Next Article in Journal
Applications of Water Jet Cutting Technology in Agricultural Engineering: A Review
Previous Article in Journal
Genesis and Significance of Late Cretaceous Granitic Magmatism in Xianghualing Tin–Polymetallic Orefield, Nanling Region, South China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of the Power Conversion Efficiency of CsPbIxBr3−x-Based Perovskite Photovoltaic Solar Cells Using ZnO and NiOx as an Inorganic Charge Transport Layer

Department of Hydrogen and Renewable Energy, Kyungpook National University, Daegu 41566, Korea
*
Author to whom correspondence should be addressed.
Submission received: 10 August 2022 / Revised: 30 August 2022 / Accepted: 6 September 2022 / Published: 7 September 2022
(This article belongs to the Section Energy Science and Technology)

Abstract

:
In this study, we analyzed the maximum power conversion efficiency (PCE) of a photovoltaic cell with an ITO/ZnO/CsPbIxBr3−x/NiOx/Au structure, using ZnO and NiOx as the inorganic charge transport layers and CsPbIxBr3−x as an absorption layer. We optimized the thickness of each layer and investigated the effects of the defect density and interface defect density. To achieve the highest PCE, the optimal thicknesses were 300 nm for the electron transport layer (ZnO), 60 nm for the hole transport layer (NiOx), and 1000 nm for the absorption layer. The absorber defect density was maintained at approximately 1015 cm−3, and the interface defect density was approximately 1011 cm−3. The highest PCE obtained through optimization of each of these factors was 23.07%. These results are expected to contribute to the performance optimization of perovskite solar cells that use inorganic charge carrier transport layers.

1. Introduction

Recently, owing to the depletion of fossil fuel energy and increasing global warming concerns, interest in renewable energy has increased. In particular, photovoltaic solar cells can be used in most places on earth, and their energy is abundant. Thus, many studies on solar cells have been conducted [1,2,3]. Specifically, since perovskite of the ABX3 structure was first discovered, solar cells have developed more rapidly. Perovskite has a long diffusion length, large absorption coefficient (104 cm−1), and low price [4,5]. In the ABX3 structure, A contains large cations such as Cs+, methylammonium (MA+, CH3NH3+), and formamidine (FA+, CH3CH2NH3+), B contains metal cations such as Pb2+ and Sn2+, and X contains anions such as Cl, Br, and I [6]. Organic and organic-inorganic hybrid perovskite are widely applied in composing solar cell devices including perovskite [7]. Recently, perovskite solar cells (PSCs) have been reported to exhibit a power conversion efficiency (PCE) of nearly 25% [8,9]. However, despite the high PCE, most organic electron transport layers (ETLs) and hole transport layers (HTLs) constituting these PSCs remain sensitive to moisture [10,11]. Owing to these characteristics, encapsulation of the entire device is required to achieve long-term reliability [12]. In addition, organic-inorganic hybrid solar cells have issues such as band alignment and recombination losses and are also a problem in terms of long-term stability [13]. Therefore, interest in all-inorganic solar cell structures, in which inorganic materials are applied to devices, is increasing [14]. Generally, metal oxide semiconductors have high electron mobility (>10 cm2/Vs) so that a separate doping process is not required, as well as a wide band gap (Eg > 3 eV) and visible light region that have excellent transmittance [15]. The p-type metal oxide semiconductors include MoO3, Cu2O, CuI, CuS, and NiOx, and the n-type metal oxide includes SnO2, ZnO, In2O3, and Ga2O3 [16,17,18]. Among the n-type materials, ZnO has a wide band gap energy (Eg = 3.26 eV) and excellent mobility, while the p-type NiOx is chemically stable and has excellent electron-blocking properties [19,20,21]. Therefore, in this study, ZnO and NiOx were selected as the ETL and HTL materials, respectively, based on these excellent characteristics and the band alignment of the energy band diagram [22]. CsPbIxBr3−x was used as the perovskite absorption layer because it has been reported to be superior to other perovskite materials in terms of its moisture, thermal, and light stabilities [23,24].
Although many related papers have reported on all-inorganic PSCs [25,26,27], few studies have systematically analyzed the effect of each factor (e.g., thickness, defect density, and interface defect density) in a solar cell device using ZnO ETL and NiOx HTL with CsPbIxBr3−x as the inorganic absorbing layer on a solar cell’s performance [28]. Therefore, in this study, the effect of each layer’s thickness, defect density, and interface defect density on solar cell performance is systematically analyzed, and an optimized solar cell structure is suggested based on an all-inorganic solar cell with a device structure of inorganic ZnO ETL/absorbing layer CsPbIxBr3−x/inorganic NiOx HTL.

2. Simulation Procedure

In this study, as shown in Figure 1, ITO/ZnO/CsPbIxBr3−x/NiOx/Au PSC devices with an n-i-p structure were employed. The simulation was performed using SCAPS-1D simulation software (Burgelman et al., 2000) [29]. This tool is based on Poisson’s equation (Equation (1)) and the continuity equation for each electron and hole (Equations (2) and (3)), and it helps to calculate the short-circuit current density (Jsc), open circuit voltage (Voc), power conversion efficiency (PCE), fill factor (FF), quantum efficiency (QE), and recombination current, among others [30]. It is also based on the drift–diffusion equation (Equation (4)), which causes carrier transport in semiconductors [31]:
𝒹 2 φ ( 𝓍 ) / 𝒹 𝓍 2 = / 0 r ( p ( 𝓍 ) 𝓃 ( 𝓍 ) + N D N A + ρ p ρ 𝓃 )
In Poisson’s equation, we have the following:
  • φ = the electrostatic potential;
  • e = electrical charge;
  • εr = relative permittivity;
  • ε0 = vacuum permittivity;
  • ND = charged impurities of the donor;
  • NA = charged impurities of the acceptor;
  • ρ p = distribution of holes;
  • ρ 𝓃 = distribution of electrons.
𝒹 J n / 𝒹 𝓍 = G R
𝒹 J p / 𝒹 𝓍 = G R
In the continuity equation, we have the following:
  • R = the recombination rate;
  • Jn = the electron current density;
  • Jp = the hole current density;
  • G = the generation rate.
J n ( 𝓍 ) = q μ n n ( 𝓍 ) ( 𝓍 ) + qD n d n ( x ) d x ,   J p ( 𝓍 ) = q μ p p ( 𝓍 ) ( 𝓍 ) qD p d p ( x ) d x
In the above drift–diffusion model, we have the following:
  • q = the charge of the electron;
  • μ n = the mobility of the electrons;
  • μ p = the mobility of the holes;
  • n ( 𝓍 ) = electron distribution;
  • p ( 𝓍 )   = hole distribution;
  • ( 𝓍 ) = the electric field;
  • Dn = the electron diffusion coefficient;
  • Dp = the hole diffusion coefficient.
Figure 1a shows the basic structure of the all-inorganic multi-layer perovskite solar cell (PSC) in the study. ITO, ZnO, CsPbIxBr3−x, NiOx, and Au were used as the front metal contact, ETL, perovskite active layer (PAL), HTL, and back metal contact, respectively. The energy band alignment of this multi-layer perovskite solar cell (PSC) is shown in Figure 1b. The work functions of ITO (used as a front contact) and Au (used as a back contact) were −4.52 eV and −5.1 eV, respectively [32,33]. The lowest unoccupied molecular orbital (LUMO) levels and the highest occupied molecular orbital (HOMO) levels of the ZnO, CsPbIxBr3−x, and NiOx were −4.1 eV, −3.95 eV, and −1.46 eV and −7.4 eV, −5.73 eV, and −5.26 eV, respectively [34,35,36,37]. Through this optimized energy band alignment, it can be seen that the electrons and holes generated in the CsPbIxBr3−x absorption layer can efficiently move to the ITO electrode and the Au electrode, respectively. This device analysis was conducted under AM1.5G spectrum (1000 W/m, T = 300 K) conditions.
Table 1 summarizes the parameters of the PSC layers used in this study with SCAPS simulation software. For the numerical values of the ETL, HTL, and absorption layer, the previously published literature was referenced [34,35,36,37]. For the initial PSC device configuration, the thickness of each layer of ZnO, CsPbIxBr3−x, and NiOx was 100 nm, 800 nm, and 50 nm, respectively. After that, optimization simulation was performed for each thickness to increase the power conversion efficiency first. The band gaps of ZnO, CsPbIxBr3−x, and NiOx were 3.3, 1.78, and 3.8 eV, respectively, and the electron affinity was 4.1, 3.95, and 1.46 eV, respectively. Relative dielectric permittivity was set to 9, 6, and 10.7, respectively. Refer to Table 1 for the remaining conduction band (CB) effective densities of the states, valence band (VB) effective densities of the states, electron and hole thermal mobilities, electron and hole mobilities, shallow uniform donor densities, shallow uniform acceptor densities, and defect densities. For the interface defect between ZnO/CsPbIxBr3−x, the defect type was set to neutral, the capture cross-section of electrons and holes was set to 1.0 × 10−15 cm2, and the total density value was 1.0 × 1011 cm−2 [30]. For the interface defect between CsPbIxBr3−x and NiOx, the defect type was set to neutral, the capture cross-sections of the electrons and holes were set to 1.0 × 10−18 cm2 and 1.0 × 10−16 cm2, respectively, and the total density value was set to 1.0 × 1012 cm−2 [37].

3. Results and Discussion

3.1. Performance of the Perovskite Solar Cell (PSC) under the Initial Conditions

Figure 2 shows the current density–voltage (JV) and QE of the ITO/ZnO/CsPbIxBr3−x/NiOx/Au PSC device under the initial conditions. The Voc, Jsc, and FF were measured to be 1.03 V, 23.86 mA/cm2, and 67.27%, respectively. The FF and PCE values were calculated using Equations (5) and (6) [38]. Accordingly, the PCE was determined to be 16.61%. Based on these basic performance indicators, from Section 3.2, the effect of the layer thickness, defect density, and interface defect density on the PCE were systematically analyzed:
FF = V max J max V oc J sc
PCE = V oc J sc FF P in
In these equations, we use the following:
  • FF = the fill factor;
  • Vmax = the maximum voltage;
  • Jmax = the maximum current density;
  • Voc = the open-circuit voltage;
  • Jsc = the short-circuit current density;
  • PCE = power conversion efficiency;
  • Pin = the incident power of the sun.

3.2. Effect of the ZnO Thickness

ZnO applied as an ETL in an n-i-p device as the first layer that absorbed the incident light. Therefore, the thickness of the ZnO layer was optimized in the range of 100–1000 nm. The current density–voltage (J–V) curves with respect to the ZnO thickness are shown in Figure 3. At a ZnO thickness of 300 nm, the highest values of Voc, Jsc, FF, and PCE were 1.04 V, 23.3 mA/cm2, 80.2%, and 19.4%, respectively. Then, as the ZnO thickness increased, each output parameter values showed a saturated behavior. The reason that the PCE increased with the ETL thickness from 100 nm to 300 nm may be attributed to the increase in dissociation of the photo-generated holes and electrons and reduced recombination in the photovoltaic solar cells [39]. When the thickness exceeded 300 nm, as recombination also increased as the thickness of ZnO increased, the PCE did not significantly increase and was considered to show a behavior of saturation. A similar behavior can be found in [39].

3.3. Effect of the Perovskite Absorption Layer (PAL) Thickness

In a photovoltaic solar cell, the absorption layer absorbs light and creates a hole-electron pair. Thickness optimization is important because the amount of absorbed light can vary depending on the thickness of the absorption layer. In this study, such analysis was conducted in the PAL thickness range of 100–1000 nm. Figure 4 presents the current density–voltage (J–V) curves for various PAL thicknesses. In Figure 5, each output parameter value (i.e., Voc, Jsc, FF, and PCE) is shown according to the thickness of the PAL.
As the PAL thickness increased, the Voc and Jsc exhibited increasing behavior. However, the FF value decreased at PAL thicknesses of above 500 nm. This behavior was somewhat different from those of the Voc and Jsc. This can be determined from the relationship between the thickness of the absorption layer and resistance and is expressed as shown in Equations (7) and (8) [40].
As the thickness of the absorption layer increased, the resistance of the absorption layer also increased, owing to an increase in the series resistance. This increased series resistance led to a decrease in the FF value, as given by Equation (9) [41,42]. Therefore, the FF value was considered to decrease above a PAL thickness of 500 nm:
R absorber = ρ t A 2
R series ρ t + R s L 2 2
FF series = FF 0 ( 1 r series )
Here, we use the following:
  • Rabsorber = the absorber resistance;
  • ρ = the absorber layer resistivity;
  • t = the absorber layer’s bulk region thickness;
  • A = the base area;
  • Rseries = the series resistance;
  • L = length;
  • Rs = sheet resistance;
  • FFseries = the FF affected by the series resistance;
  • F0 = the initial FF;
  • rseries = normalized series resistance (Rseries Jsc/Voc).
In the case of a PCE above 800 nm, it appeared to be saturated. At a PAL thickness of 1000 nm, the PCE showed the highest value of 19.7%. The more light was absorbed, the more hole-electron pairs which could be created, which increased each output parameter [43]. However, if the thickness was too thick, the carrier movement path and recombination current increased. Thus, it was necessary to select an appropriate thickness [44]. Accordingly, 1000 nm was selected as the optimal PAL thickness.

3.4. Effect of the NiOx Thickness

In this study, NiOx was used as a hole transport layer (HTL), and its thickness was analyzed in the range of 30–60 nm. The current density–voltage (J–V) data for varying NiOx thicknesses are depicted in Figure 6. The Voc did not change significantly until 60 nm, whereas the Jsc increased from 22.8 to 25.6 mA/cm2. These results indicate that the PCE increased from 18.5% to 20.3%.

3.5. Device Performance Optimization of a PSC

Based on the optimization of the ETL, HTL, and perovskite layer thicknesses, the thicknesses for each layer were 300 nm for ZnO, 1000 nm for PAL, and 60 nm for NiOx. As shown in Figure 7a,b, the optimal Voc, Jsc, FF, and PCE values simulated using each of these optimized layers increased from 1.03 to 1.05 V, 23.86 to 25.56 mA/cm2, 67.27% to 75.75%, and 16.61% to 20.26%, respectively, when compared with those of the initial reference device structure. With this optimized structure, the effects of the PAL defect density and the interface defect density between ZnO/CsPbIxBr3−x and CsPbIxBr3−x/NiOx are also analyzed in the following section.

3.6. Effect of the PAL Defect Density

The defect density in the absorption layer is one of the major factors affecting the device performance because it can act as a trap when the hole and electron carriers move to each electrode [45]. A corresponding analysis was performed in the range of 1011–1018 cm3, and the results are shown in Figure 8 and Figure 9. Figure 8 shows each output parameter, such as the Voc, Jsc, FF, and PCE, according to the defect density. All output parameters showed a constant value until the defect density was 1015/cm3 but showed a sharp decrease when it exceeded 1015/cm3. Figure 9 presents the current density–voltage (JV) curves for the various PAL defect densities. As the defect density increased, the Voc, Jsc, FF, and PCE decreased significantly from 1.05 to 0.99 V, 25.6 to 11.9 mA/cm2, 76.1% to 45.8%, and 20.4% to 5.44%, respectively. These results were thought to be caused by Shockley–Read–Hall (SRH) recombination [46]. According to the SRH recombination, as the defect density increases, the carrier lifetime and diffusion length decrease. This causes an increase in SRH recombination, as shown in Figure 9. Thus, the PCE of this device decreased.
Moreover, it can be seen from Figure 10 that the recombination current increased rapidly above 1016 cm−3. Therefore, in this study, the PAL defect density was kept below 1015 cm3.

3.7. Effect of the Interface Defect Density

Because the holes and electron carriers generated by light must pass through the interface to reach each electrode, the interface defect density is important for device performance. The effects of the ZnO/PAL (CsPbIxBr3−x) interface and PAL (CsPbIxBr3−x)/NiOx interface defect density were investigated in the interface defect density range of 108–1014 cm−3. Figure 11a,b illustrates the current density–voltage (J–V) curves for the ZnO/PAL and PAL/NiOx interfaces, respectively.
In the case of the ZnO/PAL interface, as the interface defect density increased from 108 cm−3 to 1014 cm−3, each of the output parameters (Voc, Jsc, FF, and PCE) decreased from 1.06 V, 25.6 mA/cm2, 76.0%, and 20.5% to 0.96 V, 25.0 mA/cm2, 72.4%, and 17.3%, respectively.
In the case of PAL/NiOx, as the interface defect density increased from 108 cm−3 to 1014 cm−3, the Voc, Jsc, FF, and PCE decreased from 1.13 V, 25.6 mA/cm2, 81.8, and 23.7% to 1.01 V, 23.8 mA/cm2, 56.0, and 13.4%, respectively.
Even if the carriers have a diffusion length that can reach each the electrode without recombination, severe recombination can occur because of an increase in the interface defect density [47]. Therefore, when stacking layers, a reduction in the interface defect density is recommended.
In this analysis, corresponding values of 1011 cm−3 or less were considered because both interface defect densities exhibited sharp decreases in efficiency above 1012 cm−3. Therefore, the interface defect density was fixed at 1011 cm−3 for both the ZnO/PAL and PAL/NiOx interfaces. Following this optimization process, the Voc, Jsc, and FF obtained were 1.09 V, 25.58 mA/cm2, and 82.89%, respectively, and as a result, a maximum PCE of 23.07% was realized.

3.8. The Effect of Back Contact Materials

The work function is defined as the minimum energy of the photon required for an electron to come out of a metal surface [48]. In order to find out the effect of the work function of the back contact, as summarized in Table 2, various electrodes such as Cu, Ag, Fe, and Au were used [33].
Figure 12 shows the current density–voltage (J–V) curves for various back contact materials. It can be seen that the output parameter value increased as the work function value increased. The Voc increased from 0.87 V to 1.04 V, and the PCE increased from 17.36% to 23.07%. This is consistent with the results of the previous literature, which reported that solar cell efficiency increased as the work function value increased [49]. As the work function value increased, the barrier height value of the majority carriers decreased, which made the electrode more of an ohmic type. Therefore, the Voc and PCE values increased. Accordingly, the best results were obtained when Au, which had the largest work function of 5.1 eV, was used as the back electrode in this experiment.

4. Conclusions

In this study, the layer thickness, defect density, and interface defect density for an all-inorganic PSC device were optimized using CsPbIxBr3−x, ZnO, and NiOx as the inorganic absorber layer, ETL, and HTL, respectively. The optimal thicknesses were determined to be 300 nm for ZnO, 1000 nm for the absorption layer, and 60 nm for NiOx. Based on this optimized structure, a Voc of 1.05 V, Jsc of 25.56 mA/cm2, FF of 75.75%, and PCE of 20.26% were achieved. In addition, the effects of the absorber defect density and interface defect density between ETL/PAL and PAL/NiOx were analyzed. Based on this, maintaining the absorber layer defect density at approximately 1015 cm−3 and the interface defect density at 1011 cm−3 or less is recommended. Through optimization, we achieved a Voc of 1.09 V, Jsc of 25.58 mA/cm2, FF of 82.89%, and PCE of 23.07%. Thus, this study on the optimization of each factor of a solar cell is expected to improve the performance of inorganic PSC devices.

Author Contributions

Conceptualization, H.S. and B.-S.J.; software, H.S.; methodology, H.S. and B.-S.J.; formal analysis, H.S. and B.-S.J.; writing—original draft preparation, H.S.; writing—review and editing, B.-S.J.; supervision, B.-S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Industrial Technology Innovation Program funded by the Ministry of Trade, Industry, and Energy (MOTIE, Korea) (No. 20015778).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gross, R.; Leach, M.; Bauen, A. Progress in Renewable Energy. Environ. Int. 2003, 29, 105–122. [Google Scholar] [CrossRef]
  2. Ludin, N.A.; Mustafa, N.I.; Hanafiah, M.M.; Ibrahim, M.A.; Asri Mat Teridi, M.; Sepeai, S.; Zaharim, A.; Sopian, K. Prospects of Life Cycle Assessment of Renewable Energy from Solar Photovoltaic Technologies: A Review. Renew. Sustain. Energy Rev. 2018, 96, 11–28. [Google Scholar] [CrossRef]
  3. Green, M.A.; Hishikawa, Y.; Dunlop, E.D.; Levi, D.H.; Hohl-Ebinger, J.; Ho-Baillie, A.W.Y. Solar Cell Efficiency Tables (Version 51). Prog. Photovolt. 2018, 26, 3–12. [Google Scholar] [CrossRef]
  4. Park, N.-G. Perovskite solar cells: An emerging photovoltaic technology. Mater. Today 2015, 18, 65–72. [Google Scholar] [CrossRef]
  5. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef]
  6. Becker, M.; Klüner, T.; Wark, M. Formation of hybrid ABX3 perovskite compounds for solar cell application: First-principles calculations of effective ionic radii and determination of tolerance factors. Dalton Trans. 2017, 46, 3500–3509. [Google Scholar] [CrossRef]
  7. Luo, S.; Daoud, W.A. Recent progress in organic-inorganic halide perovskite solar cells: Mechanisms and materials design. J. Mater. Chem. A 2012, 3, 8992–9010. [Google Scholar] [CrossRef]
  8. Kim, M.; Jeong, J.; Lu, H.; Lee, T.-K.; Eickemeyer, T.; Liu, Y.; Choi, I.-W.; Choi, S.-J.; Jo, Y.-H.; Kim, H.-B.; et al. Conformal Quantum dot-SnO2 layers as electron transporters for efficient perovskite solar cells. Science 2022, 375, 302–306. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, T.; Wang, F.; Kim, H.-B.; Choi, I.-W.; Wang, C.; Cho, E.; Konefal, R.; Puttisong, Y.; Terado, K.; Kobera, L.; et al. Ion-modulated radical doping of spiro-OMeTAD for more efficient and stable perovskite solar cells. Science 2022, 375, 495–501. [Google Scholar] [CrossRef]
  10. Smecca, E.; Numata, Y.; Deretzis, I.; Pellegrino, G.; Boninelli, S.; Miyasaka, T.; La Magna, A.; Alberti, A. Stability of Solution-Processed MAPbI3 and FAPbI3 Layers. Phys. Chem. Chem. Phys. 2016, 18, 13413–13422. [Google Scholar] [CrossRef]
  11. Yang, J.; Kelly, T.L. Decomposition and Cell Failure Mechanisms in Lead Halide Perovskite Solar Cells. Inorg. Chem. 2017, 56, 92–101. [Google Scholar] [CrossRef]
  12. Uddin, A.; Upama, M.B.; Yi, H.; Duan, L. Encapsulation of Organic and Perovskite Solar Cells: A Review. Coatings 2019, 9, 65. [Google Scholar] [CrossRef]
  13. Zheng, Y.; Yang, X.; Su, R.; Wu, P.; Gong, Q.; Zhu, R. High-Performance CsPbIxBr3-x All-Inorganic Perovskite Solar Cells with Efficiency over 18% via Spontaneous Interfacial Manipulation. Adv. Funct. Mater. 2020, 30, 2000457. [Google Scholar] [CrossRef]
  14. Liang, J.; Wang, C.; Wang, Y.; Xu, Z.; Lu, Z.; Ma, Y.; Zhu, H.; Hu, Y.; Xiao, C.; Yi, X.; et al. All-Inorganic Perovskite Solar Cells. J. Am. Chem. Soc. 2016, 138, 15829–15832. [Google Scholar] [CrossRef] [PubMed]
  15. Takagi, A.; Nomura, K.; Ohta, H.; Yanagi, H.; Kamiya, T.; Hirano, M.; Hosono, H. Carrier transport and electronic structure in amorphous oxide semiconductor, a-InGaZnO4. Thin Solid Films 2005, 486, 38–41. [Google Scholar] [CrossRef]
  16. Yuan, H.; Zhao, Y.; Duan, J.; Wang, Y.; Yang, X.; Tang, Q. All-Inorganic CsPbBr3 Perovskite Solar Cell with 10.26% Efficiency by Spectra Engineering. J. Mater. Chem. A 2018, 6, 24324–24329. [Google Scholar] [CrossRef]
  17. Petti, L.; Münzenrieder, N.; Vogt, C.; Faber, H.; Büthe, L.; Cantarella, G.; Bottacchi, F.; Anthopoulos, T.D.; Tröster, G. Metal Oxide Semiconductor Thin-Film Transistors for Flexible Electronics. Appl. Phys. Rev. 2016, 3, 021303. [Google Scholar] [CrossRef]
  18. Chen, J.; Park, N.G. Inorganic Hole Transporting Materials for Stable and High Efficiency Perovskite Solar Cells. J. Phys. Chem. C 2018, 122, 14039–14063. [Google Scholar] [CrossRef]
  19. Craciun, V.; Elders, J.; Gardeniers, J.G.E.; Boyd, I.W. Characteristics of High Quality ZnO Thin Films Deposited by Pulsed Laser Deposition. Appl. Phys. Lett. 1994, 65, 2963–2965. [Google Scholar] [CrossRef]
  20. Lee, W.; Lee, C.; Yu, H.; Kim, D.J.; Wang, C.; Woo, H.Y.; Oh, J.H.; Kim, B.J. Side Chain Optimization of Naphthalenediimide-Bithiophene-Based Polymers to Enhance the Electron Mobility and the Performance in All-Polymer Solar Cells. Adv. Funct. Mater. 2016, 26, 1543–1553. [Google Scholar] [CrossRef]
  21. Hossain, M.I.; Hasan, A.K.M.; Qarony, W.; Shahiduzzaman, M.; Islam, M.A.; Ishikawa, Y.; Uraoka, Y.; Amin, N.; Knipp, D.; Akhtaruzzaman, M.; et al. Electrical and Optical Properties of Nickel-Oxide Films for Efficient Perovskite Solar Cells. Small Methods 2020, 4, 2000454. [Google Scholar] [CrossRef]
  22. Hussain, I.; Tran, H.P.; Jaksik, J.; Moore, J.; Islam, N.; Uddin, M.J. Functional Materials, Device Architecture, and Flexibility of Perovskite Solar Cell. Emerg. Mater. 2018, 1, 133–154. [Google Scholar] [CrossRef]
  23. Zeng, Q.; Zhang, X.; Liu, C.; Feng, T.; Chen, Z.; Zhang, W.; Zheng, W.; Zhang, H.; Yang, B. Inorganic CsPbI2Br Perovskite Solar Cells: The Progress and Perspective. Sol. RRL 2018, 3, 1800239. [Google Scholar] [CrossRef]
  24. Ho-Baillie, A.; Zhang, M.; Lau, C.F.J.; Ma, F.J.; Huang, S. Untapped Potentials of Inorganic Metal Halide Perovskite Solar Cells. Joule 2019, 3, 938–955. [Google Scholar] [CrossRef]
  25. Liang, J.; Zhao, P.; Wang, C.; Wang, Y.; Hu, Y.; Zhu, G.; Ma, L.; Liu, J.; Jin, Z. CsPb0.9Sn0.1IBr2 Based All-Inorganic Perovskite Solar Cells with Exceptional Efficiency and Stability. J. Am. Chem. Soc. 2017, 139, 14009–14012. [Google Scholar] [CrossRef]
  26. Liu, C.; Li, W.; Li, H.; Wang, H.; Zhang, C.; Yang, Y.; Gao, X.; Xue, Q.; Yip, H.; Fan, J. Structurally Reconstructed CsPbI2Br perovskite for Highly Stable and Square-Centimeter All-Inorganic Perovskite Solar Cells. Adv. Energy Mater. 2018, 9, 1803572. [Google Scholar] [CrossRef]
  27. Guo, Z.; Zhao, S.; Liu, A.; Kamata, Y.; Teo, S.; Yang, S.; Xu, Z.; Hayase, S.; Ma, T. Niobium Incorporation into CsPbI2Br for stable and Efficient All-Inorganic Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2019, 11, 19994–20003. [Google Scholar]
  28. Liu, J.; Lei, M.; Zhang, W.; Wang, G. A Multifunctional CuSCN Interlayer in Carbon Electrode-Based CsPbIBr2 All-Inorganic Perovskite Solar Cells for Boosting the Efficiency and Stability. Sustain. Energy Fuels 2020, 4, 4249–4256. [Google Scholar] [CrossRef]
  29. Decock, M.B.K.; Niemegeers, A.; Verschraegen, J.; Degrave, S. SCAPS Manual; University of Gent: Gent, Belgium, 2016. [Google Scholar]
  30. Mandadapu, U.; Vedanayakam, S.V.; Thyagarajan, K.; Babu, B.J. Optimisation of High Efficiency Tin Halide Perovskite Solar Cells Using SCAPS-1D. Int. J. Simul. Process. Model. 2018, 13, 221–227. [Google Scholar] [CrossRef]
  31. Ben, G.; Streetman, S.B. Solid State Electronic Devices, 5th ed.; Prentice Hall: Hoboken, NJ, USA, 2000; pp. 127–130. [Google Scholar]
  32. Song, Y.; Yan, L.; Zhou, Y.; Song, B.; Li, Y. Lowering the Work Function of ITO by Covalent Surface Grafting of Aziridine: Application in Inverted Polymer Solar Cells. Adv. Mater. Interfaces 2015, 2, 1400397. [Google Scholar] [CrossRef]
  33. Sawicka-Chudy, P.; Starowicz, Z.; Wisz, G.; Yavorskyi, R.; Zapukhlyak, Z.; Bester, M.; Głowa, Ł.; Sibiński, M.; Cholewa, M. Simulation of TiO2/CuO Solar Cells with SCAPS-1D Software. Mater. Res. Express 2019, 6, 085918. [Google Scholar] [CrossRef]
  34. Bhavsar, K.; Lapsiwala, P.B. Numerical Simulation of Perovskite Solar Cell with Different Material as Electron Transport Layer Using SCAPS-1D Software. Semicond. Phys. Quantum Electron. Optoelectron. 2021, 24, 341–347. [Google Scholar] [CrossRef]
  35. Martínez, N.; Pinzón, C.; Casas, G.; Alvira, F.; Cappelletti, M. Computational Study of Inverted All-Inorganic Perovskite Solar Cells Based on CsPbIxBr3-x Absorber Layer with Band Gap of 1.78 eV. Energy 2020. [Google Scholar] [CrossRef]
  36. Hossain, M.I.; Alharbi, F.H.; Tabet, N. Copper Oxide as Inorganic Hole Transport Material for Lead Halide Perovskite Based Solar Cells. J. Sol. Energy 2015, 120, 370–380. [Google Scholar] [CrossRef]
  37. Anwar, F.; Mahbub, R.; Satter, S.S.; Ullah, S.M. Effect of Different HTM Layers and Electrical Parameters on ZnO Nanorod-Based Lead-Free Perovskite Solar Cell for High-Efficiency Performance. J. Spectrosc. 2017, 2017, 9846310. [Google Scholar] [CrossRef]
  38. Kitai, A. Principles of Solar Cells, LEDs and Diodes: The Role of the PN Junction; John Wiley & Sons: Hoboken, NJ, USA, 2011; p. 170. [Google Scholar]
  39. Gupta, S.K.; Sharma, A.; Banerjee, S.; Gahlot, R.; Aggarwal, N.; Deepak; Garg, A. Understanding the role of thickness and morphology of the constituent layers on the performance of inverted organic solar cells. Sol. Energy Mater. Sol. Cells 2013, 116, 135–143. [Google Scholar] [CrossRef]
  40. Jannat, F.; Ahmed, S.; Alim, M.A. Performance Analysis of Cesium Formamidinium Lead Mixed Halide Based Perovskite Solar Cell with MoOx as Hole Transport Material via SCAPS-1D. Optik 2021, 228, 166202. [Google Scholar] [CrossRef]
  41. Jao, M.H.; Liao, H.C.; Su, W.F. Achieving a High Fill Factor for Organic Solar Cells. J. Mater. Chem. A 2016, 4, 5784–5801. [Google Scholar] [CrossRef]
  42. Khanna, A.; Mueller, T.; Stangl, R.A.; Hoex, B.; Basu, P.K.; Aberle, A.G.; Fill Factor Loss, A. Analysis Method for Silicon Wafer Solar Cells. IEEE J. Photovolt. 2013, 3, 1170–1177. [Google Scholar] [CrossRef]
  43. Khoshsirat, N.; Yunus, N.A.M.; Hamidon, M.N.; Shafie, S.; Amin, N. Analysis of Absorber Layer Properties Effect on CIGS Solar Cell Performance Using SCAPS. Optik 2015, 126, 681–686. [Google Scholar] [CrossRef]
  44. Liu, D.; Gangishetty, M.K.; Kelly, T.L. Effect of CH3NH3PbI3 Thickness on Device Efficiency in Planar Heterojunction Perovskite Solar Cells. J. Mater. Chem. A 2014, 2, 19873–19881. [Google Scholar] [CrossRef] [Green Version]
  45. Abdelaziz, S.; Zekry, A.; Shaker, A.; Abouelatta, M. Investigating the Performance of Formamidinium Tin-Based Perovskite Solar Cell by SCAPS Device Simulation. Opt. Mater. 2020, 101, 109738. [Google Scholar] [CrossRef]
  46. Wen, X.; Feng, Y.; Huang, S.; Huang, F.; Cheng, Y.B.; Green, M.; Ho-Baillie, A. Defect Trapping States and Charge Carrier Recombination in Organic–Inorganic Halide Perovskites. J. Mater. Chem. C 2016, 4, 793–800. [Google Scholar] [CrossRef]
  47. Sherkar, T.S.; Momblona, C.; Gil-Escrig, L.; Ávila, J.; Sessolo, M.; Bolink, H.J.; Koster, L.J.A. Recombination in Perovskite Solar Cells: Significance of Grain Boundaries, Interface Traps, and Defect Ions. ACS Energy Lett. 2017, 2, 1214–1222. [Google Scholar] [CrossRef] [PubMed]
  48. Anwar, F.; Afrin, S.; Satter, S.; Mahbub, R.; Ullah, S. Simulation and Performance Study of Nanowire CdS/CdTe Solar Cell. Int. J. Renew. Energy Res. 2017, 7, 885–893. [Google Scholar]
  49. Thahab, S.M.; Hassan, H.A.; Hassan, Z. Effects of Metal Work Function and Operating Temperatures on the Electrical Properties of Contacts to n-type GaN. In Proceedings of the IEEE International Conference on Semiconductor Electronics, Kuala Lumpur, Malaysia, 29 October 2006–1 December 2006; pp. 816–819. [Google Scholar]
Figure 1. (a) The basic structure of an all-inorganic multi-layer perovskite solar cell (PSC). (b) The energy band alignment of an all-inorganic perovskite solar cell (PSC).
Figure 1. (a) The basic structure of an all-inorganic multi-layer perovskite solar cell (PSC). (b) The energy band alignment of an all-inorganic perovskite solar cell (PSC).
Applsci 12 08987 g001
Figure 2. (a) Current density–voltage (J–V) and (b) quantum efficiency (QE) curve for ITO/ZnO/CsPbIxBr3−x/NiOx/Au structure.
Figure 2. (a) Current density–voltage (J–V) and (b) quantum efficiency (QE) curve for ITO/ZnO/CsPbIxBr3−x/NiOx/Au structure.
Applsci 12 08987 g002
Figure 3. Current density–voltage (J–V) curves for various ZnO thicknesses.
Figure 3. Current density–voltage (J–V) curves for various ZnO thicknesses.
Applsci 12 08987 g003
Figure 4. Current density–voltage (J–V) curves for various PAL thicknesses.
Figure 4. Current density–voltage (J–V) curves for various PAL thicknesses.
Applsci 12 08987 g004
Figure 5. Effect of PAL thickness on (a) Voc, (b) Jsc, (c) FF, and (d) PCE.
Figure 5. Effect of PAL thickness on (a) Voc, (b) Jsc, (c) FF, and (d) PCE.
Applsci 12 08987 g005
Figure 6. Current density–voltage (J–V) curves for various NiOx thicknesses.
Figure 6. Current density–voltage (J–V) curves for various NiOx thicknesses.
Applsci 12 08987 g006
Figure 7. Current density–voltage (J–V) and QE curves (a) before and (b) after thickness optimization.
Figure 7. Current density–voltage (J–V) and QE curves (a) before and (b) after thickness optimization.
Applsci 12 08987 g007aApplsci 12 08987 g007b
Figure 8. Effect of PAL defect density on (a) Voc, (b) Jsc, (c) FF, and (d) PCE.
Figure 8. Effect of PAL defect density on (a) Voc, (b) Jsc, (c) FF, and (d) PCE.
Applsci 12 08987 g008
Figure 9. Current density–voltage (JV) curves for various PAL defect densities.
Figure 9. Current density–voltage (JV) curves for various PAL defect densities.
Applsci 12 08987 g009
Figure 10. Recombination current density with varying PAL defect densities.
Figure 10. Recombination current density with varying PAL defect densities.
Applsci 12 08987 g010
Figure 11. Current density–voltage (J–V) curves for varying interface defect densities at (a) ZnO/PAL and (b) PAL/NiOx interfaces.
Figure 11. Current density–voltage (J–V) curves for varying interface defect densities at (a) ZnO/PAL and (b) PAL/NiOx interfaces.
Applsci 12 08987 g011aApplsci 12 08987 g011b
Figure 12. Comparison of the current density-voltage (J–V) behavior for different work functions of back contact.
Figure 12. Comparison of the current density-voltage (J–V) behavior for different work functions of back contact.
Applsci 12 08987 g012
Table 1. Physical parameters of ZnO, CsPbIxBr3−x, and NiOx used in the ITO/ZnO/CsPbIxBr3−x/NiOx/Au PSC.
Table 1. Physical parameters of ZnO, CsPbIxBr3−x, and NiOx used in the ITO/ZnO/CsPbIxBr3−x/NiOx/Au PSC.
ParameterZnO [34]CsPbIxBr3−x [35]NiOx [36]
Thickness (nm)10080050
Band gap (eV)3.31.783.8
Electron affinity (eV)4.13.951.46
Dielectric permittivity (relative)9610.7
CB effective density of states (cm−3)4.0 × 10181.1 × 10202.8 × 1019
VB effective density of states (cm−3)1.0 × 10198.0 × 10191.0 × 1019
Electron thermal velocity (cm/s)1.0 × 1071.0 × 1071.0 × 107
Hole thermal velocity (cm/s)1.0 × 1071.0 × 1071.0 × 107
Electron mobility (cm2/Vs)1001612
Hole mobility (cm2/Vs)25162.8
Shallow uniform donor density ND (cm−3)1.0 × 1018--
Shallow uniform acceptor density NA (cm−3)-1.0 × 10151.0 × 1018
Defect density Nt (cm−3)1.0 × 10141.0 × 10141.0 × 1014
Table 2. Work function of various back contact materials for PSC.
Table 2. Work function of various back contact materials for PSC.
MaterialCuAgFeAu
Work function (eV)4.64.74.85.1
PCE (%)17.3619.6521.6223.07
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Son, H.; Jeong, B.-S. Optimization of the Power Conversion Efficiency of CsPbIxBr3−x-Based Perovskite Photovoltaic Solar Cells Using ZnO and NiOx as an Inorganic Charge Transport Layer. Appl. Sci. 2022, 12, 8987. https://0-doi-org.brum.beds.ac.uk/10.3390/app12188987

AMA Style

Son H, Jeong B-S. Optimization of the Power Conversion Efficiency of CsPbIxBr3−x-Based Perovskite Photovoltaic Solar Cells Using ZnO and NiOx as an Inorganic Charge Transport Layer. Applied Sciences. 2022; 12(18):8987. https://0-doi-org.brum.beds.ac.uk/10.3390/app12188987

Chicago/Turabian Style

Son, Hyojung, and Byoung-Seong Jeong. 2022. "Optimization of the Power Conversion Efficiency of CsPbIxBr3−x-Based Perovskite Photovoltaic Solar Cells Using ZnO and NiOx as an Inorganic Charge Transport Layer" Applied Sciences 12, no. 18: 8987. https://0-doi-org.brum.beds.ac.uk/10.3390/app12188987

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