Next Article in Journal
Compact Wideband Four-Port MIMO Antenna for Sub-6 GHz and Internet of Things Applications
Previous Article in Journal
Towards a Highly Efficient ZnO Based Nanogenerator
Previous Article in Special Issue
Modeling and Simulation of Tin Sulfide (SnS)-Based Solar Cell Using ZnO as Transparent Conductive Oxide (TCO) and NiO as Hole Transport Layer (HTL)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Power-Conversion Efficiency of Lead-Free Perovskite Solar Cells: A Theoretical Investigation

1
Department of Chemistry, Faculty of Science and Arts, and Promising Centre for Sensors and Electronic Devices (PCSED), Najran University, Najran 11001, Saudi Arabia
2
Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA
3
Department of Applied Sciences, Galgotias College of Engineering and Technology, Greater Noida 201306, India
4
Institute of Advanced Materials, IAAM, Gammalkilsvägen 18, 590 53 Ulrika, Sweden
5
Photonics and Photovoltaic Research Lab, Department of Physics and Material Science, Madan Mohan Malaviya University of Technology, Gorakhpur 273010, India
6
Department of Electrical Engineering, College of Engineering, Najran University, Najran 11001, Saudi Arabia
7
Department of Materials Science, University of Patras, 265 04 Patras, Greece
*
Authors to whom correspondence should be addressed.
Adjunct Professor at the Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA.
These authors contributed equally and should be considered as first author.
Micromachines 2022, 13(12), 2201; https://0-doi-org.brum.beds.ac.uk/10.3390/mi13122201
Submission received: 10 October 2022 / Revised: 19 November 2022 / Accepted: 9 December 2022 / Published: 12 December 2022
(This article belongs to the Special Issue Advances in Perovskite Devices)

Abstract

:
Solar cells based on lead-free perovskite have demonstrated great potential for next-generation renewable energy. The SCAPS-1D simulation software was used in this study to perform novel device modelling of a lead-free perovskite solar cell of the architecture ITO/WS2/CH3NH3SnI3/P3HT/Au. For the performance evaluation, an optimization process of the different parameters such as thickness, bandgap, doping concentration, etc., was conducted. Extensive optimization of the thickness and doping density of the absorber and electron transport layer resulted in a maximum power-conversion efficiency of 33.46% for our designed solar cell. Because of the short diffusion length and higher defect density in thicker perovskite, an absorber thickness of 1.2 µm is recommended for optimal solar cell performance. Therefore, we expect that our findings will pave the way for the development of lead-free and highly effective perovskite solar cells.

1. Introduction

A photovoltaic solar cell is a promising renewable and non-polluted source of energy [1]. In terms of research and development, halide-based perovskite solar cells (PSCs) are the fastest growing photovoltaic technology [2,3,4]. The first potential report of halide-based perovskite solar cells appeared in 2009. In such studies, organic−inorganic (CH3NH3PbI3) hybrid perovskite has been used as a light-sensitizer in a dye-sensitized solar cell and a power conversion efficiency of 3.8% has been achieved. The ability of halide perovskites to operate not only as powerful light absorbers but also as efficient electron and hole conductors was demonstrated in the literature [5,6,7] with power conversion efficiencies of 10.9% and 9.7%, respectively. These findings led to the start of a global effort to improve perovskite cell efficiency beyond 20%. The Pb-based perovskite solar cells (PSCs) have gained a lot of potential over silicon-based solar cells due to their ease of fabrication and low cost. Lead-based halide perovskite also exhibits several desirable optoelectronic properties such as a high absorption coefficient, a comparatively moderate bandgap, and a high charge diffusion length [7,8,9].
Considering these great advances towards high performance, the toxicity and lack of stability of the absorber layer in lead-based PSCs remains a serious problem for their widespread commercialization [10,11,12]. One potential way to address these concerns is to replace the hazardous toxic lead in the PSC’s CH3NH3PbI3 absorber layer with a suitable non-toxic element. The methylammonium tin iodide CH3NH3SnI3 has, amongst others, become a viable alternative due to its eco-friendliness, as well as its smaller effective mass of holes [13], excellent mobility [14], narrower bandgap, and wider visible absorption spectrum [15].
The study of the characteristics of perovskite solar cell materials, as well as their control through accurate modelling, can lead to the production of efficient and cost-effective solar cells [16,17]. To build cost-effective, efficient, and lead-free PSCs, the electrical and optical properties of CH3NH3SnI3 must be explored using modelling before dealing with very complex fabrication processes. The main objective of this research is to design lead-free PSCs with an improved efficiency. In this study, we have used a solar cell capacitance simulator in one dimension (SCAPS-1D) program. To get the highest feasible PCE, different aspects were optimized. Initially, we have calibrated the device structure with previously reported experimental work [18]. Then, an investigation of the proposed device structure ITO/WS2/CH3NH3SnI3/P3HT/Au has been performed.
There has been a lot of interest in WS2 as ETL due to its potential as an electron transport layer in thin film solar cells [19]. It is easily available at a low cost and is less hazardous than other transition metal dichalcogenides (TMDC) compounds. The development of WS2 in thin film solar cells is still in its infancy compared with other photovoltaic materials [20]. Due to its superior optoelectronic properties, tungsten disulfide (WS2) has become the primary material for thin film solar cells. Its tunable bandgap is an essential feature that is usually ignored. WS2 has a large direct bandgap (>2 eV) and a small indirect bandgap (~1.3 eV) [20,21,22].
Furthermore, it exhibits high carrier mobility, good conductivity, native n-type semiconducting characteristics [23], and excellent electron conduction properties [24]. Moreover, it can be deposited through a solution process or by sputtering at low temperature [25]. Homo-polymer poly (3-hexylthiophene) (P3HT) as an HTL is one of the few viable choices for commercial Organic Photovoltaic (OPV). Its application in large-area, roll-to-roll printed solar cells has already been amply shown [26,27]. Additionally, the semi-crystalline structure of P3HT, in contrast to more amorphous polymers, is nearly unique in setting an acceptable morphological length-scale for bulk heterojunction OPV from a variety of solvents and processing conditions, as well as giving it outstanding charge transport properties [28,29]. P3HT has attracted great interest as a polymeric hole-selective material for perovskite solar cells due to its low cost [30], wide band-gap [31], relatively high hole mobility [32], high thermal stability [33], scalable solution processability [34], robust hydrophobicity, and oxygen impermeability [35].
An analysis of the impact of CH3NH3SnI3 as an absorber layer with different thicknesses and bandgap is performed. After this, the impact of ETL (WS2) with varying thicknesses and doping concentrations on the performance of PSCs has been investigated. At the end of this simulation analysis, the optimized device structure with the highest efficiency of 33.36% is obtained.

2. Device Structure and Simulation Methodology

The perovskite solar cell device structure consists of Glass/ITO/WS2/CH3NH3SnI3/P3HT/Au, as shown in Figure 1. In this device design, ITO serves as an electron transport layer, which is covered by an N-type (WS2) material. The organic and inorganic intrinsic perovskite CH3NH3SnI3 serves as the absorber layer, while the p-type P3HT serves as the hole transporting layer on which the contact is formed.
The SCAPS-1D program was used to simulate a solar cell device. It is freely available to the research community. Solar radiation of AM_1.5 spectrum (1 kW/m2) is illuminated from the ITO window layer of the perovskite device structure. To start simulation one has to insert the material parameters by selecting the option ‘set problem’, click on ‘add the layer’, then insert the electrical and optical properties of the suitable material, such as thickness, electron affinity, dielectric permittivity (relative), bandgap, electron and hole thermal velocity (cm/s), electron and hole mobility (cm2/V-s), donor density ND (cm−3), acceptor density NA (cm−3). Then, insert the absorption coefficient (α) in the absorption model and recombination model, add the recombination details, and include the defects of the material [36,37]. Similarly, add the layers of the different materials and make the different types of solar cells. After adding the parameters as mentioned above, one has to check the illumination of light and whether or not light is passing through glass substrate (FTO, AZO, ITO, etc.) and through ETL layer which has less thickness (to allow the whole light to be absorbed in the absorber layer). Check whether the connections of the voltage are connected properly or not. Then, set the working point values to room temperature (300 K) and set the frequency. Add the series and shunt resistance values, ideally the series resistance would be low and the shunt resistance would be very high.
The SCAPS-1D simulation software is superior to other simulation software because it gives good consistency between experimental and simulated results. In the present simulation, series and shunt resistances are taken as 1   Ω and 106 Ω   respectively. The effect of the dangling bond on the interface of the materials has been ignored. Benchmarking is achieved by matching the correct defect design before starting the simulation, therefore, the simulated outcome matches the corresponding experimentally tested values.
The parameters used in the simulation for the perovskite solar cell with device structure Glass/ITO/WS2/CH3NH3SnI3/P3HT/Au are listed in Table 1.

3. Results and Discussion

3.1. The Effect of the Thickness and Doping Concentration of the Absorber Layer

The impact of the active perovskite layer thickness on the efficiency of the solar device will be explored in this section. Absorber layers play a vital role in the improvement of the performance of the solar device [39,40]. All parameters, such as bandgap, thickness, and doping concertation, play an important role in optimizing performance. Figure 2 illustrates the variation of the electrical parameters with thickness, such as Voc Jsc and FF, η. The simulated outcome shows that solar cell parameters are highly dependent on the thickness of the perovskite layer. The electrical parameters, such as Jsc, and η, increase with the increase in the thickness of the perovskite layer while the FF and Voc decrease with a further rise in thickness. When increasing the absorber layer thickness, it has been observed that the Jsc, and eta performance improve but the Voc performance degrades.
Since, as the absorber layer thickness increases, more photons are absorbed, and as photons penetrate deeper into the absorber layer, more electron-hole pairs are produced which increases the performance of the device [41]. The decrease in Voc is caused by an increase in the dark saturation current, which enhances charge carrier recombination [42]. This is addressed by the photo-generated current and dark saturation current’s dependence on open-circuit voltage, which is expressed as [42,43]
V oc = kT q Ln J sc J 0 + 1
Here Jsc represents the photo-generated current density, kT/q represents the thermal voltage, and J0 represents the saturation current density
The JV characteristics have been recorded by varying the thickness of the absorber layer, as shown in Figure 3a. Current density vs. voltage characteristics show an increase with increasing absorber layer thickness. As the thickness of the absorber layer increases, the area under the curve also enhances, which results in an increase in the Jsc values because more photons fall on it [40,41,42].
The external quantum efficiency spectra of the proposed device are shown in Figure 3b. The relations between quantum efficiency (QE) and wavelength curve show that the % QE increases with varying the absorber layer thickness. As the thickness of the absorber layer increases, more light is absorbed, resulting in a large number of carriers. These excess carriers led to the increase in the Jsc and PCE of the perovskite device [41,42]. Firstly, QE increases rapidly (up to 1.2 m) with the increase in the thickness of the absorber layer and after it gets saturated (as shown in Figure 3b). Considering the short diffusion length and higher defect density in thicker perovskite, a thickness of 1.2 µm for the absorber is appropriate for obtaining optimal solar cell performance.

3.2. Effect of Absorber’s Bandgap on Solar Cell Performance

In this present section, the effect of the bandgap on the solar cell performance has been investigated, which is shown in Figure 4a. The energy bandgap of the CH3NH3SnI3 has been varied from 1.0 eV to 1.5 eV and the corresponding change in the performance is noted. FF and Voc increase with the bandgap of absorber materials while Jsc and PCE decrease with an increasing bandgap of absorber materials. The FF and open circuit voltage is proportional to the active material’s bandgap. As the bandgap widens, so does the open circuit voltage [44].
In Figure 4a the sharp decrease in PCE and Jsc was observed with the increase in the bandgap of the absorber layer. Initially, FF increases with the bandgap of the absorber layer to the particular bandgap of 1.3 eV, but with further increase, it starts to decrease.
The QE spectra of varying the bandgap of the CH3NH3SnI3 are shown in Figure 4b. QE vs. wavelength curve demonstrates that the quantum efficiency continues to increase with increasing the bandgap of the materials. Initially, maximum QE is observed at a bandgap of 1.0 eV, while with the increase in the bandgap of the absorber layer, QE goes on decreasing, which results in a decrease in the PCE and Jsc of the perovskite device.

3.3. Effect of Doping Density of ETL

The impact of the doping density of the WS2 ETL on the functional parameters of perovskite cells has been investigated. The doping density of the ETL layer has been varied from 1015 cm−3 to 1022 cm−3, as shown in Figure 5a. Initially, with increases in the doping density of the ETL material, PCE, Voc, and FF go on increasing to a certain value of 1018 cm−3 and with a further increase in the doping density, Voc and FF become saturated while PCE slightly decreases and then becomes constant.
The QE vs. wavelength curve shows almost constant variation in quantum efficiency with an increase in the doping concentration of the electron transport material (ETM) (Figure 5b). The QE with varied doping density at a wavelength range of ~400–900 nm is observed to be constant, while a slight change can be seen below 400 nm, as shown in Figure 5b.
The doping concentration of the HTL has little impact on Jsc because photogeneration occurs mostly in the absorber layer [45].
The maximum performance of the proposed device structure Glass/ITO/WS2/CH3NH3SnI3/P3HT/Au is observed at a doping density of 1018 cm−3. Therefore, 1018 cm−3 is taken as the optimum doping density for WS2 of this device structure.

3.4. Effect of Doping Density and Bandgap of HTL

The effect of the bandgap of the HTL on the performance of the perovskite solar cell has been investigated, which is shown in Figure 6a. Increases in FF, Voc, and PCE are shown with an increase in the HTL’s bandgap, however a substantial decrease in Jsc (mA/cm2) is seen. Because of the large bandgap, more energy is required to transfer an electron from the valence band to the conduction band.
Furthermore, the influence of the P3HT’s doping density on the perovskite cell’s functional characteristics has been studied. The doping density of the HTL layer has been varied from 1015 cm−3 to 1022 cm−3, as shown in Figure 6b. Initially, as the doping density of the ETL material increases, PCE, Voc, and FF increase to a certain threshold prior to getting saturated. While a small drop in Jsc (mA/cm2) has been noticed with a rise in P3HT doping concentration. Moreover, photogeneration occurs mainly in the Ch3NH3SnI3 layer, and as a result, the doping concentration of the HTL has little effect on Jsc [45].
From the graph, it is concluded that the maximum performance of the device structure glass/ITO/WS2/CH3NH3SnI3/P3HT/Au is at a doping density of 1018 cm−3. As a result, 1018 cm−3 is chosen as the best doping density for P3HT in this device configuration.

3.5. Optimized Performance

The performance of the lead-free perovskite solar cell with the device structure Glass/ITO/WS2/CH3NH3SnI3/P3HT have been studied. The optimized performance of the CH3NH3SnI3 perovskite solar cell using WS2 as an electron transporting layer is shown in Figure 7a,b. To optimize the performance, different parameters such as thickness, doping concentration, and bandgap have been varied. In this study, room temperature (300 K) is used for optimization. The best PCE with optimized parameters has been observed as 33.36%.
The light transmittance of the substrate, which is known to be over 90% for glass substrates, has a substantial impact on the efficiency of perovskite solar cells. Solar cell power is primarily affected by optical losses via a reduction in the short-circuit current. Light that could have produced an electron-hole pair but does not because it is reflected off the front surface or because it is not absorbed by the solar cell is referred to as an optical loss. The ideal situation would be for visible light (350–780 nm) to be totally absorbed, since it has enough energy to produce electron–hole pairs [46,47]. There are a number of measures to lower optical losses, including coating the top surface of the cell with anti-reflective materials, reducing top contact coverage of the cell surface, increasing absorption by thickening the active layer, reducing reflection by texturing the surface, increasing the optical path by combining surface texturing, and light trapping, among others [48,49].
The potential of these materials, such as WS2 and P3HT, to act as electron and hole transport layers in thin film solar cells could be one of the reasons for the high-power conversion efficiency [20,21,22]. Another reason could be optical effects caused by multi-reflection within the device between layers, which could result in the active layer absorbing more photons and producing larger short-circuit currents [50].
The simulated and experimental reported results are listed in Table 2. Our simulated study, according to the tabulated results, is in good agreement with experimentally reported work by various researchers working in this field.

4. Conclusions

In the present study, a novel device structure composed of Glass/ITO/WS2/CH3NH3SnI3/P3HT was investigated to boost solar cell performance. The lead-free perovskite solar cell can have good efficiency by selecting a buffer layer of wide bandgap materials that can transmit more photons to the absorber layer. Here, the impact of the various parameters on the performance of the perovskite solar cell has been studied. The optimized performance of the proposed device structure such as Voc, FF, Jsc, and PCE are 1.0997 V, 37.1778 mA, 81.59%, and 33.36% respectively. It can also be concluded that with the increase in the ETL doping concentration, the suggested structure has good performance and can compete with the existing lead-based perovskite solar cells. The present study would serve as a beneficial roadmap in developing lead-free, high-efficiency perovskite solar cells.

Author Contributions

A.U.: Data analyzation, writing—review and editing; S.: Calculations, data analyzation; writing—review and editing; P.K.S.: Data analyzation; H.A.: Calculations, data analyzation, D.K.D.: Data analyzation; A.A.I. and M.A.M.A.: Data analyzation; S.B.: Calculations, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful to the Deanship of Scientific Research at Najran University, Najran, Kingdom of Saudi Arabia for funding this work under the National Research Priority funding program, grant no. NU/NRP/SERC/11/1.

Acknowledgments

The authors are thankful to the Deanship of Scientific Research at Najran University, Najran, Kingdom of Saudi Arabia for funding this work under the National Research Priority funding program, grant no. NU/NRP/SERC/11/1. The authors also would like to thank Mark Burgelman for developing SCAPS-1D software and making it freely available to the entire research community.

Conflicts of Interest

The authors declared no potential conflict of interest.

References

  1. Mahjabin, S.; Haque, M.M.; Sobayel, K.; Jamal, M.S.; Islam, M.A.; Selvanathan, V.; Assaifan, A.K.; Alharbi, H.F.; Sopian, K.; Amin, N.; et al. Perceiving of Defect Tolerance in Perovskite Absorber Layer for Efficient Perovskite Solar Cell. IEEE Access 2020, 8, 106346–106353. [Google Scholar] [CrossRef]
  2. Ikram, M.; Malik, R.; Raees, R.; Imran, M.; Wang, F.; Ali, S.; Khan, M.; Khan, Q.; Maqbool, M. Recent advancements and future insight of lead-free non-toxic perovskite solar cells for sustainable and clean energy production: A review. Sustain. Energy Technol. Assess. 2022, 53, 102433. [Google Scholar] [CrossRef]
  3. Bello, O.O.; Emetere, M.E. Progress and limitation of lead-free inorganic perovskites for solar cell application. Sol. Energy 2022, 243, 370–380. [Google Scholar] [CrossRef]
  4. Jayan, K.D.; Sebastian, V. Comparative Study on the Performance of Different Lead-Based and Lead-Free Perovskite Solar Cells. Adv. Theory Simul. 2021, 4, 2100027. [Google Scholar] [CrossRef]
  5. Akhiro, K.; Kenjiro, T.; Yasuo, S.; Tsutomu, M. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar]
  6. Hima, A.; Lakhdar, N. Enhancement of Efficiency and Stability of CH3NH3GeI3 Solar Cells with CuSbS2. Opt. Mater. 2020, 99, 109607. [Google Scholar] [CrossRef]
  7. Son, D.Y.; Im, J.H.; Kim, H.S.; Park, N.G. 11% Efficient Perovskite Solar Cell Based on ZnO Nanorods: An Effective Charge Collection System. J. Phys. Chem. C 2014, 118, 16567–16573. [Google Scholar] [CrossRef]
  8. Yang, W.S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. High-Performance Photovoltaic Perovskite Layers Fabricated through Intramolecular Exchange. Science 2015, 348, 1234–1237. [Google Scholar] [CrossRef]
  9. Hima, A.; Lakhdar, N.; Benhaoua, B.; Saadoune, A.; Kemerchou, I.; Rogti, F. An Optimized Perovskite Solar Cell Designs for High Conversion Efficiency. Superlattices Microstruct. 2019, 129, 240–246. [Google Scholar] [CrossRef]
  10. Wang, Q.; Phung, N.; Di Girolamo, D.; Vivo, P.; Abate, A. Enhancement in Lifespan of Halide Perovskite Solar Cells. Energy Environ. Sci. 2019, 12, 865–886. [Google Scholar] [CrossRef]
  11. Shi, Z.; Jayatissa, A.H. Perovskites-Based Solar Cells: A Review of Recent Progress, Materials and Processing Methods. Materials 2018, 11, 729. [Google Scholar] [CrossRef] [PubMed]
  12. Conings, B.; Drijkoningen, J.; Gauquelin, N.; Babayigit, A.; D’Haen, J.; D’Olieslaeger, L.; Ethirajan, A.; Verbeeck, J.; Manca, J.; Mosconi, E.; et al. Intrinsic Thermal Instability of Methylammonium Lead Trihalide Perovskite. Adv. Energy Mater. 2015, 5, 1500477. [Google Scholar] [CrossRef]
  13. Kagan, C.R.; Mitzi, D.B.; Dimitrakopoulos, C.D. Organic-Inorganic Hybrid Materials as Semiconducting Channels in Thin-Film Field-Effect Transistors. Science 1999, 286, 945–947. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, L.; Xu, L. Theoretical and Experimental Research Base on the Tin Iodide Organic–Inorganic Hybrid Perovskite (CH3NH3SnI3) Tetragonal and Orthorhombic Phases for Photovoltaics. Sci. Adv. Mater. 2018, 10, 1519–1527. [Google Scholar] [CrossRef]
  15. Umari, P.; Mosconi, E.; De Angelis, F. Relativistic GW Calculations on CH3NH3PbI3 and CH3NH3SnI3 Perovskites for Solar Cell Applications. Sci. Rep. 2014, 4, 4467. [Google Scholar] [CrossRef] [Green Version]
  16. 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]
  17. 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]
  18. 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]
  19. Ballif, C.; Regula, M.; Levy, F. Optical and Electrical Properties of Semiconducting WS2 Thin Films: From Macroscopic to Local Probe Measurements. Sol. Energy Mater. Sol. Cells 1999, 57, 189–207. [Google Scholar] [CrossRef]
  20. Bruno, A.; Borriello, C.; Haque, S.A.; Minarini, C.; Di Luccio, T. Ternary Hybrid Systems of P3HT-CdSe-WS2 Nanotubes for Photovoltaic Applications. Phys. Chem. Chem. Phys. 2014, 16, 17998–18003. [Google Scholar] [CrossRef]
  21. Macchia, E.; Zak, A.; Picca, R.A.; Manoli, K.; Di Franco, C.; Cioffi, N.; Scamarcio, G.; Tenne, R.; Torsi, L. Improved Performance P-Type Polymer (P3HT)/n-Type Nanotubes (WS2) Electrolyte Gated Thin-Film Transistor. MRS Adv. 2017, 2, 3865–3872. [Google Scholar] [CrossRef]
  22. Najafi, L.; Romano, V.; Oropesa-Nuñez, R.; Prato, M.; Lauciello, S.; D’Angelo, G.; Bellani, S.; Bonaccorso, F. Hybrid Organic/Inorganic Photocathodes Based on WS 2 Flakes as Hole Transporting Layer Material. Small Struct. 2021, 2, 2000098. [Google Scholar] [CrossRef]
  23. Hankare, P.P.; Manikshete, A.H.; Sathe, D.J.; Chate, P.A.; Patil, A.A.; Garadkar, K.M. WS2 Thin Films: Opto-Electronic Characterization. J. Alloy. Compd. 2009, 479, 657–660. [Google Scholar] [CrossRef]
  24. Li, S.; Chen, Z.; Zhang, W. Dye-Sensitized Solar Cells Based on WS2 Counter Electrodes. Mater. Lett. 2012, 72, 22–24. [Google Scholar] [CrossRef]
  25. Lignier, O.; Couturier, G.; Tedd, J.; Gonbeau, D.; Salardenne, J. Photoactivity Enhancement of WS2 Sputtered Thin Films by Use of Nickel. Thin Solid Film. 1997, 299, 45–52. [Google Scholar] [CrossRef]
  26. Chatterjee, S.; Jinnai, S.; Ie, Y. Nonfullerene Acceptors for P3HT-Based Organic Solar Cells. J. Mater. Chem. A 2021, 9, 18857–18886. [Google Scholar] [CrossRef]
  27. Manceau, M.; Rivaton, A.; Gardette, J.L.; Guillerez, S.; Lemaître, N. Light-Induced Degradation of the P3HT-Based Solar Cells Active Layer. Sol. Energy Mater. Sol. Cells 2011, 95, 1315–1325. [Google Scholar] [CrossRef]
  28. Holliday, S.; Ashraf, R.S.; Wadsworth, A.; Baran, D.; Yousaf, S.A.; Nielsen, C.B.; Tan, C.H.; Dimitrov, S.D.; Shang, Z.; Gasparini, N.; et al. High-Efficiency and Air-Stable P3HT-Based Polymer Solar Cells with a New Non-Fullerene Acceptor. Nat. Commun. 2016, 7, 11585. [Google Scholar] [CrossRef] [Green Version]
  29. Mor, G.K.; Kim, S.; Paulose, M.; Varghese, O.K.; Shankar, K.; Basham, J.; Grimes, C.A. Visible to Near-Infrared Light Harvesting in TiO2 Nanotube Array-P3HT Based Heterojunction Solar Cells. Nano Lett. 2009, 9, 4250–4257. [Google Scholar] [CrossRef]
  30. Jacoby, M. Low-Cost Polymer Works Well in Perovskite Solar Cells. Chem. Eng. News 2019, 97, 11. [Google Scholar]
  31. Hrostea, L.; Girtan, M.; Mallet, R.; Leontie, L. Optical and Morphological Properties of P3HT and P3HT: PCBM Thin Films Used in Photovoltaic Applications. IOP Conf. Ser. Mater. Sci. Eng. 2018, 374, 12015. [Google Scholar] [CrossRef] [Green Version]
  32. Janasz, L.; Chlebosz, D.; Gradzka, M.; Zajaczkowski, W.; Marszalek, T.; Müllen, K.; Ulanski, J.; Kiersnowski, A.; Pisula, W. Improved Charge Carrier Transport in Ultrathin Poly (3-Hexylthiophene) Films via Solution Aggregation. J. Mater. Chem. C 2016, 4, 11488–11498. [Google Scholar] [CrossRef]
  33. Rodrigues, A.; Castro, M.C.R.; Farinha, A.S.F.; Oliveira, M.; Tomé, J.P.C.; Machado, A.V.; Raposo, M.M.M.; Hilliou, L.; Bernardo, G. Thermal Stability of P3HT and P3HT: PCBM Blends in the Molten State. Polym. Test. 2013, 32, 1192–1201. [Google Scholar] [CrossRef]
  34. Brabec, C.J.; Durrant, J.R. Solution-Processed Organic Solar Cells. MRS Bull. 2008, 33, 670–675. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, X.; Ederth, T.; Inganäs, O. In Situ Wilhelmy Balance Surface Energy Determination of Poly (3-Hexylthiophene) and Poly (3, 4-Ethylenedioxythiophene) during Electrochemical Doping–Dedoping. Langmuir 2006, 22, 9287–9294. [Google Scholar] [CrossRef]
  36. Jhuma, F.A.; Shaily, M.Z.; Rashid, M.J. Towards High-Efficiency CZTS Solar Cell through Buffer Layer Optimization. Mater. Renew. Sustain. Energy 2019, 8, 6. [Google Scholar] [CrossRef] [Green Version]
  37. Sadanand; Dwivedi, D.K. Modeling of Photovoltaic Solar Cell Based on CuSbS2 Absorber for the Enhancement of Performance. IEEE Trans. Electron Devices 2021, 68, 1121–1128. [Google Scholar] [CrossRef]
  38. Sobayel, K.; Akhtaruzzaman, M.; Rahman, K.S.; Ferdaous, M.T.; Al-Mutairi, Z.A.; Alharbi, H.F.; Alharthi, N.H.; Karim, M.R.; Hasmady, S.; Amin, N. A Comprehensive Defect Study of Tungsten Disulfide (WS2) as Electron Transport Layer in Perovskite Solar Cells by Numerical Simulation. Results Phys. 2019, 12, 1097–1103. [Google Scholar] [CrossRef]
  39. Lin, L.; Jiang, L.; Li, P.; Fan, B.; Qiu, Y.; Yan, F. Simulation of Optimum Band Structure of HTM-Free Perovskite Solar Cells Based on ZnO Electron Transporting Layer. Mater. Sci. Semicond. Process. 2019, 90, 1–6. [Google Scholar] [CrossRef]
  40. Sadanand; Dwivedi, D.K. Numerical Modeling for Earth-Abundant Highly Efficient Solar Photovoltaic Cell of Non-Toxic Buffer Layer. Opt. Mater. 2020, 109, 110409. [Google Scholar] [CrossRef]
  41. Sadanand; Dwivedi, D.K. Theoretical Investigation on Enhancement of Output Performance of CZTSSe Based Solar Cell. Sol. Energy 2019, 193, 442–451. [Google Scholar] [CrossRef]
  42. Ouslimane, T.; Et-taya, L.; Elmaimouni, L.; Benami, A. Impact of Absorber Layer Thickness, Defect Density, and Operating Temperature on the Performance of MAPbI3 Solar Cells Based on ZnO Electron Transporting Material. Heliyon 2021, 7, e06379. [Google Scholar] [CrossRef] [PubMed]
  43. Singh, P.; Ravindra, N.M. Temperature Dependence of Solar Cell Performance—An Analysis. Sol. Energy Mater. Sol. Cells 2012, 101, 36–45. [Google Scholar] [CrossRef]
  44. Polman, A.; Atwater, H.A. Photonic Design Principles for Ultrahigh-Efficiency Photovoltaics. Nat. Mater. 2012, 11, 174–177. [Google Scholar] [CrossRef]
  45. Jeyakumar, R.; Bag, A.; Nekovei, R.; Radhakrishnan, R. Influence of Electron Transport Layer (TiO2) Thickness and Its Doping Density on the Performance of CH3NH3PbI3-Based Planar Perovskite Solar Cells. J. Electron. Mater. 2020, 49, 3533–3539. [Google Scholar] [CrossRef] [Green Version]
  46. Shaikh, S.F.; Kwon, H.C.; Yang, W.; Mane, R.S.; Moon, J. Performance Enhancement of Mesoporous TiO2-Based Perovskite Solar Cells by ZnS Ultrathin-Interfacial Modification Layer. J. Alloy. Compd. 2018, 738, 405–414. [Google Scholar] [CrossRef]
  47. Kakherskyi, S.I.; Diachenko, O.V.; Opanasyuk, N.M.; Opanasyuk, A.S. Optical Losses in Glass/ITO (ZnO)/CdS/Cu2ZnSn(SxSe1−x)4 Solar Cells with Different Kesterite Composition. Trans. Electr. Electron. Mater. 2022, 23, 552–562. [Google Scholar] [CrossRef]
  48. Abdullah, M.F.; Alghoul, M.A.; Naser, H.; Asim, N.; Ahmadi, S.; Yatim, B.; Sopian, K. Research and development efforts on texturization to reduce the optical losses at front surface of silicon solar cell. Renew. Sustain. Energy Rev. 2016, 66, 380–398. [Google Scholar] [CrossRef]
  49. Selj, J.H.; Mongstad, T.T.; Søndenå, R.; Marstein, E.S. Reduction of optical losses in colored solar cells with multilayer antireflection coatings. Sol. Energy Mater. Sol. Cells 2011, 95, 2576–2582. [Google Scholar] [CrossRef]
  50. Gabr, A.M.; Trojnar, A.H.; Wilkins, M.; Hall, T.J.; Kleiman, R.N.; Hinzer, K. Optimization of Anti-Reflection Coatings for Bifacial Solar Cells with Upconversion Layers. In Proceedings of the 2014 IEEE 40th Photovoltaic Specialist Conference (PVSC), Denver, CO, USA, 8–13 June 2014; pp. 2230–2233. [Google Scholar] [CrossRef]
  51. Li, M.; Xie, Y.M.; Xu, X.; Huo, Y.; Tsang, S.W.; Yang, Q.D.; Cheng, Y. Comparison of Processing Windows and Electronic Properties between CH3NH3PbI3 Perovskite Fabricated by One-Step and Two-Step Solution Processes. Org. Electron. 2018, 63, 159–165. [Google Scholar] [CrossRef]
  52. Kim, D.I.; Lee, J.W.; Jeong, R.H.; Boo, J.H. A High-efficiency and Stable Perovskite Solar Cell Fabricated in Ambient Air Using a Polyaniline Passivation Layer. Sci. Rep. 2022, 12, 694. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, D.; Traverse, C.J.; Chen, P.; Elinski, M.; Yang, C.; Wang, L.; Young, M.; Lunt, R.R. Aqueous-Containing Precursor Solutions for Efficient Perovskite Solar Cells. Adv. Sci. 2018, 5, 1700484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kumar, M.; Raj, A.; Kumar, A.; Anshul, A. An Optimized Lead-Free Formamidinium Sn-Based Perovskite Solar Cell Design for High Power Conversion Efficiency by SCAPS Simulation. Opt. Mater. 2020, 108, 110213. [Google Scholar] [CrossRef]
Figure 1. Schematic device structure of perovskite solar cell.
Figure 1. Schematic device structure of perovskite solar cell.
Micromachines 13 02201 g001
Figure 2. Effect of the perovskite layer thickness on solar cell parameters (Jsc, Voc, FF, and η].
Figure 2. Effect of the perovskite layer thickness on solar cell parameters (Jsc, Voc, FF, and η].
Micromachines 13 02201 g002
Figure 3. (a) Current density vs. voltage characteristics and (b) quantum efficiency vs. wavelength curve with varied thickness of the absorber layer.
Figure 3. (a) Current density vs. voltage characteristics and (b) quantum efficiency vs. wavelength curve with varied thickness of the absorber layer.
Micromachines 13 02201 g003
Figure 4. (a) Effect of bandgap of perovskite on solar cell parameters (b) QE vs. wavelength.
Figure 4. (a) Effect of bandgap of perovskite on solar cell parameters (b) QE vs. wavelength.
Micromachines 13 02201 g004
Figure 5. (a) Effect of doping density of the ETL on solar cell parameters (Jsc, Voc, FF, and η ] (b) QE vs. wavelength.
Figure 5. (a) Effect of doping density of the ETL on solar cell parameters (Jsc, Voc, FF, and η ] (b) QE vs. wavelength.
Micromachines 13 02201 g005
Figure 6. (a) Effect of bandgap of P3HT on solar cell parameters (Jsc, Voc, FF, and η ] (b) Effect of the HTL doping density on the parameters of solar cell (Jsc, Voc, FF, and η ).
Figure 6. (a) Effect of bandgap of P3HT on solar cell parameters (Jsc, Voc, FF, and η ] (b) Effect of the HTL doping density on the parameters of solar cell (Jsc, Voc, FF, and η ).
Micromachines 13 02201 g006
Figure 7. (a) Current vs. voltage characteristics and (b) QE vs. wavelength of the proposed CH3NH3SnI3 based perovskites solar cell.
Figure 7. (a) Current vs. voltage characteristics and (b) QE vs. wavelength of the proposed CH3NH3SnI3 based perovskites solar cell.
Micromachines 13 02201 g007
Table 1. Simulation parameters of each layer of the proposed device structure.
Table 1. Simulation parameters of each layer of the proposed device structure.
ParametersP3HT
[28]
CH3NH3SnI3 [18]WS2
[38]
ITO
Thickness (nm)350350150100
Eg (eV)1.7001.31.8003.500
X (eV)3.5004.173.9504.000
εr3.0008.213.6009.000
Nc (1/cm3)2.0 × 10181 × 10182.2 × 10172.2 × 1018
Nv (1/cm3)2.0 × 10191 × 10182.2 × 10161.8 × 1018
Ve (cm/s)1.0 × 1071.0 × 1071.0 × 1071.0 × 107
Vh (cm/s)1.0 × 1071.0 × 1071.0 × 1071.0 × 107
µe (cm2/Vs)1.8 × 10−31.61.0 × 1022.0
µh (cm2/Vs)1.8 × 10−21.61.0 × 1021.0
ND (1/cm3)-1.0 × 10171.0 × 10182.0 × 1019
NA (1/cm3)1.0 × 10191.0 × 1017--
Table 2. Comparison with the finding of the simulated and experimentally reported results.
Table 2. Comparison with the finding of the simulated and experimentally reported results.
Device StructurePCE
(%)
FF
(%)
Voc
(Volt)
Jsc (mA/cm2)References
Zn0.75Mg0.25O/CH3NH3SnI3/MASnBr3
(simulated)
26.3382.010.9533.85[18]
ITO/PEDOT: PSS/CH3NH3SnI3/C60/
BCP/Ag (experimental)
17.1%76.411.0022.95[51]
FTO/TiO2/Perovskite/
SpiroMeOTAD/Ag(experimental)
15.164.20.9923.71[52]
ITO/PEDOT: PSS/CH3NH3PbI3−xClx/C60/
BCP/Ag (experimental)
21.180.31.0922.3[53]
FTO/TiO2/FASnI3/Spiro-OMeTAD/Au (simulated)19.0833.721.1831.20[54]
FTO/ TiO2-ZnS/Spiro-OMeTAD/Au14.90%74.431.0219.07[46]
WS2/CH3NH3SnI3/P3HT
(simulated)
33.4681.591.099737.17This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Umar, A.; Sadanand; Singh, P.K.; Dwivedi, D.K.; Algadi, H.; Ibrahim, A.A.; Alhammai, M.A.M.; Baskoutas, S. High Power-Conversion Efficiency of Lead-Free Perovskite Solar Cells: A Theoretical Investigation. Micromachines 2022, 13, 2201. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13122201

AMA Style

Umar A, Sadanand, Singh PK, Dwivedi DK, Algadi H, Ibrahim AA, Alhammai MAM, Baskoutas S. High Power-Conversion Efficiency of Lead-Free Perovskite Solar Cells: A Theoretical Investigation. Micromachines. 2022; 13(12):2201. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13122201

Chicago/Turabian Style

Umar, Ahmad, Sadanand, Pravin Kumar Singh, D. K. Dwivedi, Hassan Algadi, Ahmed A. Ibrahim, Mohsen A. M. Alhammai, and Sotirios Baskoutas. 2022. "High Power-Conversion Efficiency of Lead-Free Perovskite Solar Cells: A Theoretical Investigation" Micromachines 13, no. 12: 2201. https://0-doi-org.brum.beds.ac.uk/10.3390/mi13122201

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