Next Article in Journal
Bacterial Nanocellulose as a Scaffold for In Vitro Cell Migration Assay
Next Article in Special Issue
Solution-Processed SnO2 Quantum Dots for the Electron Transport Layer of Flexible and Printed Perovskite Solar Cells
Previous Article in Journal
Colored Surfaces Made of Synthetic Eumelanin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Numerical Investigation of a Lead-Free Inorganic Layered Double Perovskite Cs4CuSb2Cl12 Nanocrystal Solar Cell by SCAPS-1D

1
School of Optoelectronic Science and Engineering, University of Electronic Science and Technology of China, Chengdu 610054, China
2
Yangtze Delta Region Institute (Huzhou), University of Electronic Science and Technology of China, Huzhou 313001, China
3
Key Laboratory of Display Science and Technology of Sichuan Province, University of Electronic Science and Technology of China, Chengdu 610054, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2321; https://doi.org/10.3390/nano11092321
Submission received: 27 July 2021 / Revised: 31 August 2021 / Accepted: 3 September 2021 / Published: 7 September 2021
(This article belongs to the Special Issue Nanomaterials for Solar Energy Conversion and Storage)

Abstract

:
In the last decade, perovskite solar cells have made a quantum leap in performance with the efficiency increasing from 3.8% to 25%. However, commercial perovskite solar cells have faced a major impediment due to toxicity and stability issues. Therefore, lead-free inorganic perovskites have been investigated in order to find substitute perovskites which can provide a high efficiency similar to lead-based perovskites. In recent studies, as a kind of lead-free inorganic perovskite material, Cs4CuSb2Cl12 has been demonstrated to possess impressive photoelectric properties and excellent environmental stability. Moreover, Cs4CuSb2Cl12 nanocrystals have smaller effective photo-generated carrier masses than bulk Cs4CuSb2Cl12, which provides excellent carrier mobility. To date, there have been no reports about Cs4CuSb2Cl12 nanocrystals used for making solar cells. To explore the potential of Cs4CuSb2Cl12 nanocrystal solar cells, we propose a lead-free perovskite solar cell with the configuration of FTO/ETL/Cs4CuSb2Cl12 nanocrystals/HTL/Au using a solar cell capacitance simulator. Moreover, we numerically investigate the factors that affect the performance of the Cs4CuSb2Cl12 nanocrystal solar cell with the aim of enhancing its performance. By selecting the appropriate hole transport material, electron transport material, thickness of the absorber layer, doping densities, defect density in the absorber, interface defect densities, and working temperature point, we predict that the Cs4CuSb2Cl12 nanocrystal solar cell with the FTO/TiO2/Cs4CuSb2Cl12 nanocrystals/Cu2O/Au structure can attain a power conversion efficiency of 23.07% at 300 K. Our analysis indicates that Cs4CuSb2Cl12 nanocrystals have great potential as an absorbing layer towards highly efficient lead-free all-inorganic perovskite solar cells.

1. Introduction

In the last decade, lead-based perovskite solar cells (PSCs) have witnessed tremendous growth in photovoltaic applications due to their good optical and electrical properties [1]. Typical organic–inorganic hybrid PSC attained power conversion efficiency (PCE) of over 25% by 2020, and include methylammonium lead halide (MAPbX3) [2]. Despite these exciting developments, PSCs still face some challenges with respect to commercialization, e.g., their stability, and the toxic nature of lead [3,4,5]. These challenges can be addressed by developing an inorganic lead-free Perovskite absorbing layer [6,7]. Metal cations from the same family as Pb2+—Ge2+ and Sn2+—were first considered to replace Pb2+. However, Ge2+ and Sn2+ are highly susceptible to being oxidized to the tetravalent state (Ge 4+, Sn 4+) in air [8,9,10]. Subsequently, Bi3+ and Sb3+ were used as heterovalent substitutes for Pb2+ to synthesize two-dimensional or zero-dimensional chalcogenides such as Cs3Sb2X9, Cs3Bi2X9 (X = Cl, Br, I) [11,12,13,14,15]. Recently, double perovskite structure A2B’B”X6 has been proposed as a promising substitution. This structure is formed by replacing the lead ions in two adjacent lattices with a pair of nontoxic heterovalent (i.e., monovalent and trivalent) metal cations. However, the most typical double perovskite, Cs2AgBiBr6, is not suitable for photovoltaic (PV) applications due to its wide bandgap of 2.19 eV and the indirect nature of the bandgap [16,17,18].
To further explore lead-free inorganic perovskite candidates suitable for PV applications, Cs4CuSb2Cl12(CCSC) has been proposed. Experiments have demonstrated its impressive photoelectric properties, which include narrow direct bandgap (1.0 eV) and excellent environmental stability (under humidity, heat, and light conditions) [19,20]. However, the bulk CCSC has been inferred to exhibit a high electron effective mass, which results in tardy carrier mobility and consequently hinders its performance in solar cells and other optoelectronic devices [21]. Reduction of the particle size to the nanoscale has been widely demonstrated to be an effective strategy for tuning the energy band structure of materials in accordance with the quantum confinement effect [22,23,24]. In 2019, Kuang et al. fabricated CCSC nanocrystals (NCs) with an average particle size of ~3 nm using a top-down ultrasonic exfoliation technique and subsequently fabricated CCSCNC thin film (500nm) by centrifugal casting of the CCSCNC solution onto FTO glass, using oleic acid (OA) as the organic ligand [25]. The resulting NCs possessed a direct bandgap of 1.6 eV and low effective photo-generated carrier masses. In addition, CCSCNCs have been demonstrated to have excellent environmental stability (under heat, humidity, and light conditions). In 2020, Tong et al. fabricated a thin-film-based high-speed photodetector by casting high-concentration CCSCNC hexane solution, using OA, oleylamine (OAm), and 1-octadecene (ODE) as the ligands on a quartz substrate [26]. The excellent carrier mobility in the CCSCNCs were demonstrated using a high-speed photodetector, suggesting that CCSCNCs have great potential as the absorber layer for solar cells. In 2021, Ashitha P. et al. proved that CCSCNCs (~3.9 nm) can efficiently catalyze the ferricyanide reduction and dye degradation reactions as a photocatalyst, and CCSCNCs have strong absorption throughout the visible region [27]. NCs, as particles with one or more dimensions less than 100 nanometers (nm), have different properties from their bulk materials due to the quantum-confinement effect [28]. In particular, perovskite NCs are considered ideal candidates for next-generation photovoltaic applications due to their high electrical conductivity, broad absorption spectrum, variable band gap, and structural compatibility [28,29]. In recent years, perovskite NCs-based solar cells have developed rapidly, and the highest efficiency has reached 16.6% [30].
To the best of our knowledge, there are no reports on CCSCNCs for solar cells. In this paper, to explore the potential of CCSCNCs in solar cells, we propose a CCSCNC PSC with a structure involving a fluorine-doped tin oxide (FTO)/electron transport layer (ETL)/CCSCNCs/hole transport layer (HTL)/Au; then, by selecting a suitable hole transport material (HTM), electron transport material (ETM), thickness of the absorber layer, doping densities, defect density in the absorber, interface defect densities, and working temperature point, we predict that the CCSCNC solar cell with the FTO/TiO2/CCSCNCs/Cu2O/Au structure can attain a PCE of 23.07% at 300 K. In addition, we investigate the factors affecting the performance of the CCSCNCs solar cell for enhancing its performance, providing a guide for future experiments.

2. Device Structure and Simulation Parameters

2.1. Numerical Method

The numerical simulation software used in this work is solar cell capacitance simulator (SCAPS-1D 3.3.10), a one-dimension solar cell simulation software program developed by investigators at Ghent University [31]. Fundamentally, SCAPS-1D executes three sets of PV equations for the hole and electron carrier densities, respectively. These three equations are shown below:
d d x [ ε ( x ) d ψ d x ] = q [ p ( x ) n ( x ) + N d ( x ) N a ( x ) + p t ( x ) n t ( x ) ]
1 q d J n d x + R n ( x ) G ( x ) = 0
1 q d J p d x + R p ( x ) G ( x ) = 0
where x denotes the coordinate position; ψ and ε represent the electrostatic potential and the relative permittivity; n denotes the number of electrons; similarly, p denotes the number of holes; Nd represents the ionized donor concentration; Na represents the ionized acceptor density; nt is the number of trapped electrons; pt is the number of trapped holes; Jn represents the current density of electrons; Jp represents the current density of holes; G(x) is the photo-generated rate; the recombination rate of electrons is represented by Rn(x); and the recombination rate of holes is represented by Rp(x).
The optical absorption constant α of the CCSCNCs is calculated from the following equation of the optical absorption model Equation [31]:
α ( λ ) = ( A + B h v ) ( h v E g ) 1 2
where A and B are the model parameters, and Eg is the actual band gap of the material. The monochromatic photon in this work we set as 1018 s−1. The shunt resistance of the CCSCNCs solar cell is 4200 Ω∙cm2. As for the series resistance of the device, we set it as 1 Ω∙cm2. For working condition, we choose a standard AM 1.5G illumination and a continuous temperature 300 K [28].

2.2. Device Structure and Materials

The PSC structure of FTO/ETL/CCSCNCs/HTL/Au is schematically shown in Figure 1. In the structure, lead-free inorganic CCSCNCs were used as the absorber layer. The ETM and HTL will be carefully chosen for the purpose of high efficiency. The ETL is chosen from often-used materials in PSCs, e.g., TiO2, PCBM, ZnO, and IGZO. Similarly, the HTL is also chosen from often used materials in PSCs, e.g., P3HT, PEDOT:PSS, Spiro-OMETAD, Cu2O, and CuI. Table 1 and Table 2 include the basic parameters of the ETL, HTL and CCSCNCs, which were extracted from the previously published literature [19,20,21,25,26,32,33,34,35,36,37,38,39,40]. The bulk defect parameters in the absorber layer and the interface defect parameters at ETL/Absorber and Absorber/HTL interfaces are presented in Table 3.

3. Results and Discussion

We investigate the crucial factors affecting performance on the CCSCNCs solar cell and select the appropriate parameters to improve its performance. Firstly, we determine the suitable hole transport layer materials (HTMs) and electron transport materials (ETMs) for our CCSCNC PSC. Next, we optimize the absorber layer thickness to enhance the performance of the device. Finally, we investigate the effect of doping density in the transport layers, the effect of bulk defect in the absorber layer, the effect of interface defect densities at the CCSCNCs/ETL and CCSCNCs/ETL interfaces, and the effect of operating temperature. In addition, we select the appropriate values of the three parameters to enhance the performance of the device step by step. At each step of the optimization process, the other parameters are considered constants, and only the parameters to be optimized are varied. Each single optimization step is performed based on the completion of the previous optimizations.

3.1. HTMs Selection

In this selection, the HTL/CCSCNCs/TiO2/Au structure is used to select the most appropriate HTM. Figure 2 presents the J-V plots and PCE statistical graph of the device with different HTMs. The detailed performance parameters of the cells with different HTMs are shown in Table 4. The band alignment between CCSCNCs and HTMs is shown in Figure 3. The valence band offset (VBO) is defined by the following Equation (5), which denotes the difference between valence band (VB) level of HTL and that of the perovskite.
V B O = E V ( H T L ) E V ( A b s o r b e r L a y e r )
where Ev (HTL) denotes the VB level of HTL, Ev (Absorber Layer) denotes the VB level of the absorber layer. When the VB energy level of the CCSCNCs absorber layer is lower than that of HTL, the VBO becomes negative. It leads to the formation of an energy cliff between the absorber and HTL. The energy cliff can promote holes transporting HTL from the absorber layer. However, it is not the case that the larger the energy cliff is, the better it is. The difference between the absolute value of the band gap of the absorber layer and that of VBO represents the activation energy for carrier recombination, and a smaller activation energy means that the carriers are more likely to recombine [41]. Therefore, a large energy cliff will cause the CCSCNCs/Cu2O interface have a smaller carrier activation energy, resulting in enhanced recombination at the interface [42]. This can deteriorate the performance of the device. From Figure 3, it can be observed that the VBOs are −0.14 eV (P3HT), −0.24 eV(PEDOT:PSS), −0.23 eV(Spiro-OMETAD), +0.03 eV(Cu2O), −0.24 eV(CuI) respectively. The VBOs of all materials are negative, except for the VBO of Cu2O, which is positive. This indicates that all other materials form energy cliffs with the absorber layer and do not block hole transport. A positive VB for Cu2O means that an energy spike forms between the Cu2O and the absorber layer. The formation of this energy spike will hinder hole transport. In particular, small spikes have little effect on the hindrance of hole transport but can increase the activation energy of the interfacial recombination. This leads to a reduction of the interfacial recombination and thus to an improvement of the device performance. As shown in Figure 2, Cu2O used as HTM has a larger open-circuit voltage (Voc) and fill factor (FF) compared to other HTMs, because Cu2O has the smallest activation energy of the interfacial recombination. Therefore, the device used Cu2O, as the HTM possesses the highest PCE. Therefore, we obtain the optimal calculation result of the device by setting Cu2O as the HTM, which has a Voc of 1.03 V, a short circuit current density (Jsc) of 19.07 mA/cm2, an FF of 82%, and a PCE of 16.11%.

3.2. ETMs Selection

In this selection, the Cu2O/CCSCNCs/ETL/Au structure is used to select the most appropriate ETM. Figure 4 presents the J-V plots and PCE statistical graph of the device with different ETMs. The detailed performance parameters of the cells with different ETMs are shown in Table 5. The band alignment between CCSCNCs and ETMs is shown in Figure 5. The conduct band offset (CBO) is defined by the following Equation (6), which denotes the difference between conduct band (CB) level of ETL and that of the perovskite.
C B O = E C ( A b s o r b e r L a y e r ) E C ( E T L )
where Ec (Absorber Layer) denotes the CB level of the absorber layer, Ev (ETL) denotes the CB level of ETL. When the energy level of HTL is lower than that of the CCSCNCs absorber layer, the CBO becomes negative. The negative CBO is beneficial for electron transport to ETL from the absorber layer due to the formation of the energy cliffs between ETL and the absorber layer. Similarly, a small cliff is beneficial, but a large cliff enhances the interfacial recombination, resulting in the deterioration of performance [41,42]. From Figure 4, it can be observed that the TiO2 used as ETM has the highest PCE compared to other ETMs. Figure 5 shows that the CBOs are −0.16 eV(TiO2), −0.16 eV(PCBM), −0.36 eV(ZnO), and −0.42 eV(IGZO), respectively. Both TiO2 and PCBM have the smallest absolute CBO values, which are smaller than those of the other two materials. TiO2 used as ETM has a slightly higher PCE than PCBM because the dielectric constant of TiO2r = 9) is higher than that of PCBM(εr = 3.9), which is consistent with the previous literature [43]. Therefore, we select TiO2 as the most appropriate ETM. The device with TiO2 ETL has the Voc of 1.03 V, Jsc of 19.07 mA/cm2, FF of 82%, and PCE of 16.11%.

3.3. CCSCNC Thickness

Since the thicknesses of HTL and ETL have very small effects on the device performance, we only optimize the thickness of the absorber layer, which plays a decisive role in the device performance. The thickness of the CCSCNC layer is optimized within the range of 100 nm to 1000 nm. From Figure 6, it can be seen that the Voc increases with increasing absorber layer thickness and reaches a maximum value of 1.031 V at 250 nm, which should be attributed to the greater number of generated electrons and holes. However, above 250 nm, the Voc decreases due to the fact that excess absorption of photons may enhance the heat production in the device [44].
With increasing absorber layer thickness, the Jsc increases quickly, and above 400 nm, it continues to increase, but only slightly. This trend can be attributed to the fact that a thick absorber layer does not create more carriers [45]. It is noted that, as the absorber layer thickness increases, the FF decreases quickly, which is due to the increase in series resistance of the absorber layer. As a result, the PCE firstly increases, and then reaches a maximum value at 350 nm, after which the PCE gradually decreases.
The best performance occurs at a thickness of 350 nm, with Voc = 1.03 V, Jsc = 21.18 mA/cm2, FF = 77.69%, PCE = 16.94%.

3.4. Effect of the Doping Density in the Transport Layers

To understand the effect of doping in the HTL and ETL on device performance, we vary the doping density from 1016 cm−3 to 1020 cm−3.
J-V characteristics curves for different acceptor density in HTL are presented in Figure 7. The corresponding device performances are presented in Figure 8. In Figure 8, Voc, Jsc, FF increase with increasing acceptor density. The increase of the acceptor density in the HTL can increase the hole mobility and charge density, leading to a reduction in the resistivity of the HTL, resulting in increased Jsc and FF [46,47]. In addition, the increase of HTL doping density leads to the enhancement of the interfacial electric field between HTM and ETM, which increases the potential used to separate excitons and decreases the recombination rate [48]. Thus, a better extraction of electrons and holes from the absorber layer can be achieved, which improves the Voc. Therefore, the PCE increases from 15.86% to 17.15% when the acceptor density increases from 1016 cm−3 to 1020 cm−3. Therefore, we select an acceptor density NA of 1020 cm−3, and the performance parameters of the device with this acceptor density are an FF of 78.52%, Jsc of 21.20 mA/cm2, Voc of 1.03 V, and PCE of 17.15%.
Similarly, J-V characteristics curves for different donor density in ETL are presented in Figure 9. The corresponding device performances are presented in Figure 10. Figure 10 shows that Voc, Jsc, and FF increase with the increase of donor density. Similar to the doping density of HTL discussed above, the interfacial electric field between ETL and HTL is enhanced as the doping density of ETL increases, which contributes to the separation of excitons and reduces the recombination [48]. The PCE increases from 15.88% to 18.13% with an increase in donor density from 1016 cm−3 to 1020 cm−3. Therefore, we select a donor density ND of 1020 cm−3, and the performance parameters of the device with this donor density are an FF of 78.52%, Jsc of 21.21 mA/cm2, Voc of 1.09 V, and PCE of 18.13%.

3.5. Effect of the Bulk Defect Density

To analyze the effect of the bulk defect density at the interface of the absorber/transport layer on device performance, we vary the bulk defect density in the CCSCNC absorber layer from 1011 cm−3 to 1017 cm−3. The J-V characteristics curves for different bulk defect density in the absorber layer are presented in Figure 11. The corresponding device performance is presented in Figure 12. From Figure 12, it can be seen that Voc, Jsc, and FF decrease with increasing bulk defect density due to the fact that the generated electrons and holes are more easily captured by bulk defects [43]. The PCE falls from 21.3% to 4.74% when the bulk defect density increases from 1011 cm−3 to 1017 cm−3. When the bulk defect density is below 1012 cm−3, the effect of the defect density on the PCE becomes weak. Defects are inevitable in the actual perovskite absorber layer, so the device has relatively good performance at a defect density of less than 1012 cm−3. We select an optimized bulk defect density of 1012 cm−3, with an obtained Voc of 1.16 V, Jsc of 21.35 mA/cm2, FF of 86.33%, and PCE of 21.3%.

3.6. Effect of Interface Defect Density

For the interface between the perovskite absorber layer and the transport layer, it has been shown that the interface defects density increases under light, oxygen, humidity, and high temperature, thus degrading the performance of the device. Therefore, it is significant to study the effect of interface defect density on device performance [49].
We vary the interface defect density at Cu2O/CCSCNCs interface and CCSCNCs/TiO2 interface from 109 cm−3 to 1021 cm−3. Figure 13 presents the J-V curves for different interface defect densities in Cu2O/CCSCNCs interface layer. Figure 14 presents the corresponding device performance. It shows that Voc, Jsc, and FF decrease with the Cu2O/CCSCNCs interface defect density. This trend is attributed to the fact that the higher defect density at Cu2O/CCSCNCs interface brings more traps and recombination centers, which results in deteriorating performance of the cells [43]. Therefore, the PCE decreases from 21.20% to 20.81% with the Cu2O/CCSCNCs interface defect density increasing from 1019 cm−3 to 1021 cm−3. It is obvious that when the defect density at Cu2O/CCSCNCs interface layer is below 1013 cm−3, the effect of the defect density on the PCE becomes weak. Considering that defects are inevitable in the actual interface, relatively good performance can be obtained when the defect density is less than 1013 cm−3. Thus, we choose 1013 cm−3 as the defect density at Cu2O/CCSCNCs interface layer, with an obtained Voc of 1.16 V, a Jsc of 21.35 mA/cm2, an FF of 86.33%, and a PCE of 21.3%.
Figure 15 presents the J-V curves for different interface defect density in the CCSCNCs/TiO2 interface layer. Figure 16 presents the corresponding device performance. In Figure 16, similar to the CCSCNCs/TiO2 interface, Voc, Jsc, and FF decrease with the CCSCNCs/TiO2 interface defect density due to the higher defect density at the CCSCNCs/TiO2 interface resulting in more traps and recombination centers. Therefore, the PCE decreases from 23.07% to 15.18% with an increase in CCSCNCs/TiO2 interface defect density from 109 cm−3 to 1021 cm−3. It is noted that the efficiency reduction due to the increase of CCSCNCs/TiO2 interface defect density is much larger than that due to the increase of Cu2O/CCSCNCs interface defect density. Obviously, the defect density at the CCSCNCs/TiO2 interface has a remarkable influence on the device performance. This is because the number of electron–hole pairs generated at CCSCNCs/TiO2 is 10 times higher than that at Cu2O/CCSCNCs under light illumination [44]. The higher excess carrier density present at the CCSCNCs/TiO2 interface leads to a higher recombination rate. The difference between the two interfaces we obtained is consistent with the previous literature [50,51]. Considering that defects are inevitable in the actual interface, relatively good performance can be obtained when the defect density is less than 109 cm−3. Therefore, when the interface defect density at the CCSCNCs/TiO2 interface is 109 cm−3, the maximum PCE is 23.07%, while FF = 83.02%, Jsc = 21.35 mA/cm2, and Voc = 1.3 V.

3.7. Effect of Operating Temperature

The actual operating temperature of PSC typically exceeds 80 °C (353 K), and the performance of PSC is highly dependent on the operating temperature [52]. We investigate the effect of operating temperature on the device performance by varying the temperature from 300 K to 500 K. Figure 17 shows that the device performance continues to deteriorate with increasing temperature. As the temperature increases from 300 K to 500 K, the PCE of the device decreases from 23.07% to 16.12%. This phenomenon can be explained by the following Equation (7):
d V o c d T = ( V o c E g q ) T
where T denotes the working temperature, q represents the elementary charge, Eg denotes the band gap. As the temperature increases, the Voc decreases, resulting in a lower PCE of the device [53]. In addition, as the temperature increases, the device defect density in the device increases, and the carrier mobility decrease, which deteriorates the device performance [54]. Therefore, we obtain an optimal PCE of 23.07% on the device at 300 K.

3.8. Performance with the Optimized Device Structure

The final optimal device structure is FTO/TiO2/CCSCNCs/Cu2O/Au, and the thickness of the CCSCNC absorber layer is 350 nm, as shown in Figure 18. The J-V curve for the optimal CCSCNC cell structure is presented in Figure 19a, where the acceptor density in HTL and the donor density in ETL are both 1020 cm−3, and the bulk defect density in the CCSCNC absorber layer, the interface defect density at Cu2O/CCSCNCs interface, and the interface defect density at CCSCNCs/TiO2 interface are 1012 cm−3, 1013 cm−3, and 109 cm−3, respectively. The performance of the optimized device is predicted to have a PCE of 23.07%, Voc of 1.3 V, Jsc = 21.35 mA/cm2 and FF = 83.02%.
The external quantum efficiency (EQE) was also calculated. The EQE takes optical performance of the solar cell along with the ratio of charge generation with respect to incident light photons. From Figure 19b, it can be seen that 80~90% quantum efficiency is obtained in the wavelength range of 360 nm–720 nm.
Table 6 provides a performance comparison among various works on lead-free PSC. It shows that different lead-free perovskite absorbers produce different PCEs, e.g., from 13.57% to 27.43%. The corresponding Voc, Jsc, FF values are in the range of 0.8 V~1.9 V, 19.88 mA/cm2~40.05 mA/cm2, and 35.95%~87.79%, respectively. Although the values for Voc, Jsc, FF we obtained are not the highest among those lead-free PSCs, the PCE is very high. Our simulation results show that the PSC with structure of FTO/TiO2/CCSCNCs/Cu2O/Au has an exciting PCE of 23.07%. It indicates that the CCSCNCs are very suitable for the absorber layer in PSC.

4. Conclusions

In this work, we numerically explored the performance of the Cs4CuSb2Cl12 nanocrystal solar cell using SCAPS-1D. By selecting the appropriate hole transport material, electron transport material, thickness of the absorber layer, doping densities, defect density in the absorber, interface defect densities, and working temperature point, we predicted that the CCSCNCs solar cell with the FTO/TiO2/CCSCNCs/Cu2O/Au structure could attain a PCE of 23.07% at 300 K. Very high electrical parameters were obtained, with a Jsc of 21.35 mA/cm2, a Voc of 1.30 V, an FF of 83.02%, and an external quantum efficiency of 80~90% in the range of 360~720 nm. These exciting results suggest that CCSCNCs could play a momentous role as an absorbing perovskite in achieving a highly efficient lead-free inorganic perovskite solar cell technology. In addition, we investigated the factors affecting the performance of CCSCNCs solar cells in order to enhance their performance, thus providing a guide for future experiments.

Author Contributions

Original draft preparation, Y.H.; conceptualization, methodology, Y.H. and L.X.; revising draft preparation, C.Y.; reviewing, X.G.; methodology, figure, Y.H.; concept review, editing, Y.H. and L.X.; project administration, editing, C.Y.; funding acquisition, supervision, S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by Natural Science Foundation of Sichuan Province (2019YJ0172).

Data Availability Statement

The data will be made available upon reasonable request to the corresponding author.

Acknowledgments

The authors gratefully acknowledge Marc Bargeman, University of Gent, Belgium, for providing the SCAPS simulation software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  2. National Renewable Energy Laboratory. Perovskite Efficiency Chart; National Renewable Energy Laboratory: Golden, CO, USA, 2020. [Google Scholar]
  3. Zhang, T.; Li, H.; Ban, H.; Sun, Q.; Shen, Y.; Wang, M.K. Efficient CsSnI 3 -based inorganic perovskite solar cells based on a mesoscopic metal oxide framework via incorporating a donor element. J. Mater. Chem. A 2020, 8, 4118–4124. [Google Scholar] [CrossRef]
  4. Giustino, F.; Snaith, H.J. Toward lead-free perovskite solar cells. ACS Energy Lett. 2016, 1, 1233–1240. [Google Scholar] [CrossRef] [Green Version]
  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] [PubMed] [Green Version]
  6. Yang, W.S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S. High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science 2015, 348, 1234–1237. [Google Scholar] [CrossRef]
  7. Yin, W.J.; Yang, J.H.; Kang, J.; Yan, Y.; Wei, S.H. Halide perovskite materials for solar cells: A theoretical review. J. Mater. Chem. A 2015, 3, 8926–8942. [Google Scholar] [CrossRef]
  8. Fabini, D.H.; Laurita, G.; Bechtel, J.S.; Stoumpos, C.C.; Evans, H.A.; Kontos, A.G.; Raptis, Y.S.; Falaras, P.; van der Ven, A.; Kanatzidis, M.G.; et al. Dynamic stereochemical activity of the Sn2+ lone pair in perovskite CsSnBr3. J. Am. Chem. Soc. 2016, 138, 11820–11832. [Google Scholar] [CrossRef] [Green Version]
  9. Shi, Z.; Guo, J.; Chen, Y.; Li, Q.; Pan, Y.; Zhang, H.; Xia, Y.; Huang, W. Lead-free organic–inorganic hybrid perovskites for photovoltaic applications: Recent advances and perspectives. Adv. Mater. 2017, 29, 1605005. [Google Scholar] [CrossRef]
  10. Krishnamoorthy, T.; Ding, H.; Yan, C.; Leong, W.L.; Baikie, T.; Zhang, Z.; Sherburne, M.; Li, S.; Asta, M.; Mathews, N.; et al. Lead-free germanium iodide perovskite materials for photovoltaic application. J. Mater. Chem. A 2015, 3, 23829–23832. [Google Scholar] [CrossRef]
  11. Pradhan, B.; Kumar, G.S.; Sain, S.; Dalui, A.; Ghorai, U.K.; Pradhan, S.K.; Acharya, S. Size tunable cesium antimony chloride perovskite nanowires and nanorods. Chem. Mater. 2018, 30, 2135–2142. [Google Scholar] [CrossRef]
  12. Pal, J.; Manna, S.; Mondal, A.; Das, S.; Adarsh, K.V.; Nag, A. Colloidal Synthesis and Photophysics of M3Sb2I9 (M = Cs and Rb) Nanocrystals: Lead-Free Perovskites. Angew. Chem. Int. Ed. 2017, 56, 14187–14191. [Google Scholar] [CrossRef]
  13. Yang, B.; Chen, J.; Hong, F.; Mao, X.; Zheng, K.; Yang, S.; Li, Y.; Pullerits, T.; Deng, W.; Han, K. Lead-free, air-stable all-inorganic cesium bismuth halide perovskite nanocrystals. Angew. Chem. Int. Ed. 2017, 56, 12471–12475. [Google Scholar] [CrossRef]
  14. Nelson, R.D.; Santra, K.; Wang, Y.; Hadi, A.; Petrich, J.W.; Panthani, M.G. Synthesis and optical properties of ordered-vacancy perovskite cesium bismuth halide nanocrystals. Chem. Commun. 2018, 54, 3640–3643. [Google Scholar] [CrossRef] [Green Version]
  15. Pal, J.; Bhunia, A.; Chakraborty, S.; Manna, S.; Das, S.; Dewan, A.; Datta, S.; Nag, A. Synthesis and Optical Properties of Colloidal M3Bi2I9 (M = Cs, Rb) Perovskite Nanocrystals. J. Phys. Chem. C. 2018, 122, 10643–10649. [Google Scholar] [CrossRef]
  16. Pan, W.; Wu, H.; Luo, J.; Deng, Z.; Ge, C.; Chen, C.; Jiang, X.; Yin, W.-J.; Niu, G.; Zhu, L.; et al. Cs2AgBiBr6 single-crystal X-ray detectors with a low detection limit. Nat. Photonics 2017, 11, 726–732. [Google Scholar] [CrossRef]
  17. Zhou, L.; Xu, Y.-F.; Chen, B.-X.; Kuang, D.-B.; Su, C.-Y. Synthesis and Photocatalytic Application of Stable Lead-Free Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14, 1703762. [Google Scholar] [CrossRef]
  18. McClure, E.T.; Ball, M.R.; Windl, W.; Woodward, P.M. Cs2AgBiX6 (X = Br, Cl): New Visible Light Absorbing, Lead-Free Halide Perovskite Semiconductors. Chem. Mater. 2016, 28, 1348–1354. [Google Scholar] [CrossRef]
  19. Vargas, B.; Ramos, E.; Perez-Gutierrez, E.; Alonso, J.C.; Solis Ibarra, D. A Direct Bandgap Copper–Antimony Halide Perovskite. J. Am. Chem. Soc. 2017, 139, 9116–9119. [Google Scholar] [CrossRef]
  20. Singhal, N.; Chakraborty, R.; Ghosh, P.; Nag, A. Low-Bandgap Cs4CuSb2Cl12 Layered Double Perovskite: Synthesis, Reversible Thermal Changes, and Magnetic Interaction. Chem. Asian J. 2018, 13, 2085–2092. [Google Scholar] [CrossRef]
  21. Tang, G.; Xiao, Z.; Hosono, H.; Kamiya, T.; Fang, D.; Hong, J. Layered Halide Double Perovskites Cs3+nM(II)nSb2X9+3n (M = Sn, Ge) for Photovoltaic Applications. J. Phys. Chem. Lett. 2017, 9, 43–48. [Google Scholar] [CrossRef]
  22. Samadi, M.; Sarikhani, N.; Zirak, M.; Zhang, H.; Zhang, H.L.; Moshfegh, A.Z. Group 6 transition metal dichalcogenide nanomaterials: Synthesis, applications and future perspectives. Nanoscale Horiz. 2018, 3, 90–204. [Google Scholar] [CrossRef] [PubMed]
  23. Bandurin, D.A.; Tyurnina, A.V.; Yu, G.L.; Mishchenko, A.; Zolyomi, V.; Morozov, S.V.; Kumar, R.K.; Gorbachev, R.V.; Kudrynskyi, Z.R.; Pezzini, S.; et al. High electron mobility, quantum Hall effect and anomalous optical response in atomically thin InSe. Nat. Nanotechnol. 2017, 12, 223–227. [Google Scholar] [CrossRef]
  24. Akkerman, Q.A.; Motti, S.G.; Srimath Kandada, A.R.; Mosconi, E.; D’Innocenzo, V.; Bertoni, G.; Marras, S.; Kamino, B.A.; Miranda, L.; de Angelis, F.; et al. Solution Synthesis Approach to Colloidal Cesium Lead Halide Perovskite Nanoplatelets with Monolayer-Level Thickness Control. J. Am. Chem. Soc. 2016, 138, 1010–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wang, X.D.; Miao, N.H.; Liao, J.F.; Li, W.Q.; Xie, Y.; Chen, J.; Sun, Z.M.; Chen, H.Y.; Kuang, D.B. The top-down synthesis of single-layered Cs4CuSb2Cl12 halide perovskite nanocrystals for photoelectrochemical application. Nanoscale 2019, 11, 5180–5187. [Google Scholar] [CrossRef]
  26. Cai, T.; Shi, W.; Hwang, S.; Kobbekaduwa, K.; Nagaoka, Y.; Yang, H.; Hills-Kimball, K.; Zhu, H.; Wang, J.; Wang, Z.; et al. Lead-free Cs4CuSb2Cl12 layered double perovskite nanocrystals. J. Am. Chem. Soc. 2020, 142, 11927–11936. [Google Scholar] [CrossRef]
  27. Ashtiha, P.P.; Joshi, M.; Verma, D.; Jadhav, S.; Roy Choudhury, A.; Jana, D. Layered Cs4CuSb2Cl12 Nanocrystals for Sunlight-Driven Photocatalytic Degradation of Pollutants. ACS Appl. Nano Mater. 2021, 4, 1305–1313. [Google Scholar] [CrossRef]
  28. Que, M.; Zhu, L.; Guo, Y.; Que, W.; Yun, S. Toward perovskite nanocrystalline solar cells: Progress and potential. J. Mater. Chem. C 2020, 8, 5321–5334. [Google Scholar] [CrossRef]
  29. Zhou, F.; Li, Z.; Chen, H.; Wang, Q.; Ding, L.; Jin, Z. Application of perovskite nanocrystals (NCs)/quantum dots (QDs) in solar cells. Nano Energy 2020, 73, 104757. [Google Scholar] [CrossRef]
  30. Hao, M.; Bai, Y.; Zeiske, S.; Ren, L.; Liu, J.; Yuan, J.; Zarrabi, N.; Cheng, N.; Ghasemi, M.; Chen, P.; et al. Ligand-assisted cation-exchange engineering for high-efficiency colloidal Cs1−xFAxPbI3 quantum dot solar cells with reduced phase segregation. Nat. Energy 2020, 5, 79–88. [Google Scholar] [CrossRef]
  31. Burgelman, M.; Nollet, P.; Degrave, S. Modelling polycrystalline semiconductor solar cells. Thin Solid Films 2000, 361, 527–532. [Google Scholar] [CrossRef]
  32. Trinh, M.T.; Wu, X.; Niesner, D.; Zhu, X.-Y. Many-body interactions in photo-excited lead iodide perovskite. J. Mater. Chem. A 2015, 3, 9285–9290. [Google Scholar] [CrossRef]
  33. Ahmed, S.; Jannat, F.; Khan, M.A.K.; Alim, M.A. Numerical development of eco-friendly Cs2TiBr6 based perovskite solar cell with all-inorganic charge transport materials via SCAPS-1D. Optik 2021, 225, 165765. [Google Scholar] [CrossRef]
  34. Jani, M.R.; Islam, M.T.; Amin, S.M.A.; Sami, M.S.U.; Shorowordi, K.M.; Hossain, M.I.; Chowdhury, S.; Nishat, S.S.; Ahmed, S. Exploring solar cell performance of inorganic Cs2TiBr6 halide double perovskite: A numerical study. Superlattice Microst. 2020, 146, 106652. [Google Scholar] [CrossRef]
  35. Chowdhury, M.S.; Shahahmadi, S.A.; Chelvanathan, P.; Tiong, S.K.; Amin, N.; Techato, K.; Nuthammachot, N.; Chowdhury, T.; Suklueng, M. Effect of deep-level defect density of the absorber layer and n/i interface in perovskite solar cells by SCAPS-1D. Results Phys. 2020, 16, 102839. [Google Scholar] [CrossRef]
  36. Karimi, E.; Ghorashi, S.M.B. Investigation of the influence of different hole-transporting materials on the performance of perovskite solar cells. Optik 2017, 130, 650–658. [Google Scholar] [CrossRef]
  37. Jayan, K.D.; Sebastian, V. Comprehensive device modelling and performance analysis of MASnI3 based perovskite solar cells with diverse ETM, HTM and back metal contacts. Solar Energy 2021, 217, 40–48. [Google Scholar] [CrossRef]
  38. Raj, A.; Anshul, A.; Tuli, V.; Singh, P.K.; Singh, R.C.; Kumar, M. Effect of appropriate ETL on quantum efficiency of double perovskite La2NiMnO6 based solar cell device via SCAPS simulation. Mater. Today Proc. 2021, in press. [Google Scholar] [CrossRef]
  39. Lakhdar, N.; Hima, A. Electron transport material effect on performance of perovskite solar cells based on CH3NH3GeI3. Opt. Mater. 2020, 99, 109517. [Google Scholar] [CrossRef]
  40. Islam, S.; Sobayel, K.; Al-Kahtani, A.; Islam, M.A.; Muhammad, G.; Amin, N.; Shahiduzzaman, M.; Akhtaruzzaman, M. Defect Study and Modelling of SnX3-Based Perovskite Solar Cells with SCAPS-1D. Nanomaterials 2021, 11, 1218. [Google Scholar] [CrossRef]
  41. Tanaka, K.; Minemoto, T.; Takakura, H. Analysis of heterointerface recombination by Zn1− xMgxO for window layer of Cu (In, Ga) Se2 solar cells. Solar Energy 2009, 83, 477–479. [Google Scholar] [CrossRef]
  42. Turcu, M.; Rau, U. Fermi level pinning at CdS/Cu(In,Ga)(Se,S)2 interfaces: Effect of chalcopyrite alloy composition. J. Phys. Chem. Solids 2003, 64, 1591–1595. [Google Scholar] [CrossRef]
  43. Gan, Y.; Bi, X.; Liu, Y.; Qin, B.; Li, Q.; Jiang, Q.; Mo, P. Numerical investigation energy conversion performance of tin-based perovskite solar cells using cell capacitance simulator. Energies 2020, 13, 5907. [Google Scholar] [CrossRef]
  44. Tan, K.; Lin, P.; Wang, G.; Liu, Y.; Xu, Z.; Lin, Y. Controllable design of solid-state perovskite solar cells by SCAPS device simulation. Solid-State Electron. 2016, 126, 75–80. [Google Scholar] [CrossRef]
  45. 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]
  46. Abate, A.; Leijtens, T.; Pathak, S.; Teuscher, J.; Avolio, R.; Errico, M.E.; Kirkpatrik, J.; Ball, J.M.; Docampo, P.; McPhersonc, I.; et al. Lithium salts as “redox active” p-type dopants for organic semiconductors and their impact in solid-state dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2013, 15, 2572–2579. [Google Scholar] [CrossRef] [PubMed]
  47. Leijtens, T.; Lim, J.; Teuscher, J.; Park, T.; Snaith, H.J. Charge density dependent mobility of organic hole-transporters and mesoporous TiO2 determined by transient mobility spectroscopy: Implications to dye-sensitized and organic solar cells. Adv. Mater. 2013, 25, 3227–3233. [Google Scholar] [CrossRef]
  48. Trukhanov, V.A.; Bruevich, V.V.; Paraschuk, D.Y. Effect of doping on performance of organic solar cells. Phys. Rev. B 2011, 84, 205–318. [Google Scholar] [CrossRef] [Green Version]
  49. Ghadiri, M.; Kang, A.K.; Gorji, N.E. XRD characterization of graphene-contacted perovskite solar cells: Moisture degradation and dark-resting recovery. Superlattices Microstruct. 2020, 146, 106677. [Google Scholar] [CrossRef]
  50. Huang, L.; Sun, X.; Li, C.; Xu, R.; Xu, J.; Du, Y.; Wu, Y.; Ni, J.; Cai, H.; Li, J.; et al. Electron transport layer-free planar perovskite solar cells: Further performance enhancement perspective from device simulation. Sol. Energy Mater. Sol. Cells 2016, 157, 1038–1047. [Google Scholar] [CrossRef]
  51. Lin, L.; Li, P.; Jiang, L.; Kang, Z.; Yan, Q.; Xiong, H.; Lien, S.; Zhang, P.; Qiu, Y. Boosting efficiency up to 25% for HTL-free carbon-based perovskite solar cells by gradient doping using SCAPS simulation. Sol. Energy 2021, 215, 328–334. [Google Scholar] [CrossRef]
  52. Slami, A.; Bouchaour, M.; Merad, L. Comparative study of modelling of Perovskite solar cell with different HTM layers. Int. J. Mater. 2020, 7, 2313–10555. [Google Scholar] [CrossRef]
  53. Devi, N.; Parrey, K.A.; Aziz, A.; Datta, S. Numerical simulations of perovskite thin-film solar cells using a CdS hole blocking layer. J. Vac. Sci. Technol. B 2018, 36, 04G105. [Google Scholar] [CrossRef]
  54. Behrouznejad, F.; Shahbazi, S.; Taghavinia, N.; Wu, H.-P.; Wei-Guang Diau, E. A study on utilizing different metals as the back contact of CH 3 NH 3 PbI 3 perovskite solar cells. J. Mater. Chem. A 2016, 4, 13488–13498. [Google Scholar] [CrossRef]
  55. Chakraborty, K.; Choudhury, M.G.; Paul, S. Study of Physical, Optical, and Electrical Properties of Cesium Titanium (IV)-Based Single Halide Perovskite Solar Cell. IEEE J. Photovolt. 2021, 11, 386–390. [Google Scholar] [CrossRef]
  56. Singh, A.K.; Srivastava, S.; Mahapatra, A.; Baral, J.K.; Pradhan, B. Performance optimization of lead free-MASnI3 based solar cell with 27% efficiency by numerical simulation. Opt. Mater. 2021, 117, 111193. [Google Scholar] [CrossRef]
  57. 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]
  58. Khattak, Y.H.; Baig, F.; Shuja, A.; Beg, S.; Soucase, B.M. Numerical analysis guidelines for the design of efficient novel nip structures for perovskite solar cell. Sol. Energy 2020, 207, 579–591. [Google Scholar] [CrossRef]
  59. Coulibaly, A.B.; Oyedele, S.O.; Kre, N.A.; Aka, B. Comparative study of lead-free perovskite solar cells using different hole transporter materials. Model. Numer. Simul. Mater. Sci. 2019, 9, 97–107. [Google Scholar] [CrossRef] [Green Version]
  60. Ahmad, O.; Rashid, A.; Ahmed, M.W.; Nasir, M.F.; Qasim, I. Performance evaluation of Au/p-CdTe/Cs2TiI6/n-TiO2/ITO solar cell using SCAPS-1D. Opt. Mater. 2021, 117, 111105. [Google Scholar] [CrossRef]
Figure 1. Schematic of the CCSCNCs-based perovskite solar cell.
Figure 1. Schematic of the CCSCNCs-based perovskite solar cell.
Nanomaterials 11 02321 g001
Figure 2. (a) Comparison of J-V characteristic curves for different HTMs as HTL; (b) PCE for different HTMs as HTL.
Figure 2. (a) Comparison of J-V characteristic curves for different HTMs as HTL; (b) PCE for different HTMs as HTL.
Nanomaterials 11 02321 g002
Figure 3. Band alignment between HTL materials and CCSCNCs.
Figure 3. Band alignment between HTL materials and CCSCNCs.
Nanomaterials 11 02321 g003
Figure 4. (a) Comparison of the J-V characteristic curves for different HTMs as HTL; (b) PCE for different ETMs as ETL.
Figure 4. (a) Comparison of the J-V characteristic curves for different HTMs as HTL; (b) PCE for different ETMs as ETL.
Nanomaterials 11 02321 g004
Figure 5. Band alignment between ETL materials and CCSCNCs.
Figure 5. Band alignment between ETL materials and CCSCNCs.
Nanomaterials 11 02321 g005
Figure 6. (a) Change in Voc and Jsc against CCSCNC absorber layer thickness variation; (b) change in PCE and FF against CCSCNC absorber layer thickness variation.
Figure 6. (a) Change in Voc and Jsc against CCSCNC absorber layer thickness variation; (b) change in PCE and FF against CCSCNC absorber layer thickness variation.
Nanomaterials 11 02321 g006
Figure 7. Comparison of J-V characteristic curves for different acceptor density (NA).
Figure 7. Comparison of J-V characteristic curves for different acceptor density (NA).
Nanomaterials 11 02321 g007
Figure 8. (a) Change in Voc and Jsc against acceptor density (NA) variation; (b) change in PCE and FF against HTL doping density (NA) variation.
Figure 8. (a) Change in Voc and Jsc against acceptor density (NA) variation; (b) change in PCE and FF against HTL doping density (NA) variation.
Nanomaterials 11 02321 g008
Figure 9. Comparison of J-V characteristic curves for different donor density (ND).
Figure 9. Comparison of J-V characteristic curves for different donor density (ND).
Nanomaterials 11 02321 g009
Figure 10. (a) Change in Voc and Jsc against ETL donor density (ND) variation; (b) change in PCE and FF against ETL doping density (ND) variation.
Figure 10. (a) Change in Voc and Jsc against ETL donor density (ND) variation; (b) change in PCE and FF against ETL doping density (ND) variation.
Nanomaterials 11 02321 g010
Figure 11. Comparison of J-V characteristic curves for different bulk defect density.
Figure 11. Comparison of J-V characteristic curves for different bulk defect density.
Nanomaterials 11 02321 g011
Figure 12. (a) Change in Voc and Jsc against absorber layer bulk defect density variation; (b) change in PCE and FF against absorber layer bulk defect density variation.
Figure 12. (a) Change in Voc and Jsc against absorber layer bulk defect density variation; (b) change in PCE and FF against absorber layer bulk defect density variation.
Nanomaterials 11 02321 g012
Figure 13. Comparison of J-V characteristic curves for different Cu2O/CCSCNCs interface layer defect density.
Figure 13. Comparison of J-V characteristic curves for different Cu2O/CCSCNCs interface layer defect density.
Nanomaterials 11 02321 g013
Figure 14. (a) Change in Voc and Jsc against Cu2O/CCSCNCs interface defect density variation; (b) change in PCE and FF against Cu2O/CCSCNCs interface defect density variation.
Figure 14. (a) Change in Voc and Jsc against Cu2O/CCSCNCs interface defect density variation; (b) change in PCE and FF against Cu2O/CCSCNCs interface defect density variation.
Nanomaterials 11 02321 g014
Figure 15. Comparison of J-V characteristic curves for different CCSCNCs/TiO2 interface layer defect density.
Figure 15. Comparison of J-V characteristic curves for different CCSCNCs/TiO2 interface layer defect density.
Nanomaterials 11 02321 g015
Figure 16. (a) Change in Voc and Jsc against CCSCNCs/TiO2 interface defect density variation; (b) change in PCE and FF against CCSCNCs/TiO2 interface defect density variation.
Figure 16. (a) Change in Voc and Jsc against CCSCNCs/TiO2 interface defect density variation; (b) change in PCE and FF against CCSCNCs/TiO2 interface defect density variation.
Nanomaterials 11 02321 g016
Figure 17. J-V characteristic curves of the device for different operating temperatures.
Figure 17. J-V characteristic curves of the device for different operating temperatures.
Nanomaterials 11 02321 g017
Figure 18. Schematic of the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure.
Figure 18. Schematic of the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure.
Nanomaterials 11 02321 g018
Figure 19. (a) J-V curve for the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure; (b) EQE curve for the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure.
Figure 19. (a) J-V curve for the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure; (b) EQE curve for the optimized device with FTO/TiO2/CCSCNCs/Cu2O/Au structure.
Nanomaterials 11 02321 g019
Table 1. Input parameters for CCSCNCs absorber layer and HTL of the PSC structure.
Table 1. Input parameters for CCSCNCs absorber layer and HTL of the PSC structure.
ParametersCCSCNCsP3HTPEDOT:PSSSpiro-OMETADCu2OCuI
Thickness, d (nm)2003030303030
Band gap, Eg (eV)1.622.23.062.173.1
Electron affinity, χ (eV)3.743.22.92.053.22.1
Permittivity, εr103337.16.5
Effective density states at CB, NC (cm−3)4.5 × 10181 × 10202.2 × 10152.8 × 10192.5 × 10182.2 × 1019
Effective density states at VB, NV (cm−3)1.6 × 10181 × 10201.8 × 10181 × 10191.8 × 10191.8 × 1019
Electron mobility, μn (cm−2/V·s)2.50.0001101× 10−4200100
Hole mobility, μp (cm−2/V·s)2.50.0001102× 10−48043.9
Density of n-type doping, ND (cm−3)1 × 101300000
Density of p-type doping, NA (cm−3)1 × 10131 × 10163.17 × 10141 × 10189.1 × 10211 × 1018
Defect density, Nt (cm−3)1 × 10151 × 10151 × 10141 × 10141 × 10141 × 1014
electron thermal velocity (cm/s)1 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
hole thermal velocity (cm/s)1 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
Capture cross-section electrons (cm2)1 × 10−141 × 10−151 × 10−151 × 10−151 × 10−151 × 10−15
Capture cross-section holes (cm2)1 × 10−141 × 10−151 × 10−151 × 10−151 × 10−151 × 10−15
Reference[19,20,21,25,26,32,33,34,37,40][33][35][36][34][38]
Table 2. Input parameters for ETL of the PSC structure.
Table 2. Input parameters for ETL of the PSC structure.
ParametersTiO2PCBMZnOIGZO
Thickness, d (nm)30303030
Band gap, Eg (eV)3.223.33.05
Electron affinity, χ (eV)4.13.94.14.16
Permittivity, εr93.9910
Effective density states at CB NC (cm−3)2.2 × 10182.5 × 10214 × 10185 × 1018
Effective density states at VB NV (cm−3)1 × 10192.5 × 10211 × 10191 × 1018
Electron mobility, μn (cm−2/V·s)200.210015
Hole mobility, μp (cm−2/V·s)100.2250.2
Density of n-type doping, ND (cm−3)1 × 10182.93 × 10171 × 10181 × 1017
Density of p-type doping, NA (cm−3)001 × 1050
Defect density, Nt (cm−3)1 × 10151 × 10152 × 10171 × 1015
electron thermal velocity (cm/s)1 × 1071 × 1071 × 1071 × 107
hole thermal velocity (cm/s)1 × 1071 × 1071 × 1071 × 107
Capture cross-section electrons (cm2)2 × 10141 × 10151 × 10152 × 1014
Capture cross-section holes (cm2)2 × 10141 × 10151 × 10152 × 1014
Reference[33][38][39][37]
Table 3. Input parameters of defect inside the absorber and interface defect layers.
Table 3. Input parameters of defect inside the absorber and interface defect layers.
ParametersETL/AbsorberAbsorber/HTLCCSCNCs
Defect typeNeutralNeutralNeutral
Capture cross-section for electrons (cm2)1 × 10−191 × 10−181 × 10−15
Capture cross-section for holes (cm2)1 × 10−181 × 10−191 × 10−15
Energetic distributionGaussianGaussianGaussian
Energy level with respect to Ev0.60.60.6
Characteristic energy (eV)0.10.10.1
Total density (cm−3)1 × 10131 × 10131 × 1015
Reference[33][33][33,40]
Table 4. Performance parameters of different HTMs as HTL.
Table 4. Performance parameters of different HTMs as HTL.
Hole Transport MaterialVoc (V)Jsc (mA/cm2)FF (%)PCE (%)
P3HT1.0119.0770.4713.54
PEDOT:PSS1.0021.1467.4414.21
Spiro-OMETAD1.0219.2176.5415.00
Cu2O1.0319.0782.0116.11
CuI1.0319.2279.6615.73
Table 5. Performance parameters of different ETMs as ETL.
Table 5. Performance parameters of different ETMs as ETL.
Electron Transport MaterialVoc (V)Jsc (mA/cm2)FF (%)PCE (%)
TiO21.03 19.07 82.01 16.11
PCBM1.02 19.25 81.74 16.01
ZnO1.02 19.07 81.97 15.87
IGZO0.90 19.06 80.35 13.80
Table 6. Performance analysis of various lead-free PSC.
Table 6. Performance analysis of various lead-free PSC.
Electron Transport MaterialVoc (V)Jsc (mA/cm2)FF (%)PCE (%)FormRef
Cs2TiBr61.919.8835.9513.57simulation[55]
Cs2TiI61.7422.744116.31simulation[55]
MASnI31.20325.9787.7927.43simulation[56]
MASnI30.9632.4876.423.86simulation[43]
FASnI31.8131.233.7219.08simulation[57]
MASnBr30.831.8884.8921.66simulation[55]
MAPbI3---25.15simulation[51]
MAPbI3 1.0524.4886.3126.96simulation[58]
MASnI30.8440.0570.8223.76simulation[59]
Cs2TiI61.3925.0843.1715.06simulation[60]
Cs4CuSb2Cl12 nanocrystals1.3021.3583.0223.07simulationThis 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

He, Y.; Xu, L.; Yang, C.; Guo, X.; Li, S. Design and Numerical Investigation of a Lead-Free Inorganic Layered Double Perovskite Cs4CuSb2Cl12 Nanocrystal Solar Cell by SCAPS-1D. Nanomaterials 2021, 11, 2321. https://doi.org/10.3390/nano11092321

AMA Style

He Y, Xu L, Yang C, Guo X, Li S. Design and Numerical Investigation of a Lead-Free Inorganic Layered Double Perovskite Cs4CuSb2Cl12 Nanocrystal Solar Cell by SCAPS-1D. Nanomaterials. 2021; 11(9):2321. https://doi.org/10.3390/nano11092321

Chicago/Turabian Style

He, Yizhou, Liyifei Xu, Cheng Yang, Xiaowei Guo, and Shaorong Li. 2021. "Design and Numerical Investigation of a Lead-Free Inorganic Layered Double Perovskite Cs4CuSb2Cl12 Nanocrystal Solar Cell by SCAPS-1D" Nanomaterials 11, no. 9: 2321. https://doi.org/10.3390/nano11092321

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