Next Article in Journal
Laser-Based Manufacturing of Ceramics: A Review
Previous Article in Journal
Design of a Low-Frequency Dielectrophoresis-Based Arc Microfluidic Chip for Multigroup Cell Sorting
Previous Article in Special Issue
High-Efficiency CsPbBr3 Light-Emitting Diodes using One-Step Spin-Coating In Situ Dynamic Thermal Crystallization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comprehensive Study of CsSnI3-Based Perovskite Solar Cells with Different Hole Transporting Layers and Back Contacts

by
Seyedeh Mozhgan Seyed-Talebi
1,2,*,
Mehrnaz Mahmoudi
2 and
Chih-Hao Lee
1,*
1
Department of Engineering and System Science, National Tsing Hua University, Hsinchu 300044, Taiwan
2
Department of Chemistry, Faculty of Science, Shahid Rajaee Teacher Training University, Tehran 1678815811, Iran
*
Authors to whom correspondence should be addressed.
Micromachines 2023, 14(8), 1562; https://doi.org/10.3390/mi14081562
Submission received: 18 July 2023 / Revised: 2 August 2023 / Accepted: 4 August 2023 / Published: 6 August 2023
(This article belongs to the Special Issue Perovskite Materials and Devices: Past, Present and Future)

Abstract

:
By an abrupt rise in the power conservation efficiency (PCE) of perovskite solar cells (PSCs) within a short span of time, the instability and toxicity of lead were raised as major hurdles in the path toward their commercialization. The usage of an inorganic lead-free CsSnI3-based halide perovskite offers the advantages of enhancing the stability and degradation resistance of devices, reducing the cost of devices, and minimizing the recombination of generated carriers. The simulated standard device using a 1D simulator like solar cell capacitance simulator (SCAPS) with Spiro-OMeTAD hole transporting layer (HTL) at perovskite thickness of 330 nm is in good agreement with the previous experimental result (12.96%). By changing the perovskite thickness and work operating temperature, the maximum efficiency of 18.15% is calculated for standard devices at a perovskite thickness of 800 nm. Then, the effects of replacement of Spiro-OMeTAD with other HTLs including Cu2O, CuI, CuSCN, CuSbS2, Cu2ZnSnSe4, CBTS, CuO, MoS2, MoOx, MoO3, PTAA, P3HT, and PEDOT:PSS on photovoltaic characteristics were calculated. The device with Cu2ZnSnSe4 hole transport in the same condition shows the highest efficiency of 21.63%. The back contact also changed by considering different metals such as Ag, Cu, Fe, C, Au, W, Ni, Pd, Pt, and Se. The outcomes provide valuable insights into the efficiency improvement of CsSnI3-based PSCs by Spiro-OMeTAD substitution with other HTLs, and back-contact modification upon the comprehensive analysis of 120 devices with different configurations.

1. Introduction

Recently, perovskite semiconductors as absorber layers in perovskite solar cell (PSC) devices have drawn wide interest owing to their exceptional optical and electrical properties [1,2,3,4,5,6]. Within a few years, the power conservation efficiency (PCE) of perovskite devices boosted rapidly up to the PCE of 25.8% in 2020 for single-junction perovskite devices, which is comparable with the silicon-based classical photovoltaic devices (PCE of 26.7%) [7]. Further experimental findings showed the maximum PCE of 26.1% in 2021 [8,9], and multi-junction perovskite/silicon tandem solar cells achieved a certified PCE of 32.5% [10]. The cesium tin iodide (CsSnI3) is an eco-friendly inorganic solution for energy harvesting solar cell technology [11,12,13,14,15,16,17]. Because of the desirable band gap (Eg ≈ 1.3 eV) to achieve the ideal Shockley–Queisser theoretical PCE value (highest efficiency of 33.7% at 1.34 eV) than the band gap of lead-based perovskite absorbers (1.5 eV to 2.3 eV), strong visible light absorption coefficient (≈104 cm−1), effective local hole mobility of ∼400 cm2/(V·s) for mitigating the recombination rate, small binding energies of Wannier excitons (~10–20 meV), and easy charge separation, the CsSnI3 is a more conducive lead-free inorganic perovskite material to achieve the excellent device performance than lead-based hybrid perovskite materials [18,19,20,21].
In 2012, the first reported CsSnI3-based PSC measured the efficiency of 0.9% [22]. In 2014, an improved PCE of 2.02% was reported experimentally [23]. The facile oxidation of Sn2+ to Sn4+, high self-doping effect, and Sn vacancy result in fast degradation of the CsSnI3 absorber and instability of CsSnI3-based PSCs. Inspiringly, several effective approaches have been suggested to enhance the stability, and quality of CsSnI3 thin films, through additive engineering, composition engineering, fabricating the perovskite quantum dots, developing a two-step sequential deposition method [24,25,26,27,28]. An inverted CsSnI3-based PSC with efficiency >10%, was experimentally reported in previous literature [29]. These reported PCEs are much lower than the best reported results for lead-based PSCs and the theoretical ideal Shockley–Queisser limit.
Recently, an increasing number of theoretical studies have been carried out to design more efficient inorganic lead-free PSC devices with different electron transporting layer (ETL) and hole transporting layer (HTL) in both normal and inverted structures [30,31,32,33,34,35]. The P-type semiconductors usually regarded as HTLs for extracting the photogenerated holes from the perovskite absorber layer and blocking the carrier recombination. Conventionally, inorganic/organic HTLs have been categorized depending on their chemical structure and contents. Due to great benefits such as the easy and low cost preparation process, high chemical stability, and good hole mobility, inorganic HTLs attracted attention as potential candidates for usage in stable PSCs. Although, organic HTMs usually exhibited higher PCE than inorganic HTMs, their low conductivity and/or hole mobility induce instability of organic HTM-based PSC devices. Therefore, adding dopants is necessary for most of the effective organic HTMs to improve their conductivity and/or hole mobility. Herein, several typical inorganic HTLs such as CuI, CuSCN, Cu2O, CuO, MoS2 and Cu2ZnSnSe4 (CZTSe) studied in comparison to organic HTLs such as Spiro-MeOTAD, PEDOT:PSS, P3HT, and PTAA to provide some enlightenment for design a more efficient nontoxic lead-free CsSnI3-based PSC that can use for further promote PSCs performance, scalable manufacturing, and commercial application using low cost and easy preparation HTMs. Among them, Cu2ZnSnSe4 (CZTSe) showed the highest efficiency. Cu2ZnSnSe4 is a p-type semiconductor with high hole mobility (15.1 cm2v−1s−1), low resistivity (0.33 Ω-cm), earth-abundant elemental constituents, and non-toxic properties, which was regarded as a promising inorganic HTL in solar cells [36,37]. The simulated standard device in the present study confirmed the photovoltaic characteristic parameters of the experimentally reported results of the CsSnI3-based device with the highest PCE of 12.96%. By optimizing the perovskite absorber thickness, the resultant data proposed a greater thickness for improving the PCE of CsSnI3-based device in experiments. According to best of our search, the previous theoretical and experimental studies did not mention this issue. The HTL modification study in this paper is accomplished by considering an optimized thickness for a perovskite layer that was extracted from the agreement between experiment and simulation results. Herein, we focused on the optimization of a CsSnI3-based perovskite solar cell with Spiro-OMeTAD HTL as the standard cell. The final generation-recombination rate, current density–voltage and quantum efficiency of optimized simulated standard device with and without Spiro-OMeTAD HTL compared with each other. Then, the effect of Spiro-OMeTAD replacement with other HTLs is investigated.

2. Theoretical Methods and Device Structures

Theoretical simulation is a valuable tool for designing and characterizing devices under different conditions. Moreover, simulation reduces the cost and time associated with their experimental investigation by optimizing the necessary parameters and material properties before doing experiments.

2.1. Computational Simulation Details

In the present study, the solar cell capacitance simulator structures (SCAPS-1D) software version 3.3.08 is utilized to find the photovoltaic characteristics of short-circuit current density (Jsc), open-circuit voltage (Voc), fill factor (FF), and power conversion efficiency (PCE) by resolving coupled one-dimensional Poisson’s, and continuity semiconductor equations. SCAPS-1D is most popular software among photovoltaic researchers as its simulation results demonstrated a close match with the experimental results [38]. A solar cell device in SCAPS is defined as number of semiconductor layers, in which the transportation of electrons and holes consider by solving the carrier continuity equations and Poisson’s equations iteratively for both electrons and holes. The Poisson’s equation (Equation (1)) relates the charges to the electrostatic potential by following equation:
d 2 Ψ d x 2 = q ε 0 ε r [ p x n x + N D + x N A + p t x n t x ]
Here, Ψ represents the electrostatic potential, q is the electron charge, ε0 represents the permittivity of free space, εr is the relative permittivity of the semiconductor material, p(n) is the hole (electron) concentration, NA (ND+) is the density of the ionized acceptors (donors), nt (pt) is the trapped electron (hole), and x is the position coordinate. Because of the simultaneous investigation of recombination, generation, drift, and diffusion, the continuity equations are used. The continuity equations for the change in the concentration of electrons (Equation (2)) and holes (Equation (3)) are given as [39]:
d n p d t = G n n p n p 0 τ n + n p μ n d E d x + μ n E d n p d x + D n d 2 n p d x 2
d p n d t = G p p n p n 0 τ p + p n μ p d E d x + μ p E d p n d x + D p d 2 p n d x 2
where Gn (Gp) is the electron (hole) generation rate, np (pn) is the electron (hole) concentration in the p-region (n-region), np0 (pn0) is the equilibrium electron (hole) concentration in the p-region (n-region), τn and τp denote electron and hole lifetime, μp (μn) is the hole (electron) mobility, E is the electric field, and Dn (Dp) is the electron (hole) diffuse on coefficient.
The photovoltaic parameters such as PCE, Voc, Jsc, and FF are related to each other by the following equations [38]:
F F = P m a x P i n = I m a x × V m a x V o c × I s c
P C E = F F × V o c × I s c P i n
where Pmax is the maximum power achievable from the point on the current–voltage (I–V) curve of a solar module under illumination that the product of current and voltage is maximum.

2.2. Device Structure

In most of the perovskite device configurations, the perovskite absorber layer is sandwiched between an electron transport layer (ETL), and a hole transport layer (HTL). According to the semiconductor physics, ETL and HTL are responsible for preventing photogenerated charge carriers from recombination and boosting charge transportation by creating the carrier separation paths to deliver electrons and holes from perovskite absorber to the electrodes and go into the external circuit through the metal (Au) contacts.
The simulated CsSnI3-based perovskite solar cell in the present study has a normal (n-i-p) structure with TiO2 ETL. The exact values of the necessary input parameters for simulation using SCAPS-1D, are difficult to obtain by simulation. The input parameters extracted from literatures are listed in Table 1 [34,40,41,42]. The schematic illustration of device configuration and corresponding energy levels of different layers in simulated device and the band alignment of different HTLs with a perovskite absorber layer are shown in Figure 1.

3. Results and Discussions

3.1. Photovoltaic Characterization of Standard Device

The photocurrent density–voltage (J–V) and quantum efficiency measurements of the simulated standard PSC device with and without Spiro-OMeTAD recorded using SCAPS-1D software are presented in Figure 2. The fill-factor (FF), short circuit current (Jsc), open circuit voltage (Voc), and power conversion efficiency (PCE) were deduced from the J-V characteristics. A standard simulated AM 1.5 G irradiation system (1000 W.m2) and a bias potential scan from 0.0 V to +1.0 V was applied with a scan speed of 100 mV. s1 for characterization of PSCs. According to the previous literature, the main responsibility of HTL is effecting on Voc [41]. Therefore, a disparity appears between series and recombination resistances (Rs and Rrec, respectively) in simulated standard devices with and without HTL. Effect of perovskite thickness modification on photovoltaic parameters of standard optimized device presented in Figure 2. The highest efficiency of 18.06% was obtained for a CsSnI3 perovskite layer with a thickness of 800 nm.
By increasing the absorber layer thickness, more amount of generated carriers push to move from the traps and the charge separation ability of perovskite absorber increases. Therefore, the generated current density increases because of the higher rate of producing the photogenerated carriers [33,43,44]. Although, the generation per volume unit decreases (Figure 3a). As shown in Figure 3b, higher increase in the absorber layer thickness results in creating the additional traveling paths for carriers to transfer through the absorber. Therefore, the recombination rate and series resistance decrease by creating a bigger barrier between HTL and ETL. Thus, a thicker perovskite layer produces a weaker built-in electric field, which leads to a reduction of Voc and FF by decreasing the recombination and cell series resistance [45].
By increasing absorber layer thickness, the current density of electron and holes increase (Figure 3c). Although, the total current density in Figure 2b shows a saturated amount for devices contain a perovskite absorber thicker than 800 nm. The results plotted in Figure 3d–f demonstrate energy levels considered fixed and the thickness is the only parameter that is modified in this section.

3.2. Photovoltaic Characterization of Standard Device vs. Experimental Results and HTL Free Device

The simulated CsSnI3-based standard device with/without HTL compared in Figure 4a. The contact resistance between the metal electrode and perovskite layer causes increasing the series resistance. Therefore, the fill factor and current density of the HTL-free device decrease in comparison to the standard device with spiro-OMeTAD HTL. The HTL acts as a separator layer between perovskite absorber and metal back electrode to enhance the stability and efficiency of perovskite device (Figure 4a). The existence of HTL leads to increasing the amount of collected holes from photo-generated carriers by incident photons at a given energy (QE), and photon absorption at longer wavelengths (Figure 4b). The theoretical characterization results of standard PSC device simulated by SCAPS software at perovskite thickness of 330 nm are in good agreement with the previous experimental report of 12.96% (Figure 4c) [46].
The simulation results in Figure 4c and data listed in Table 2 demonstrate the considered thickness in experimentally fabricated devices is important issue for more efficient CsSnI3-based PSC devices. According to the current calculation results, increasing the perovskite thickness from 330 nm to 800 nm improves the efficiency of fabricated devices from 12.96% to 18.11% at room temperature. The thickness of fabricated HTL in devices that can assess HTLs between 300 nm to 800 nm is important for improving efficiency.
The operating temperature for all of devices was considered fixed at 300 K in the present study. The effect of work operating temperature modification on the optimized standard device parameters is investigated. According to the data shown in Figure 5, the highest efficiency of 18.15% is calculated at 320 K for the standard device. By increasing the temperature, the efficiency will decrease rapidly. Because of the increase in the number of charge carriers reaching to the electrodes, a continuous increase in the saturation current is observable. Increasing the reverse saturation current, due to higher surface recombination, causes the Voc to decline with an increase in temperature, while FF increases up initially, then abruptly starts to decline. The main reason for the fill factor falling is due to the increased resistance of the device.

3.3. Photovoltaic Characterization of Simulated Device with HTL Modification

In the simulated CsSnI3-based perovskite devices with normal structure, the existence of a hole transport layer helps the photogenerated holes to transfer easily from the absorber layer to metal back contact and also prevent holes recombination with electrons by rapid reflecting them to the circuit through metal back contact. It is noticeable that the input parameters of Spiro-OMeTAD in SCAPS software extract from experiments data. In experiment, the Spiro-OMeTAD is doped with Li-TFSI and tBP dopants for increasing the favorite properties of HTL. Therefore, Spiro-OMeTAD is very expensive. Herein, some proposed p-type semiconductors as HTL in the literature include CuI, CuSbS2, CBTS, MoS2, MnO2, P3HT, Cu2O, CuSCN, CZTSe, CuO, MoOx, PTTA, and PEDOT:PSS investigated to find an appropriate replacement for Spiro-OMeTAD. The resultant J–V curves and external quantum efficiency (QE) for simulated devices with different HTLs and configurations of FTO/TiO2/CsSnI3/HTL/Au plotted in Figure 6.
Among all of the devices with different HTLs, simulated devices with Cu2O, CuSCN, Cu2ZnSnSe4 (CZTSe), CBTS, CuO, MoS2, and MoOx showed better efficiency than Spiro-OMeTAD and the device with Cu2ZnSnSe4 (CZTSe) as an HTL layer showed the highest efficiency of 21.63% (Table 2) while the minimum PCE achieved for a device with PEDOT:PSS HTL (Figure 6a). The favorable band alignment of Cu2O offers a higher current density while lower current density obtained from a device with MoS2 HTL owing to unsatisfactory band alignment with perovskite absorber and lower hole transportation to metal electrodes. Therefore, the HTL’s band alignment influences the flow of photogenerated holes in CsSnI3-based PSCs. The corresponding QE as a function of wavelength (λ) is studied in the range of 400 nm–1000 nm as shown in Figure 6b. In this study, the ETL, light absorption, and incident photons to the perovskite layer were fixed. The QE started to increase from 400 nm and dropped after 820 nm, corresponding to the band edge of each active material. The maximum QE was obtained for the device with a CZTSe HTL as expected from J–V characteristics and the minimum for the device with a MoS2 HTL.

3.4. Photovoltaic Characterization of Simulated Device with HTL by Modification of Metal Back Contact

As discussed in the previous section, in the normal structures the photogenerated holes extracted through the metal back contact to the external circuit. Herein, ten different metals of Ag, Cu, Fe, C, Au, W, Ni, Pd, Pt, and Se are used as back contact in simulated devices to investigate the best combination of each HTL for having more efficient devices values among 120 different configurations. The resultant photovoltaic characteristic data are listed in Figure 7. It is noticeable to experimentalists that the fabricated devices with Se, Pt, and Pd back contacts revealed higher efficiency than commonly used Au electrodes for most of the devices with different HTLs. The most characteristic parameter essentially affected by the metal back contact modification in simulated PSCs is the fill factor, due to modification in surface recombination and series resistance created by connection between metal back contact and HTLs.

4. Conclusions

The present work reported a comparison between different CsSnI3-based PSC devices with modified HTLs and considering different metal back contacts in a normal structure. The simulated devices were characterized by SCAPS software. The standard device with Spiro-OMeTAD HTL showed good overlap with experiment results in 330 nm thickness. The effect of perovskite thickness and working temperature on photovoltaic parameters of devices investigated. Spiro-OMeTAD replaced with ten different HTLs. The device with Cu2ZnSnSe4 (CZTSe) as an HTL layer showed the highest efficiency of 21.63%. The metal back contact also changed by modifying the work function of the back contact among 120 configurations. Thus, several promising and competitive configurations for high efficient CsSnI3-based PSCs are proposed.

Author Contributions

Conceptualization, S.M.S.-T. and C.-H.L.; methodology, S.M.S.-T. and M.M.; formal analysis and investigation, S.M.S.-T. and M.M.; writing—original draft, S.M.S.-T.; project administration C.-H.L.; supervision, C.-H.L.; validation, S.M.S.-T., M.M. and C.-H.L.; resources C.-H.L.; writing—review and editing, S.M.S.-T. and C.-H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research funded by National Science and Technology Council (NSTC), Taiwan (grant No. NSTC112-2811-M-007-010), and Iran National Science Foundation (INSF), Iran (Proposal No. 99014923).

Data Availability Statement

All data are included in the manuscript.

Acknowledgments

The authors would like to acknowledge Marc Burgelman, the Ghent University, in Belgium, for providing the SCAPS-1D software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tu, Y.; Wu, J.; Xu, G.; Yang, X.; Cai, R.; Gong, Q.; Zhu, R.; Huang, W. Perovskite Solar Cells for Space Applications: Progress and Challenges. Adv. Mater. 2021, 33, 2006545. [Google Scholar] [CrossRef]
  2. Basumatary, P.; Agarwal, P. A short review on progress in perovskite solar cells. Mater. Res. Bull. 2022, 149, 111700. [Google Scholar]
  3. Deng, K.; Li, L. Optical Design in Perovskite Solar Cells. Small Methods 2020, 4, 1900150. [Google Scholar]
  4. Murugadoss, G.; Thangamuthu, R. Metals doped cesium based all inorganic perovskite solar cells: Investigations on Structural, morphological and optical properties. Sol. Energy 2019, 179, 151–163. [Google Scholar] [CrossRef]
  5. Seyed-Talebi, S.M.; Kazeminezhad, I.; Shahbazi, S.; Diau, G. Efficiency and Stability Enhancement of Fully Ambient Air Processed Perovskite Solar Cells Using TiO2 Paste with Tunable Pore Structure. Adv. Mater. Interfaces 2020, 7, 1900939. [Google Scholar] [CrossRef]
  6. Seyed-Talebi, S.M.; Kazeminezhad, I. Performance improvement of fully ambient air fabricated perovskite solar cells in an anti-solvent process using TiO2 hollow spheres. J. Colloid Interface Sci. 2020, 562, 125–132. [Google Scholar] [PubMed]
  7. Min, H.; Lee, D.Y.; Kim, J.; Kim, G.; Lee, K.S.; Kim, J.; Paik, M.J.; Kim, Y.K.; Kim, K.S.; Kim, M.G.; et al. Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes. Nature 2021, 598, 444–450. [Google Scholar] [PubMed]
  8. Best Research-Cell Efficiency Chart. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 17 July 2023).
  9. Saikia, D.; Betal, A.; Bera, J.; Sahu, S. Progress and challenges of halide perovskite-based solar cell—A brief review. Mater. Sci. Semicond. Process. 2022, 150, 106953. [Google Scholar]
  10. Iqbal, S.; Riaz, K.; Imran, H.; Khattak, Y.H.; Baig, F.; Ahmad, Z. Computational modelling of monolithically stacked perovskite/ silicon tandem solar cells using monofacial and bifacial designs. Optik 2020, 206, 163427. [Google Scholar]
  11. Wu, T.; Liu, X.; Luo, X.; Lin, X.; Cui, D.; Wang, Y.; Segawa, H.; Zhang, Y.; Han, L. Lead-free tin perovskite solar cells. Joule 2021, 5, 863–886. [Google Scholar]
  12. Nasti, G.; Abate, A. Tin Halide Perovskite (ASnX3) Solar Cells: A Comprehensive Guide toward the Highest Power Conversion Efficiency. Adv. Energy Mater. 2020, 10, 1902467. [Google Scholar] [CrossRef]
  13. Yu, B.B.; Chen, Z.; Zhu, Y.; Wang, Y.; Han, B.; Chen, G.; Zhang, X.; Du, Z.; He, Z. Heterogeneous 2D/3D tin-halides perovskite solar cells with certified conversion efficiency breaking 14%. Adv. Mater. 2021, 33, 2102055. [Google Scholar]
  14. Dimesso, L.; Das, C.; Stöhr, M.; Mayer, T.; Jaegermann, W. Properties of cesium tin iodide (Cs-Sn-I) systems after annealing under different atmospheres. Mater. Chem. Phys. 2017, 197, 27–35. [Google Scholar] [CrossRef]
  15. Chung, I.; Song, J.H.; Im, J.; Androulakis, J.; Malliakas, C.D.; Li, H.; Freeman, A.J.; Kenney, J.T.; Kanatzidis, M.G. CsSnI3: Semiconductor or Metal? High Electrical Conductivity and Strong Near-Infrared Photoluminescence from a Single Material. High Hole Mobility and Phase-Transitions. J. Am. Chem. Soc. 2012, 134, 8579–8587. [Google Scholar] [CrossRef]
  16. Zhang, M.; Chen, K.; Wei, Y.; Hu, W.; Cai, Z.; Zhu, J.; Ye, Q.; Ye, F.; Fang, Z.; Yang, L.; et al. Strategies for Optimizing the Morphology of CsSnI3 Perovskite Solar Cells. Crystals 2023, 13, 410. [Google Scholar] [CrossRef]
  17. Luna, P.S.; Pokharel, U.; Huisman, B.A.H.; Koster, L.J.A.; Palazon, F.; Bolink, H.J. Vacuum-Deposited Cesium Tin Iodide Thin Films with Tunable Thermoelectric Properties. ACS Appl. Energy Mater. 2022, 5, 10216–10223. [Google Scholar] [CrossRef]
  18. Shockley, W.; Queisser, H.J. Detailed Balance Limit of Efficiency of P-N Junction Solar Cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  19. Kamarudin, M.A.; Hirotani, D.; Wang, Z.; Hamada, K.; Nishimura, K.; Shen, Q.; Toyoda, T.; Iikubo, S.; Minemoto, T.; Yoshino, K.; et al. Suppression of Charge Carrier Recombination in Lead-Free Tin Halide Perovskite via Lewis Base Post-treatment. J. Phys. Chem. Lett. 2019, 10, 5277–5283. [Google Scholar] [CrossRef]
  20. Mitzi, D.B.; Feild, C.A.; Schlesinger, Z.; Laibowitz, R.B. Transport, Optical, and Magnetic Properties of the Conducting Halide Perovskite CH3NH3SnI3. J. Solid State Chem. 1995, 114, 159–163. [Google Scholar] [CrossRef]
  21. Aktas, E.; Rajamanickam, N.; Pascual, J.; Hu, S.; Aldamasy, M.H.; Girolamo, D.D.; Li, W.; Nasti, G.; Ferrero, E.M.; Wakamiya, A.; et al. Challenges and strategies toward long-term stability of lead-free tin-based perovskite solar cells. Commun. Mater. 2022, 3, 104. [Google Scholar]
  22. Chen, Z.; Wang, J.J.; Ren, Y.; Yu, C.; Shum, K. Schottky solar cells based on CsSnI3 thin-films. Appl. Phys. Lett. 2012, 101, 093901. [Google Scholar]
  23. Kumar, M.H.; Dharani, S.; Leong, W.; Boix, P.P.; Baikie, T.; Ding, H.; Graetzel, M.; Mhaisalkar, S.G.; Mathews, N. Lead-Free Halide Perovskite Solar Cells with High Photocurrents Realized Through Vacancy Modulation. Adv. Mater. 2014, 26, 7122–7127. [Google Scholar] [PubMed]
  24. Wang, G.; Chang, J.; Bi, J.; Lei, M.; Wang, C.; Qiao, Q. Inorganic CsSnI3 Perovskite Solar Cells: The Progressand Future Prospects. Sol. RRL 2022, 6, 2100841. [Google Scholar]
  25. Qiu, X.; Cao, B.; Yuan, S.; Chen, X.; Qiu, Z.; Jiang, Y.; Ye, Q.; Wang, H.; Zeng, H.; Liu, J.; et al. From unstable CsSnI3 to air-stable Cs2SnI6: A lead-free perovskite solar cell light absorber with bandgap of 1.48 eV and high absorption coefficient. Sol. Energy Mater. Sol. Cells 2017, 159, 227–234. [Google Scholar]
  26. Ye, T.; Wang, K.; Hou, Y.; Yang, D.; Smith, N.; Magill, B.; Yoon, J.; Mudiyanselage, R.R.H.H.; Khodaparast, G.A.; Wang, K.; et al. Ambient-Air-Stable Lead-Free CsSnI3 Solar Cells with Greater than 7.5% Efficiency. J. Am. Chem. Soc. 2021, 143, 4319–4328. [Google Scholar]
  27. Lin, S.; Zhang, B.; Lü, T.Y.; Zheng, C.; Pan, H.; Chen, H.; Lin, C.; Li, X.; Zhou, J. Inorganic Lead-Free B-γ-CsSnI3 Perovskite Solar Cells Using Diverse Electron-Transporting Materials: A Simulation Study. ACS Omega 2021, 6, 26689–26698. [Google Scholar]
  28. Abate, A. Stable Tin-Based Perovskite Solar Cells. ACS Energy Lett. 2023, 8, 1896–1899. [Google Scholar]
  29. Ye, T.; Wang, X.; Wang, K.; Ma, S.; Yang, D.; Hou, Y.; Yoon, J.; Wang, K.; Priya, S. Localized Electron Density Engineering for Stabilized B-γCsSnI3-Based Perovskite Solar Cells with Efficiencies >10%. ACS Energy Lett. 2021, 6, 1480–1489. [Google Scholar]
  30. Sarkar, D.K.; Hasan, A.K.M.; Mottakin, M.; Selvanathan, V.; Sobayel, K.; Islam, M.A.; Muhammad, G.; Aminuzzaman, M.; Shahiduzzaman, M.; Sopian, K.; et al. Lead free efficient perovskite solar cell device Optimization and defect study using Mg doped CuCrO2 as HTL and WO3 as ETL. Sol. Energy 2022, 243, 215–224. [Google Scholar]
  31. Xie, L.; Zhang, L.; Hua, Y.; Ding, L. Inorganic electron-transport materials in perovskite solar cells. J. Semicond 2022, 43, 040201. [Google Scholar]
  32. Seyed-Talebi, S.M.; Beheshtian, J. Lead-Free Inorganic Cesium Tin-Germanium Triiodide Perovskites for Photovoltaic Application. Int. J. Energy Power Eng. 2021, 15, 252–257. [Google Scholar]
  33. Imani, S.; Seyed-Talebi, S.M.; Beheshtian, J.; Diau, E.W.G. Simulation and characterization of CH3NH3SnI3-based perovskite solar cells with different Cu-based hole transporting layers. Appl. Phys. A 2023, 129, 143. [Google Scholar] [CrossRef]
  34. Hossain, M.K.; Toki, G.F.I.; Kuddus, A.; Rubel, M.H.K.; Hossain, M.M.; Bencherif, H.; Rahman, M.F.; Islam, M.R.; Mushtaq, M. An extensive study on multiple ETL and HTL layers to design and simulation of high-performance lead-free CsSnCl3-based perovskite solar cells. Sci. Rep. 2023, 13, 2521. [Google Scholar] [CrossRef] [PubMed]
  35. Lin, L.; Jones, T.W.; Yang, C.J.; Duffy, N.W.; Li, J.; Zhao, L.; Chi, B.; Wang, X.; Wilson, G.J. Inorganic Electron Transport Materials in Perovskite Solar Cells. Adv. Funct. Mater. 2021, 31, 2008300. [Google Scholar]
  36. Niemegeers, A.; Burgelman, M.; Degrave, S.; Verschraegen, J.; Decock, K. SCAPS (Version: 3.3. 08) Manual November 7, 2018; University of Gent: Gent, Belgium, 2020. Available online: https://scaps.elis.ugent.be (accessed on 17 July 2023).
  37. Ashebir, G.Y.; Dong, C.; Wan, Z.; Qi, J.; Chen, J.; Zhao, Q.; Chen, W.; Wang, M. Solution-processed Cu2ZnSnSe4 nanoparticle film as efficient hole transporting layer for stable perovskite solar cells. J. Phys. Chem. Solids 2019, 129, 204–208. [Google Scholar]
  38. Li, J.; Huang, J.; Ma, F.; Sun, H.; Cong, J.; Privat, K.; Webster, R.F.; Cheong, S.; Yao, Y.; Chin, R.L.; et al. Unveiling microscopic carrier loss mechanisms in 12% efficient Cu2ZnSnSe4 solar cells. Nat. Energy 2022, 7, 754–764. [Google Scholar]
  39. Ravidas, B.K.; Roy, M.K.; Samajdar, D.P. Investigation of photovoltaic performance of lead-free CsSnI3-based perovskite solar cell with different hole transport layers: First Principle Calculations and SCAPS-1D Analysis. Sol. Energy 2023, 249, 163–173. [Google Scholar] [CrossRef]
  40. Jiang, M.; Tang, J. Simulated development and optimized performance of narrow-bandgap CsSnI3-based all-inorganic perovskite solar cells. J. Phys. D Appl. Phys. 2021, 54, 465104. [Google Scholar] [CrossRef]
  41. Pindolia, G.; Shinde, S.M.; Jha, P.K. Optimization of an inorganic lead free RbGeI3 based perovskite solar cell by SCAPS-1D simulation. Sol. Energy 2022, 236, 802–821. [Google Scholar]
  42. Srivastava, S.; Singh, A.K.; Kumar, P.; Pradhan, B. Comparative performance analysis of lead-free perovskites solar cells by numerical simulation. J. Appl. Phys. 2022, 131, 175001. [Google Scholar]
  43. Juarez-Perez, E.J.; Wuβler, M.; Fabregat-Santiago, F.; Lakus, K.; Wollny, K.; Mankel, E.; Mayer, T.; Jaegermann, W.; Mora-Sero, I. Role of the Selective Contacts in the Performance of Lead Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 680–685. [Google Scholar] [CrossRef] [PubMed]
  44. Madan, J.; Singh, K.; Pandey, R. Comprehensive device simulation of 23.36% efficient two-terminal perovskite-PbS CQD tandem solar cell for low-cost applications. Sci. Rep. 2021, 11, 19829. [Google Scholar] [CrossRef] [PubMed]
  45. Ni, Z.; Bao, C.; Liu, Y.; Jiang, Q.; Wu, Q.; Chen, S.; Dai, X.; Chen, B.; Hartweg, B.; Yu, Z.; et al. Resolving spatial and energetic distributions of trap states in metal halide perovskite solar cells. Science 2020, 397, 1352–1358. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, L.-J.; Lee, C.-R.; Chuang, Y.-J.; Wu, Z.-H.; Chen, C. Synthesis and Optical Properties of Lead-Free Cesium Tin Halide Perovskite Quantum Rods with High-Performance Solar Cell Application. J. Phys. Chem. Lett. 2016, 7, 5028–5035. [Google Scholar]
Figure 1. Schematic presentation of (a) device configuration, and (b) corresponding energy band levels of different layers of simulated inverted CsSnI3-based PSCs.
Figure 1. Schematic presentation of (a) device configuration, and (b) corresponding energy band levels of different layers of simulated inverted CsSnI3-based PSCs.
Micromachines 14 01562 g001
Figure 2. The effect of perovskite thickness modification on photovoltaic parameters of (a) current density, (b) open circuit voltage, (c) fill factor, and (d) efficiency of simulated standard devices with Spiro-OMeTAD HTL.
Figure 2. The effect of perovskite thickness modification on photovoltaic parameters of (a) current density, (b) open circuit voltage, (c) fill factor, and (d) efficiency of simulated standard devices with Spiro-OMeTAD HTL.
Micromachines 14 01562 g002
Figure 3. The effect of perovskite thickness modification on (a) generation rate, (b) recombination rate of photo-generated carriers per volume unit in different layers of simulated devices with different perovskite absorber thickness, (c) current density of electrons (Jn) and holes (Jp), and conduction band position (Ec), valence band position (Ev), Fermi energy levels for device with perovskite thickness of (d) 100 nm, (e) 330 nm, and (f) 800 nm.
Figure 3. The effect of perovskite thickness modification on (a) generation rate, (b) recombination rate of photo-generated carriers per volume unit in different layers of simulated devices with different perovskite absorber thickness, (c) current density of electrons (Jn) and holes (Jp), and conduction band position (Ec), valence band position (Ev), Fermi energy levels for device with perovskite thickness of (d) 100 nm, (e) 330 nm, and (f) 800 nm.
Micromachines 14 01562 g003
Figure 4. (a) The current density–voltage (J-V) curves, and (b) quantum efficiency curves of devices with/without Spiro-OMeTAD as the HTL. (c) Comparison between J–V curves of simulated device and experimental results.
Figure 4. (a) The current density–voltage (J-V) curves, and (b) quantum efficiency curves of devices with/without Spiro-OMeTAD as the HTL. (c) Comparison between J–V curves of simulated device and experimental results.
Micromachines 14 01562 g004
Figure 5. The effect of work temperature on the photovoltaic parameters (a) open circuit voltage, (b) short circuit current density, (c) fill factor, and (d) efficiency of simulated standard PSC.
Figure 5. The effect of work temperature on the photovoltaic parameters (a) open circuit voltage, (b) short circuit current density, (c) fill factor, and (d) efficiency of simulated standard PSC.
Micromachines 14 01562 g005
Figure 6. (a) The current density–voltage (J–V), and (b) quantum efficiency curves of simulated devices with different HTLs.
Figure 6. (a) The current density–voltage (J–V), and (b) quantum efficiency curves of simulated devices with different HTLs.
Micromachines 14 01562 g006
Figure 7. Optimization of photovoltaic parameters of simulated PSCs with different metal back contacts for different HTLs.
Figure 7. Optimization of photovoltaic parameters of simulated PSCs with different metal back contacts for different HTLs.
Micromachines 14 01562 g007
Table 1. Input parameters of different layers of FTO, TiO2, CsSnI3, HTLs for simulated devices in the present study by SCAPS-1D software.
Table 1. Input parameters of different layers of FTO, TiO2, CsSnI3, HTLs for simulated devices in the present study by SCAPS-1D software.
ParametersNt (1/cm3)ND (cm−3)NA (cm−3)εµp (cm2v−1s−1)µn (cm2v−1s−1)Nv (cm−3)NC (cm−3)χ (eV)Eg (eV)Thickness (µm)
FTO 1.0 × 10+15 1.0 × 10+15 0 9 10 20 1.8 × 10+19 2.0 × 10+18 4.0 3.5 0.500
TiO2 1.0 × 10+17 5.0 × 10+19 0 9 10 20 2.0 × 10+20 1.0 × 10+21 4.0 3.2 0.150
CsSnI3 1.0 × 10+13 0.0 1.8 × 10+16 18 4.37 4.37 1.0 × 10+19 1.0 × 10+19 4.47 1.27 0.803
Spiro-OMeTAD 1.0 × 10+15 0.0 2.0 × 10+18 3 2.0 × 10−4 2.0 × 10−4 1.9 × 10+19 2.2 × 10+18 3 2.45 0.200
CuSCN 1.0 × 10+15 0.0 1.0 × 10+18 10.0 25.0 100 1.8 × 10+18 2.8 × 10+19 1.4 3.8 0.200
Cu2O 1.0 × 10+15 0.0 1.0 × 10+18 7.5 80 200 1.0 × 10+19 2.0 × 10+19 3.4 2.2 0.200
CuI 1.0 × 10+15 0.0 1.0 × 10+18 6.5 43.9 100 1.0 × 10+19 2.8 × 10+19 2.1 3.1 0.200
CuSbS2 1.0 × 10+15 0.0 1.0 × 10+18 14.6 49 49 1.0 × 10+19 2.0 × 10+18 4.2 1.58 0.200
CZTSe 1.0 × 10+15 0.0 5.0 × 10+16 9.1 40 145 1.8 × 10+19 2.2 × 10+18 1.00 4.46 0.200
CBTS 1.0 × 10+15 0.0 1.0 × 10+18 5.4 10 30 1.8 × 10+19 2.2 × 10+18 3.6 1.9 0.200
CuO 1.0 × 10+15 0.0 1.0 × 10+18 18.1 0.1 100 5.5 × 10+20 2.2 × 10+19 4.07 1.51 0.200
MoS2 1.0 × 10+15 0.0 1.0 × 10+17 13.6 150 100 1.8 × 10+19 2.2 × 10+18 4.2 1.29 0.200
MoOX 1.0 × 10+15 0.0 5.0 × 10+17 10 2.5 30 2.5 × 10+17 3.2 × 10+16 2.3 3.2 0.200
MoO3 1.5 × 10+17 0.0 1.0 × 10+18 5.7 1.0 × 10−4 1.0 × 10−4 4.343 × 10+19 4.386 × 10+19 2.3 3 0.200
PTAA 1.0 × 10+15 0.0 1.0 × 10+18 9 40 1 1.0 × 10+21 1.0 × 10+21 2.3 2.96 0.200
P3HT 1.0 × 10+15 0.0 1.0 × 10+18 3 1.86 × 10−2 1.8 × 10−3 2.0 × 10+21 2.0 × 10+21 3.5 1.7 0.200
PEDOT:PSS 1.0 × 10+15 0.0 1.0 × 10+18 3 4.5 × 10−2 4.5 × 10−2 1.8 × 10+19 2.2 × 10+18 3.4 1.6 0.200
Table 2. The photovoltaic parameters of a simulated device with different HTLs.
Table 2. The photovoltaic parameters of a simulated device with different HTLs.
HTLVoc (V)Jsc (mA.cm−2)FFPCE (%)
HTL-free 0.54 32.15 66.58 11.62
Spiro-OMeTAD (exp) 0.86 23.20 65.00 12.96
Spiro-OMeTAD (330 nm) 0.86 23.09 65.02 12.96
Spiro-OMeTAD (800 nm) 0.83 34.34 63.13 18.06
Cu2O 0.83 34.51 65.79 18.92
CuI 0.83 34.31 59.46 17.02
CuSCN 0.83 34.38 72.79 20.87
CuSbS2 0.72 34.47 53.00 13.23
Cu2ZnSnSe4 0.83 34.39 75.43 21.63
CBTS 0.83 34.41 72.25 21.59
CuO 0.83 34.41 73.16 20.92
MoS2 0.80 33.52 74.16 19.85
MoOX 0.83 34.50 65.76 18.91
MoO3 0.83 34.32 60.87 17.42
PTTA 0.83 34.28 52.67 15.06
P3HT 0.83 34.24 46.44 13.26
PEDOT:PSS 0.83 34.20 38.71 11.02
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Seyed-Talebi, S.M.; Mahmoudi, M.; Lee, C.-H. A Comprehensive Study of CsSnI3-Based Perovskite Solar Cells with Different Hole Transporting Layers and Back Contacts. Micromachines 2023, 14, 1562. https://doi.org/10.3390/mi14081562

AMA Style

Seyed-Talebi SM, Mahmoudi M, Lee C-H. A Comprehensive Study of CsSnI3-Based Perovskite Solar Cells with Different Hole Transporting Layers and Back Contacts. Micromachines. 2023; 14(8):1562. https://doi.org/10.3390/mi14081562

Chicago/Turabian Style

Seyed-Talebi, Seyedeh Mozhgan, Mehrnaz Mahmoudi, and Chih-Hao Lee. 2023. "A Comprehensive Study of CsSnI3-Based Perovskite Solar Cells with Different Hole Transporting Layers and Back Contacts" Micromachines 14, no. 8: 1562. https://doi.org/10.3390/mi14081562

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