquid medium annealing for fabricating durable perovskite solar cells with improved reproducibility. Science 2021, 373, 561–567. [Google Scholar] [CrossRef]
  18. Lee, K.-M.; Chan, S.-H.; Chiu, W.-H.; Ahn, S.; Ting, C.-C.; Chang, Y.-H.; Suryanarayanan, V.; Wu, M.-C.; Liu, C.-Y. Reducing defects in organic-lead halide perovskite film by delayed thermal annealing combined with KI/I2 for efficient perovskite solar cells. Nanomaterials 2021, 11, 1607. [Google Scholar] [CrossRef]
  19. Park, H.H. Efficient and stable perovskite solar cells based on inorganic hole transport materials. Nanomaterials 2021, 12, 112. [Google Scholar] [CrossRef]
  20. Wu, J.; Li, Y.; Li, Y.; Xie, W.; Shi, J.; Li, D.; Cheng, S.; Meng, Q. Using hysteresis to predict the charge recombination properties of perovskite solar cells. J. Mater. Chem. A 2021, 9, 6382–6392. [Google Scholar] [CrossRef]
  21. Yang, D.; Zhang, G.; Lai, R.; Cheng, Y.; Lian, Y.; Rao, M.; Huo, D.; Lan, D.; Zhao, B.; Di, D. Germanium-lead perovskite light-emitting diodes. Nat. Commun. 2021, 12, 4295. [Google Scholar] [CrossRef] [PubMed]
  22. Dos Santos Rosa, E.H.; Kowalski, E.L.; Ribeiro Barrozo Toledo, L.F. Simulation of organic solar cells’s power conversion efficiency. Sol. Energy 2021, 221, 483–487. [Google Scholar] [CrossRef]
  23. Wei, H.; Qiu, P.; Li, Y.; He, Y.; Peng, M.; Zheng, X.; Liu, X. Challenges and strategies of all-inorganic lead-free halide perovskite solar cells. Ceram. Int. 2022, 48, 5876–5891. [Google Scholar] [CrossRef]
  24. Li, J.; Cao, H.-L.; Jiao, W.-B.; Wang, Q.; Wei, M.; Cantone, I.; Lü, J.; Abate, A. Biological impact of lead from halide perovskites reveals the risk of introducing a safe threshold. Nat. Commun. 2020, 11, 310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Montoya De Los Santos, I.; Cortina-Marrero, H.J.; Ruíz-Sánchez, M.A.; Hechavarría-Difur, L.; Sánchez-Rodríguez, F.J.; Courel, M.; Hu, H. Optimization of CH3NH3PbI3 perovskite solar cells: A theoretical and experimental study. Sol. Energy 2020, 199, 198–205. [Google Scholar] [CrossRef]
  26. Cao, W.; Hu, Z.; Lin, Z.; Guo, X.; Su, J.; Chang, J.; Hao, Y. Defects and doping engineering towards high performance lead-free or lead-less perovskite solar cells. J. Energy Chem. 2022, 68, 420–438. [Google Scholar] [CrossRef]
  27. Chandra, P.; Mandal, S.K. Morphology controlled (CH3NH3)3Bi2Cl9 thin film for lead free perovskite solar cell. Phys. B Condens. Matter 2022, 625, 413536. [Google Scholar] [CrossRef]
  28. Bai, F.; Hu, Y.; Hu, Y.; Qiu, T.; Miao, X.; Zhang, S. Lead-free, air-stable ultrathin Cs3Bi2I9 perovskite nanosheets for solar cells. Sol. Energy Mater. Sol. Cells 2018, 184, 15–21. [Google Scholar] [CrossRef]
  29. Turkevych, I.; Kazaoui, S.; Ito, E.; Urano, T.; Yamada, K.; Tomiyasu, H.; Yamagishi, H.; Kondo, M.; Aramaki, S. Photovoltaic rudorffites: Lead-free silver bismuth halides alternative to hybrid lead halide perovskites. ChemSusChem 2017, 10, 3754–3759. [Google Scholar] [CrossRef]
  30. 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]
  31. Ünlü, F.; Deo, M.; Mathur, S.; Kirchartz, T.; Kulkarni, A. Bismuth-based halide perovskite and perovskite-inspired light absorbing materials for photovoltaics. J. Phys. D. Appl. Phys. 2022, 55, 113002. [Google Scholar] [CrossRef]
  32. Toshniwal, A.; Kheraj, V. Development of organic-inorganic tin halide perovskites: A review. Sol. Energy 2017, 149, 54–59. [Google Scholar] [CrossRef]
  33. Giustino, F.; Snaith, H.J. Toward lead-free perovskite solar cells. ACS Energy Lett. 2016, 1, 1233–1240. [Google Scholar] [CrossRef] [Green Version]
  34. Abu Baker, A.M.; Boltaev, G.S.; Iqbal, M.; Pylnev, M.; Hamdan, N.M.; Alnaser, A.S. Giant third-order nonlinear response of mixed perovskite nanocrystals. Materials 2022, 15, 389. [Google Scholar] [CrossRef]
  35. Pitaro, M.; Tekelenburg, E.K.; Shao, S.; Loi, M.A. Tin halide perovskites: From fundamental properties to solar cells. Adv. Mater. 2022, 34, 2105844. [Google Scholar] [CrossRef]
  36. Singh, S.; Kabra, D. Comparative study of recombination dynamics in optimized composition of Sn- versus Pb-based perovskite solar cells. ACS Appl. Mater. Interfaces 2021, 13, 42297–42306. [Google Scholar] [CrossRef]
  37. Roknuzzaman, M.; Ostrikov, K.; Wang, H.; Du, A.; Tesfamichael, T. Towards lead-free perovskite photovoltaics and optoelectronics by ab-initio simulations. Sci. Rep. 2017, 7, 14025. [Google Scholar] [CrossRef]
  38. Le Corre, V.M.; Stolterfoht, M.; Perdigón Toro, L.; Feuerstein, M.; Wolff, C.; Gil-Escrig, L.; Bolink, H.J.; Neher, D.; Koster, L.J.A. Charge transport layers limiting the efficiency of perovskite solar cells: How to optimize conductivity, doping, and thickness. ACS Appl. Energy Mater. 2019, 2, 6280–6287. [Google Scholar] [CrossRef] [Green Version]
  39. Braun, S.; Salaneck, W.R.; Fahlman, M. Energy-level alignment at organic/metal and organic/organic interfaces. Adv. Mater. 2009, 21, 1450–1472. [Google Scholar] [CrossRef]
  40. Deepthi Jayan, K. Design and comparative performance analysis of high-efficiency lead-based and lead-free perovskite solar cells. Phys. Status Solidi 2022, 219, 2100606. [Google Scholar] [CrossRef]
  41. Deepthi Jayan, K.; Sebastian, V. Modelling and comparative performance analysis of tin based mixed halide perovskite solar cells with IGZO and CuO as charge transport layers. Int. J. Energy Res. 2021, 45, 16618–16632. [Google Scholar] [CrossRef]
  42. Chen, Q.; Ni, Y.; Dou, X.; Yoshinori, Y. The effect of energy level of transport layer on the performance of ambient air prepared perovskite solar cell: A SCAPS-1D simulation study. Crystals 2022, 12, 68. [Google Scholar] [CrossRef]
  43. Rombach, F.M.; Haque, S.A.; Macdonald, T.J. Lessons learned from spiro-OMeTAD and PTAA in perovskite solar cells. Energy Environ. Sci. 2021, 14, 5161–5190. [Google Scholar] [CrossRef]
  44. Mahapatra, B.; Krishna, R.V.; Laxmi; Patel, P.K. Design and optimization of CuSCN/CH3NH3PbI3/TiO2 perovskite solar cell for efficient performance. Opt. Commun. 2022, 504, 127496. [Google Scholar] [CrossRef]
  45. Farooq, W.; Alshahrani, T.; Kazmi, S.A.A.; Iqbal, J.; Khan, H.A.; Khan, M.; Raja, A.A.; ur Rehman, A. Materials optimization for thin-film copper indium gallium selenide (CIGS) solar cell based on distributed braggs reflector. Optik 2021, 227, 165987. [Google Scholar] [CrossRef]
  46. Yang, H.-Y.; Rho, W.-Y.; Lee, S.; Kim, S.; Hahn, Y.-B. TiO2 nanoparticles/nanotubes for efficient light harvesting in perovskite solar cells. Nanomaterials 2019, 9, 326. [Google Scholar] [CrossRef] [Green Version]
  47. Kung, P.; Li, M.; Lin, P.; Chiang, Y.; Chan, C.; Guo, T.; Chen, P. A review of inorganic hole transport materials for perovskite solar cells. Adv. Mater. Interfaces 2018, 5, 1800882. [Google Scholar] [CrossRef]
  48. Gil, B.; Yun, A.J.; Lee, Y.; Kim, J.; Lee, B.; Park, B. Recent progress in inorganic hole transport materials for efficient and stable perovskite solar cells. Electron. Mater. Lett. 2019, 15, 505–524. [Google Scholar] [CrossRef]
  49. Arumugam, G.M.; Karunakaran, S.K.; Liu, C.; Zhang, C.; Guo, F.; Wu, S.; Mai, Y. Inorganic hole transport layers in inverted perovskite solar cells: A review. Nano Sel. 2021, 2, 1081–1116. [Google Scholar] [CrossRef]
  50. Xu, B.; Zhu, H.; Bao, H.; Cao, X.; Dong, Y.; Zhang, Y.; Yin, G.; Li, X.; Wang, S. Bifunctional spiro-fluorene/heterocycle cored hole-transporting materials: Role of the heteroatom on the photovoltaic performance of perovskite solar cells. Chem. Eng. J. 2022, 431, 133371. [Google Scholar] [CrossRef]
  51. Roy, P.; Raoui, Y.; Khare, A. Design and simulation of efficient tin based perovskite solar cells through optimization of selective layers: Theoretical insights. Opt. Mater. 2022, 125, 112057. [Google Scholar] [CrossRef]
  52. Singh, N.; Agarwal, A.; Agarwal, M. Performance evaluation of lead–free double-perovskite solar cell. Opt. Mater. 2021, 114, 110964. [Google Scholar] [CrossRef]
  53. Kanoun, A.-A.; Kanoun, M.B.; Merad, A.E.; Goumri-Said, S. Toward development of high-performance perovskite solar cells based on CH3NH3GeI3 using computational approach. Sol. Energy 2019, 182, 237–244. [Google Scholar] [CrossRef]
  54. Song, W.; Rakocevic, L.; Thiruvallur Eachambadi, R.; Qiu, W.; Bastos, J.P.; Gehlhaar, R.; Kuang, Y.; Hadipour, A.; Aernouts, T.; Poortmans, J. Improving the morphology stability of spiro-OMeTAD films for enhanced thermal stability of perovskite solar cells. ACS Appl. Mater. Interfaces 2021, 13, 44294–44301. [Google Scholar] [CrossRef] [PubMed]
  55. Ibanez, J.G.; Rincón, M.E.; Gutierrez-Granados, S.; Chahma, M.; Jaramillo-Quintero, O.A.; Frontana-Uribe, B.A. Conducting polymers in the fields of energy, environmental remediation, and chemical–chiral sensors. Chem. Rev. 2018, 118, 4731–4816. [Google Scholar] [CrossRef] [PubMed]
  56. Anrango-Camacho, C.; Pavón-Ipiales, K.; Frontana-Uribe, B.A.; Palma-Cando, A. Recent advances in hole-transporting layers for organic solar cells. Nanomaterials 2022, 12, 443. [Google Scholar] [CrossRef]
  57. Wang, J.; Yu, H.; Hou, C.; Zhang, J. Solution-Processable PEDOT:PSS:α-In 2 Se 3 with enhanced conductivity as a hole transport layer for high-performance polymer solar cells. ACS Appl. Mater. Interfaces 2020, 12, 26543–26554. [Google Scholar] [CrossRef]
  58. Lee, S.W.; Lee, Y.S.; Heo, J.; Siah, S.C.; Chua, D.; Brandt, R.E.; Kim, S.B.; Mailoa, J.P.; Buonassisi, T.; Gordon, R.G. Improved Cu2O-based solar cells using atomic layer deposition to control the Cu oxidation state at the p-n junction. Adv. Energy Mater. 2014, 4, 1301916. [Google Scholar] [CrossRef]
  59. Murali, D.S.; Kumar, S.; Choudhary, R.J.; Wadikar, A.D.; Jain, M.K.; Subrahmanyam, A. Synthesis of Cu2O from CuO thin films: Optical and electrical properties. AIP Adv. 2015, 5, 047143. [Google Scholar] [CrossRef]
  60. Scanlon, D.O.; Morgan, B.J.; Watson, G.W.; Walsh, A. Acceptor levels in p-Type Cu2O: Rationalizing theory and experiment. Phys. Rev. Lett. 2009, 103, 096405. [Google Scholar] [CrossRef]
  61. Zuo, C.; Ding, L. Solution-processed Cu2O and CuO as hole transport materials for efficient perovskite solar cells. Small 2015, 11, 5528–5532. [Google Scholar] [CrossRef] [PubMed]
  62. He, Y.; Xu, L.; Yang, C.; Guo, X.; Li, S. Design and numerical investigation of a lead-free inorganic layered double perovskite Cs4CuSb2Cl12 nanocrystal solar cell by SCAPS-1D. Nanomaterials 2021, 11, 2321. [Google Scholar] [CrossRef] [PubMed]
  63. Samiul Islam, M.; Sobayel, K.; Al-Kahtani, A.; Islam, M.A.; Muhammad, G.; Amin, N.; Shahiduzzaman, M.; Akhtaruzzaman, M. Defect study and modelling of SnX3-based perovskite solar cells with SCAPS-1D. Nanomaterials 2021, 11, 1218. [Google Scholar] [CrossRef] [PubMed]
  64. Yao, H.; Liu, L. Design and optimize the performance of self-powered photodetector based on PbS/TiS3 heterostructure by SCAPS-1D. Nanomaterials 2022, 12, 325. [Google Scholar] [CrossRef]
  65. Moiz, S.A. Optimization of hole and electron transport layer for highly efficient lead-free Cs2TiBr6-based perovskite solar cell. Photonics 2021, 9, 23. [Google Scholar] [CrossRef]
  66. Piñón Reyes, A.C.; Ambrosio Lázaro, R.C.; Monfil Leyva, K.; Luna López, J.A.; Flores Méndez, J.; Heredia Jiménez, A.H.; Muñoz Zurita, A.L.; Severiano Carrillo, F.; Ojeda Durán, E. Study of a lead-free perovskite solar cell using CZTS as HTL to achieve a 20% PCE by SCAPS-1D simulation. Micromachines 2021, 12, 1508. [Google Scholar] [CrossRef]
  67. Zyoud, S.H.; Zyoud, A.H.; Ahmed, N.M.; Prasad, A.R.; Khan, S.N.; Abdelkader, A.F.I.; Shahwan, M. Numerical modeling of high conversion efficiency FTO/ZnO/CdS/CZTS/MO thin film-based solar cells: Using SCAPS-1D software. Crystals 2021, 11, 1468. [Google Scholar] [CrossRef]
  68. Mottakin, M.; Sobayel, K.; Sarkar, D.; Alkhammash, H.; Alharthi, S.; Techato, K.; Shahiduzzaman, M.; Amin, N.; Sopian, K.; Akhtaruzzaman, M. Design and modelling of eco-friendly CH3NH3SnI3-based perovskite solar cells with suitable transport layers. Energies 2021, 14, 7200. [Google Scholar] [CrossRef]
  69. Raoui, Y.; Ez-Zahraouy, H.; Tahiri, N.; El Bounagui, O.; Ahmad, S.; Kazim, S. Performance analysis of MAPbI3 based perovskite solar cells employing diverse charge selective contacts: Simulation study. Sol. Energy 2019, 193, 948–955. [Google Scholar] [CrossRef]
  70. Lin, L.; Jiang, L.; Li, P.; Xiong, H.; Kang, Z.; Fan, B.; Qiu, Y. Simulated development and optimized performance of CsPbI3 based all-inorganic perovskite solar cells. Sol. Energy 2020, 198, 454–460. [Google Scholar] [CrossRef]
  71. Patel, P.K. Device simulation of highly efficient eco-friendly CH3NH3SnI3 perovskite solar cell. Sci. Rep. 2021, 11, 3082. [Google Scholar] [CrossRef] [PubMed]
  72. Rono, N.; Merad, A.E.; Kibet, J.K.; Martincigh, B.S.; Nyamori, V.O. A theoretical investigation of the effect of the hole and electron transport materials on the performance of a lead-free perovskite solar cell based on CH3NH3SnI3. J. Comput. Electron. 2021, 20, 993–1005. [Google Scholar] [CrossRef]
  73. 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] [PubMed] [Green Version]
  74. Abd Mutalib, M.; Ahmad Ludin, N.; Nik Ruzalman, N.A.A.; Barrioz, V.; Sepeai, S.; Mat Teridi, M.A.; Su’ait, M.S.; Ibrahim, M.A.; Sopian, K. Progress towards highly stable and lead-free perovskite solar cells. Mater. Renew. Sustain. Energy 2018, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  75. Tailor, N.K.; Kar, S.; Mishra, P.; These, A.; Kupfer, C.; Hu, H.; Awais, M.; Saidaminov, M.; Dar, M.I.; Brabec, C.; et al. Advances in lead-free perovskite single crystals: Fundamentals and applications. ACS Mater. Lett. 2021, 3, 1025–1080. [Google Scholar] [CrossRef]
  76. 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. Chinese J. Phys. 2022, 76, 94–109. [Google Scholar] [CrossRef]
  77. Al-Hattab, M.; Moudou, L.; Khenfouch, M.; Bajjou, O.; Chrafih, Y.; Rahmani, K. Numerical simulation of a new heterostructure CIGS/GaSe solar cell system using SCAPS-1D software. Sol. Energy 2021, 227, 13–22. [Google Scholar] [CrossRef]
  78. Houimi, A.; Gezgin, S.Y.; Mercimek, B.; Kılıç, H.Ş. Numerical analysis of CZTS/n-Si solar cells using SCAPS-1D. A comparative study between experimental and calculated outputs. Opt. Mater. 2021, 121, 111544. [Google Scholar] [CrossRef]
  79. Moiz, S.A.; Alahmadi, A.N.M. Design of dopant and lead-free novel perovskite solar cell for 16.85% efficiency. Polymers 2021, 13, 2110. [Google Scholar] [CrossRef]
  80. Burgelman, M.; Decock, K.; Khelifi, S.; Abass, A. Advanced electrical simulation of thin film solar cells. Thin Solid Films 2013, 535, 296–301. [Google Scholar] [CrossRef]
  81. Verschraegen, J.; Burgelman, M. Numerical modeling of intra-band tunneling for heterojunction solar cells in scaps. Thin Solid Films 2007, 515, 6276–6279. [Google Scholar] [CrossRef]
  82. Hao, L.; Zhou, M.; Song, Y.; Ma, X.; Wu, J.; Zhu, Q.; Fu, Z.; Liu, Y.; Hou, G.; Li, T. Tin-based perovskite solar cells: Further improve the performance of the electron transport layer-free structure by device simulation. Sol. Energy 2021, 230, 345–354. [Google Scholar] [CrossRef]
  83. Deepthi Jayan, K.; Sebastian, V. Comprehensive device modelling and performance analysis of MASnI3 based perovskite solar cells with diverse ETM, HTM and back metal contacts. Sol. Energy 2021, 217, 40–48. [Google Scholar] [CrossRef]
  84. 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]
  85. 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]
  86. Noel, N.K.; Abate, A.; Stranks, S.D.; Parrott, E.S.; Burlakov, V.M.; Goriely, A.; Snaith, H.J. Enhanced photoluminescence and solar cell performance via lewis base passivation of organic–inorganic lead halide perovskites. ACS Nano 2014, 8, 9815–9821. [Google Scholar] [CrossRef]
  87. Kearney, K.; Seo, G.; Matsushima, T.; Adachi, C.; Ertekin, E.; Rockett, A. Computational analysis of the interplay between deep level traps and perovskite solar cell efficiency. J. Am. Chem. Soc. 2018, 140, 15655–15660. [Google Scholar] [CrossRef]
  88. Joshi, P.H.; Zhang, L.; Hossain, I.M.; Abbas, H.A.; Kottokkaran, R.; Nehra, S.P.; Dhaka, M.; Noack, M.; Dalal, V.L. The physics of photon induced degradation of perovskite solar cells. AIP Adv. 2016, 6, 115114. [Google Scholar] [CrossRef] [Green Version]
  89. Adinolfi, V.; Yuan, M.; Comin, R.; Thibau, E.S.; Shi, D.; Saidaminov, M.I.; Kanjanaboos, P.; Kopilovic, D.; Hoogland, S.; Lu, Z.-H.; et al. The in-gap electronic state spectrum of methylammonium lead iodide single-crystal perovskites. Adv. Mater. 2016, 28, 3406–3410. [Google Scholar] [CrossRef]
  90. De Quilettes, D.W.; Vorpahl, S.M.; Stranks, S.D.; Nagaoka, H.; Eperon, G.E.; Ziffer, M.E.; Snaith, H.J.; Ginger, D.S. Impact of microstructure on local carrier lifetime in perovskite solar cells. Science 2015, 348, 683–686. [Google Scholar] [CrossRef] [Green Version]
  91. Chowdhury, M.S.; Shahahmadi, S.A.; Chelvanathan, P.; Tiong, S.K.; Amin, N.; Techato, K.; Nuthammachot, N.; Chowdhury, T.; Suklueng, M. Effect of deep-level defect density of the absorber layer and n/i interface in perovskite solar cells by SCAPS-1D. Results Phys. 2020, 16, 102839. [Google Scholar] [CrossRef]
  92. Walukiewicz, W.; Rey-Stolle, I.; Han, G.; Jaquez, M.; Broberg, D.; Xie, W.; Sherburne, M.; Mathews, N.; Asta, M. Bistable amphoteric native defect model of perovskite photovoltaics. J. Phys. Chem. Lett. 2018, 9, 3878–3885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Jiang, Q.; Zhao, Y.; Zhang, X.; Yang, X.; Chen, Y.; Chu, Z.; Ye, Q.; Li, X.; Yin, Z.; You, J. Surface passivation of perovskite film for efficient solar cells. Nat. Photonics 2019, 13, 460–466. [Google Scholar] [CrossRef]
  94. Yoo, J.J.; Wieghold, S.; Sponseller, M.C.; Chua, M.R.; Bertram, S.N.; Hartono, N.T.P.; Tresback, J.S.; Hansen, E.C.; Correa-Baena, J.-P.; Bulović, V.; et al. An interface stabilized perovskite solar cell with high stabilized efficiency and low voltage loss. Energy Environ. Sci. 2019, 12, 2192–2199. [Google Scholar] [CrossRef] [Green Version]
  95. 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]
  96. Wolff, C.M.; Caprioglio, P.; Stolterfoht, M.; Neher, D. Nonradiative recombination in perovskite solar cells: The role of interfaces. Adv. Mater. 2019, 31, 1902762. [Google Scholar] [CrossRef] [Green Version]
  97. Samantaray, M.R.; Rana, N.K.; Kumar, A.; Ghosh, D.S.; Chander, N. Stability study of large-area perovskite solar cells fabricated with copper as low-cost metal contact. Int. J. Energy Res. 2022, 46, 1250–1262. [Google Scholar] [CrossRef]
  98. Çaldıran, Z. Modification of Schottky barrier height using an inorganic compound interface layer for various contact metals in the metal/p-Si device structure. J. Alloys Compd. 2021, 865, 158856. [Google Scholar] [CrossRef]
  99. Lin, C.-H.; Li, T.-Y.; Cheng, B.; Liu, C.; Yang, C.-W.; Ke, J.-J.; Wei, T.-C.; Li, L.-J.; Fratalocchi, A.; He, J.-H. Metal contact and carrier transport in single crystalline CH3NH3PbBr3 perovskite. Nano Energy 2018, 53, 817–827. [Google Scholar] [CrossRef] [Green Version]
  100. Zai, H.; Ma, Y.; Chen, Q.; Zhou, H. Ion migration in halide perovskite solar cells: Mechanism, characterization, impact and suppression. J. Energy Chem. 2021, 63, 528–549. [Google Scholar] [CrossRef]
  101. Cheng, Y.; Liu, X.; Guan, Z.; Li, M.; Zeng, Z.; Li, H.; Tsang, S.; Aberle, A.G.; Lin, F. Revealing the degradation and self-healing mechanisms in perovskite solar cells by sub-bandgap external quantum efficiency spectroscopy. Adv. Mater. 2021, 33, 2006170. [Google Scholar] [CrossRef] [PubMed]
  102. Roy, P.; Tiwari, S.; Khare, A. An investigation on the influence of temperature variation on the performance of tin (Sn) based perovskite solar cells using various transport layers and absorber layers. Results Opt. 2021, 4, 100083. [Google Scholar] [CrossRef]
  103. Jiang, X.; Li, H.; Zhou, Q.; Wei, Q.; Wei, M.; Jiang, L.; Wang, Z.; Peng, Z.; Wang, F.; Zang, Z.; et al. One-step synthesis of SnI 2·(DMSO) x adducts for high-performance tin perovskite solar cells. J. Am. Chem. Soc. 2021, 143, 10970–10976. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Unit cell of tin-based perovskite.
Figure 1. Unit cell of tin-based perovskite.
Crystals 12 00699 g001
Figure 2. Schematic diagram of the studied perovskite solar cell (a) and energy band alignment (b).
Figure 2. Schematic diagram of the studied perovskite solar cell (a) and energy band alignment (b).
Crystals 12 00699 g002
Figure 3. Influence of the bulk defect concentration (Nt) on the performance parameters of the CH3NH3SnI3-based solar cell with various hole-transporting layers: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Figure 3. Influence of the bulk defect concentration (Nt) on the performance parameters of the CH3NH3SnI3-based solar cell with various hole-transporting layers: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Crystals 12 00699 g003
Figure 4. Influence of the concentration of interfacial defects (Ni) at the TiO2/CH3NH3SnI3 (IL1) and CH3NH3SnI3/HTL (IL2) interfaces: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Figure 4. Influence of the concentration of interfacial defects (Ni) at the TiO2/CH3NH3SnI3 (IL1) and CH3NH3SnI3/HTL (IL2) interfaces: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Crystals 12 00699 g004
Figure 5. Influence of the concentration of interfacial defects (Ni) at the TiO2/CH3NH3SnI3 (IL1) interface: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Figure 5. Influence of the concentration of interfacial defects (Ni) at the TiO2/CH3NH3SnI3 (IL1) interface: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Crystals 12 00699 g005
Figure 6. Influence of the concentration of interfacial defects (Ni) at the CH3NH3SnI3/HTL (IL2) interface: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Figure 6. Influence of the concentration of interfacial defects (Ni) at the CH3NH3SnI3/HTL (IL2) interface: (a) JSC, (b) VOC, (c) FF, (d) PCE.
Crystals 12 00699 g006
Figure 7. Perovskite solar cells efficiency depending on the choice of a back metal contact.
Figure 7. Perovskite solar cells efficiency depending on the choice of a back metal contact.
Crystals 12 00699 g007
Figure 8. Current-voltage characteristics of perovskite solar cells with different hole-transport layers and a Pt back contact.
Figure 8. Current-voltage characteristics of perovskite solar cells with different hole-transport layers and a Pt back contact.
Crystals 12 00699 g008
Figure 9. Influence of temperature on JSC (a), VOC (b), FF (c), PCE (d) for the perovskite solar cells with different hole-transport layers.
Figure 9. Influence of temperature on JSC (a), VOC (b), FF (c), PCE (d) for the perovskite solar cells with different hole-transport layers.
Crystals 12 00699 g009
Figure 10. Current-voltage characteristics of perovskite solar cells with different hole-transport layers.
Figure 10. Current-voltage characteristics of perovskite solar cells with different hole-transport layers.
Crystals 12 00699 g010
Figure 11. Current-voltage characteristic of a simulated photovoltaic cell FTO/TiO2/CH3NH3SnI3/Cu2O/Au with different thicknesses of the absorbing layer.
Figure 11. Current-voltage characteristic of a simulated photovoltaic cell FTO/TiO2/CH3NH3SnI3/Cu2O/Au with different thicknesses of the absorbing layer.
Crystals 12 00699 g011
Figure 12. Influence of the perovskite absorbing layer thickness on JSC (a), VOC (b), FF (c), PCE (d).
Figure 12. Influence of the perovskite absorbing layer thickness on JSC (a), VOC (b), FF (c), PCE (d).
Crystals 12 00699 g012
Table 1. Parameters of solar cells.
Table 1. Parameters of solar cells.
ParametersFTOTiO2CH3NH3SnI3Spiro-OMeTADPEDOT:PSSCu2O
Thickness (nm)50050 *300–1300 *50 *50 *50 *
Band gap (eV)3.503.201.303.061.802.17
Electron affinity (eV)4.004.264.172.053.403.20
Relative dielectric permittivity9.009.008.203.0018.007.10
Conduction band effective density of states (cm−3)2.20 × 10182.20 × 10181 × 10182.20 × 10182.20 × 10182.00 × 1017
Valence band effective density of states (cm−3)1.80 × 10191.80 × 10191 × 10181.80 × 10191.80 × 10191.10 × 1019
Electron thermal velocity (cm/s)107107107107107107
Hole thermal velocity (cm/s)107107107107107107
Electron mobility (cm2/Vs)20.0020.001.602.00 × 10−44.50 × 10−2200.00
Hole mobility (cm2/Vs)10.0010.001.602.00 × 10−44.50 × 10−280.00
Shallow donor density ND (cm−3)2.00 × 101910180000
Shallow acceptor density NA (cm−3)001.00 × 10141.00 × 10181.00 × 10201.00 × 1018
References[53,82][83][82,83][53][84,85][53,83]
* In this study.
Table 2. Work function of a back contact metal.
Table 2. Work function of a back contact metal.
Back Contact MetalsAlTiCrAgNiCuCAuPt
Work Function (eV)4.284.334.504.504.604.705.005.105.65
References[98][99][98][97][98][97][68][98][99]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Omarova, Z.; Yerezhep, D.; Aldiyarov, A.; Tokmoldin, N. In Silico Investigation of the Impact of Hole-Transport Layers on the Performance of CH3NH3SnI3 Perovskite Photovoltaic Cells. Crystals 2022, 12, 699. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050699

AMA Style

Omarova Z, Yerezhep D, Aldiyarov A, Tokmoldin N. In Silico Investigation of the Impact of Hole-Transport Layers on the Performance of CH3NH3SnI3 Perovskite Photovoltaic Cells. Crystals. 2022; 12(5):699. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050699

Chicago/Turabian Style

Omarova, Zhansaya, Darkhan Yerezhep, Abdurakhman Aldiyarov, and Nurlan Tokmoldin. 2022. "In Silico Investigation of the Impact of Hole-Transport Layers on the Performance of CH3NH3SnI3 Perovskite Photovoltaic Cells" Crystals 12, no. 5: 699. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12050699

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