Next Article in Journal
Transport Sustainability Index: An Application Multicriteria Analysis
Previous Article in Journal
Smart Energy Management Strategy for Microgrids Powered by Heterogeneous Energy Sources and Electric Vehicles’ Storage
Previous Article in Special Issue
Evaluation of the Provincial Carbon Neutrality Capacity of the Middle and Lower Yellow River Basin based on the Entropy Weight Matter-Element Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient CsPbBr3 Perovskite Solar Cells with Storage Stability > 340 Days

School of Science, China University of Geosciences, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(20), 7740; https://doi.org/10.3390/en15207740
Submission received: 9 September 2022 / Revised: 12 October 2022 / Accepted: 13 October 2022 / Published: 19 October 2022
(This article belongs to the Special Issue New Insights into Geo-Energy, Geo-Resources and Environment)

Abstract

:
For CsPbBr3 perovskite materials, it is especially important to reduce interface defects, suppress non-radiative recombination, and improve morphology to achieve highly efficient and stable CsPbBr3 perovskite solar cells (PSCs). Herein, we reported a facile but highly efficient approach in additive engineering for improving the efficiency and stability of CsPbBr3 PSCs. It was found that phenethylammonium iodide can passivate interface defects, suppress non-radiative recombination, and increase the grain sizes of CsPbBr3 films by optimizing crystal quality and interface contact. As a result, a carbon-based CsPbBr3 PSC with power conversion efficiency > 8.51%, storage stability > 340 days, and excellent harsh stability under high temperature and humidity, has been achieved.

1. Introduction

Perovskite materials have shown enormous potential in photovoltaic power generation [1] due to their excellent characteristics, such as a tunable band gap, high light absorption coefficient, high carrier mobility, and weak exciton binding energy [2]. The power conversion efficiency (PCE) of perovskite solar cells (PSCs) has been improved from 3.8% to 25.7% in 11 years [3,4,5]. However, the long-term instability of organic–inorganic hybrid perovskite materials upon exposure to moisture, oxygen, and heat caused by organic cations [4] remains an issue. To improve the stability of perovskite materials, inorganic cesium ion (Cs+) was introduced to replace the organic counterparts to make all-inorganic perovskite materials, which was demonstrated to remarkably boost stability, especially in full-brominated CsPbBr3 [6,7]. CsPbBr3 was first applied into PSCs in 2015, presenting a PCE of 5.95% with outstanding moisture, oxygen, and thermal stability under harsh conditions [8]. Then, in 2016, Liu’s group abandoned expensive Spiro-OMeTAD hole transport layers (HTL) and Au/Ag electrodes, designed the carbon-based CsPbBr3 PSC, and obtained a PCE of 6.7% [8,9,10,11,12]. In 2018, Tang’s group creatively put forward a multi-step spin-coating method to improve the crystal quality and morphology of the CsPbBr3 active layers, resulting in a certified PCE of 9.72% [8,9,10,11,12,13]. Recently, the PCE of CsPbBr3 PSCs was further increased to over 10% via a doping process, interface modification, or spectra engineering [8,9,10,11,12,14]. Although it improved rapidly, the performance of CsPbBr3 PSCs is still unsatisfactory due to the wide band gap and severe carrier recombination of CsPbBr3 films, which will cause a decline in PCE and stability [10]. Furthermore, the wide band gap of CsPbBr3 leads to a large optical absorption loss, which significantly hinders further PCE improvements of CsPbBr3 PSCs [3,12]. A widely reported solution to this impasse is partially substituting Br- with I- to reduce the band gap; however, the long-term stability declines significantly after the substitution [12,15,16]. So, long-term stability is still a problem for CsPbBr3 PSCs when widening optical absorption [11,15,17,18,19]. Regarding the severe carrier recombination in CsPbBr3 PSCs, a mass of defects at the interface and grain boundaries usually acts as non-radiative recombination centers. Furthermore, different from organic–inorganic hybrid perovskite materials, a higher annealing temperature over 200 °C is necessary for CsPbBr3 films, which leads to a rapid growth of crystals, resulting in a poor crystal quality with lots of defects, uneven grain size, and large traps. Thus, for the CsPbBr3 perovskite materials, it is especially important to reduce interface defects, suppress non-radiative recombination, and improve morphology to achieve highly efficient CsPbBr3 PSCs [11,20]. Recently, it has been widely reported that phenethylammonium iodide (PEAI) can passivate interface defects, suppress non-radiative recombination [21,22], and increase the grain sizes of organic–inorganic hybrid perovskites by doping into the precursor solution [23]. However, the application of PEAI in the synthesis of CsPbBr3 perovskite film has never reported.
Herein, CsPbBr3 films were synthesized by a multi-step spin-coating method, and PEAI was introduced into the CsPbBr3 films for the first time via doping into the PbBr2 precursor solution. It was found that PEAI can passivate interface defects, suppress non-radiative recombination, and increase the grain sizes of CsPbBr3 films by optimizing crystal quality and interface contact. As a result, a carbon-based CsPbBr3 PSC with a PCE > 8.51%, storage stability > 340 days, and excellent harsh stability under high temperature and humidity has been achieved. The whole preparation process of CsPbBr3 PSCs was carried out in ambient air without humidity control.

2. Materials and Methods

We performed all experiments with the air at room temperature without humidity and temperature control.

2.1. Substrate and Precursor Solutions Preparation

We cleaned the dioxide transparent conductive glass doped with fluoride (FTO) by ultrasonic successively dipped in acetone, isopropanol, ethanol, deionized water, and isopropanol for 15 min. Then, we cleaned the FTO substrates by plasma. After that, we deposited the compact TiO2 (c-TiO2) from a solution of titanium tetraisopropoxide and hydrochloric acid in anhydrous ethanol by spin-coating at 3000 rpm for 30 s. We annealed the c-TiO2 layer at 500 °C for 30 min. We diluted mesoporous TiO2-layer (m-TiO2) pastes (30 NR-D, Dyesol) in ethanol (1:4, weight ratio), then spin-coated at 5000 rpm for 45 s, and finally annealed at 500 °C for 30 min, as described in our previous article [24]. We prepared the PbBr2 precursor solution by dissolving 367 mg PbBr2 in 1 mL N,N-Dimethylformamide (DMF) under active stirring for 12 h. We doped the PbBr2 precursor solutions of the experimental groups by PEAI at concentrations of 5, 10, and 15 mg/mL.

2.2. Multi-Step Spin-Coating Process of CsPbBr3 Films

The multi-step spin-coating process of CsPbBr3 films is illustrated in Figure S1 in the Supplementary Materials. We placed the FTO/c-TiO2/m-TiO2 substrates and the prepared PbBr2 precursor solution on a 90 °C hot plate to preheat. We spin-coated the PbBr2 solution on the substrates at a speed of 2000 rpm for 30 s. We placed the substrates with the spin-coated PbBr2 solution on a hot plate at 90 °C and annealed for 30 min. Subsequently, we spin-coated the 0.07 M CsBr solution on the PbBr2 layers 5 times. Each time, we annealed those substrates on a hot plate at 250 °C for 5 min, which was also reported in our previous article [25].

2.3. Device Fabrication

We scrapped the carbon electrode on the active layers. The area of each battery electrode is 0.01 cm2. We obtained the solar cell devices after annealing 120 °C for 15 min. Schematic structure of the carbon-based mesoscopic PSC is shown in Figure 1a.

2.4. Characterization

We characterized the surface morphology of perovskite films by scanning electron microscope (SEM). We tested the structure and components by X-ray diffraction technique (XRD, D8 fucox) and X-ray photoelectron spectroscopy (XPS). We determined the light absorbance spectra by UV–vis absorption spectra. We conducted Photoluminescence (PL) and time-resolved photoluminescence (TRPL) spectra to characterize the carrier transport properties. We measured dark J-V and space-charge limited-current (SCLC) curves to analyze defect-related information. We conducted current density–voltage (J-V) and monochromatic incident photon-to-electron conversion efficiency (EQE) experiments to characterize the performance of PSCs.

3. Results and Discussion

XRD spectra of CsPbBr3 films without and with PEAI doping in different concentration have been performed as shown in Figure 1b. The optical setup and scan parameters of XRD scans for all the four samples are unified. Three characteristic peaks centered at around 15.2°, 21.7°, 26.5°, and 30.8° correspond to (100), (110), (111), and (200) planes of CsPbBr3 phase, respectively. Obviously, all the four spectra exhibit almost identical peak positions, indicating the introduction of PEAI does not change the crystal structure and lattice constant of CsPbBr3 films. However, the (100) and (200) diffraction peaks of CsPbBr3 phase are dramatically enhanced after doping PEAI at 10 mg/mL. This is clear evidence for the optimization of crystallinity by PEAI doping at 10 mg/mL [14,26]. Furthermore, the appearance of the characteristic peaks of CsPb2Br5 and Cs4PbBr6 phases demonstrates that there are residuals of PbBr and CsBr in the four CsPbBr3 films.
SEM has been introduced to characterize the surface morphology of CsPbBr3 films. Figure 2a–d show top-view SEM images of CsPbBr3 films undoped and doped with PEAI at 5, 10, and 15 mg/mL. We observed that the grain sizes of the undoped CsPbBr3 film are small, and they increase with the further concentration of the PEAI additive. The grain sizes of the CsPbBr3 films are undoped and doped with PEAI at 5, 10, and 15 mg/mL are 0.45, 0.67, 1.11, and 1.13 μm, respectively, which are obtained by the “Nanomeasurer” software. It is reported that most of these small crystal grains in the undoped CsPbBr3 film are the phase of CsPb2Br5, which agrees with our previous XRD results [1]. The introduction of PEAI affects the kinetics of the crystallization process of CsPbBr3 films, inhibits the formation of CsPb2Br5, and finally increases the grain size [9]. In addition, for PEAI-doped CsPbBr3 films, it is obvious that two-dimensional (2D) perovskite material has formed at the grain boundaries, which is reported to generate in the crystallization process of CsPbBr3 and could passivate interface defects at the grain boundaries [26]. The formation of 2D perovskite material caused by the addition of PEAI is not significantly visible when the concentration of PEAI is low, such as at 5 mg/mL, as shown in Figure 2b. However, by adding an excessive amount of PEAI, the 2D perovskite material is clearly observable, as presented in Figure 2d. Figure 2e,f present the cross-sectional morphological properties of CsPbBr3 films undoped and doped with PEAI at 10 mg/mL, respectively. The thickness of the CsPbBr3 film after PEAI doping is ~650 nm, the same as the undoped film. We can see that the undoped CsPbBr3 film exhibits an uneven surface and obvious pinholes. The extrusion of the grains against each other results in partial convex grains, leaving notable pinholes between the substrate and CsPbBr3 film. On the contrary, the PEAI-doped CsPbBr3 film presents uniform and full coverage. The increased grain size, reduced interface defects, and uniform and full coverage of PEAI-doped CsPbBr3 films will promote light capture and charge transportation.
Combined XRD and SEM results confirm that the optimal doping concentration of PEAI is 10 mg/mL; therefore, CsPbBr3 films doped with PEAI at 10 mg/mL were used for the following film characterization and PSC application.
To further verify the elemental composition and chemical structure of the CsPbBr3 films without and with PEAI doping, XPS has been performed. Figure 3a–c are the high-resolution XPS spectra of Cs 3d, Pb 4f, and Br 3d, respectively. We can see that Pb and Br peaks shift to higher binding energies upon doping PEAI into CsPbBr3 films, while peaks of Cs 3d3/2 and Cs 3d5/2 present nearly no variations. It is known that large-radius PEA+ cations play a supporting role in the perovskite crystallization process and promote the oriented growth of PEAI-doped CsPbBr3 films [6,14], leading to the formation of 2D perovskite material, which may cause the distortion of the PbBr6 octahedron and then increase the electron density of Pb and Br atoms. Therefore, for PEAI-doped CsPbBr3 film, the XPS spectra of Pb 4f and Br 3d present blue shifts. In addition, based on the Pb 4f and Br 3d nuclear-energy-level spectra in Figure 3d, the Pb:X (X = Br) ratio of the undoped CsPbBr3 film can be estimated to be 1:1.665 (calculated by Avantage software), indicating that there is a serious lack of Br on the surface, which may lead to obvious point defects. After the introduction of PEAI, the Pb:X (X = Br, I) ratio increased to 1:1.689, demonstrating that the point defects on the perovskite surface have been partially passivated and the Br vacancies are likely to be filled by a small amount of I in PEAI [24].
UV–vis absorbance spectra have been carried out, and the results are shown in Figure 4a. Obviously, there is no change in the absorption edge of both perovskite films, indicating that the PEAI doping does not affect the band gap of CsPbBr3 films [2,27]. In addition, a slight increase in the optical absorption is observed for the PEAI-doped CsPbBr3 film, as shown in Figure 4a, which can be explained as follows: the uniform and full coverage of PEAI-doped CsPbBr3 film will promote light capture and lead to improved optical absorption. Steady PL and TRPL measurements are presented in Figure 4a,b, respectively. Here, the samples used for steady PL and TRPL tests are CsPbBr3 films directly deposited on the Al2O3 substrates without an electron transport layer (ETL) or HTL. As a result, light-induced carriers in an excited state cannot be extracted quickly, leading to radiation recombination. Here, higher steady PL intensity and slower TRPL decay rates demonstrate fewer traps or defects. We can see that the steady PL spectra of CsPbBr3 films undoped and doped with PEAI exhibit an identical peak centered at about 532 nm, demonstrating that the PEAI doping does not affect the crystal structure of CsPbBr3 films, which fully agrees with the UV–vis results. In addition, the PEAI-doped perovskite film exhibits a remarkably enhanced PL intensity compared with the undoped film, implying that the trap and defect states in CsPbBr3 films are reduced. The TRPL spectra in Figure 4b can be well fitted by a bi-exponential decay function [11,22], and the corresponding lifetimes τ1, τ2, and τave are obtained. After PEAI treatment, both τ1, τ2, and τave increase greatly, from 15.12, 539.87, and 514.68 ns to 61.54, 700.80, and 635.60 ns, respectively, which can be attributed to the suppression of defect-related non-radiative recombination after PEAI doping, which is consistent with the steady PL results.
The dark J–V measurements have been conducted with the architecture of FTO/c-TiO2/m-TiO2/CsPbBr3/carbon, and the corresponding results are shown in Figure 5. It can be seen from Figure 5a that the dark current of the PEAI-doped device is about two orders of magnitude lower than the undoped device, implying reduced charge carrier recombination losses, which can be due to the reduced interface defects in the CsPbBr3 layer after PEAI doping, as shown in previous SEM and XPS results [20,28]. The SCLC curves, which were measured with FTO/CsPbBr3/Au architectures under dark conditions, are displayed in Figure 5b. The onset voltage of the trap-filled limit (VTFL) can be obtained from the SCLC curves, and then we can calculate the trap state density (Nt) according to the following equation [6]:
N t = 2 ε r ε 0 V T F L q L 2
In this equation, L is the thickness of perovskite films, εr is the relative dielectric constant for CsPbBr3, and ε0 is the vacuum permittivity. From the SCLC curves in Figure 5b, the VTLF of the undoped and PEAI-doped CsPbBr3 layers are 1.0727 and 0.7915 V, respectively. Combined with Formula (1), the Nt of the undoped and PEAI-doped CsPbBr3 layers are calculated to be 9.907 × 1015 and 7.310 × 1015 cm−3, respectively. It can be inferred that the trap states that lead to carrier recombination are significantly reduced after PEAI treatment, cross-checking the above-mentioned conclusion [6].
A series of CsPbBr3 PSCs are fabricated with an architecture of FTO/compct-TiO2/mesoporous-TiO2/CsPbBr3/carbon. The active area is 0.1 cm2, and the whole preparation process of the PSCs was finished in ambient air. The dependences of photovoltaic performances on PEAI concentrations are reflected by Figure S2 in the Supplementary Materials, and the corresponding photovoltaic parameters are summarized in the inset of Figure S2, which shows that the optimal doping concentration of PEAI is 10 mg/mL, consistent with the XRD and SEM characterization above.
Figure 6a shows the characteristic J–V curves of PSCs based on CsPbBr3 films without and with PEAI doping at 10 mg/mL, and the corresponding photovoltaic parameters, short-circuit current density (JSC), open-circuit voltage (VOC), fill factor (FF), and PCE are summarized in the inset of Figure 6a. Obviously, the photoelectric conversion capability of devices has been tremendously improved upon PEAI doping, yielding a champion PCE of 8.51%. Specifically, VOC rises from 1.28 V to 1.39 V, JSC from 7.49 mA/cm2 to 8.22 mA/cm2, FF from 68% to 74%, and the PCE is enhanced by 26.87% compared with 6.55% for the undoped device. All the photovoltaic parameters increase notably, indicating that the total series resistance of the entire device is reduced by PEAI doping, which can be attributed to the reduced traps and defects, as confirmed above. From Figure 6a, we also noticed that a smaller J–V hysteresis can be achieved in the PEAI-doped PSC under forward and reverse scan directions. The hysteresis index [29], which is the ratio of PCE under reverse and forward scan, is decreased from 1.21 to 1.11 after PEAI doping, implying the passivation of interface defects in the PEAI-doped CsPbBr3 layer. It was reported that ion migration tended to occur at the defects/grain boundary in the active layer, which leads to J–V hysteresis [23,27]. The IPCE spectrum of the champion device is shown in Figure 6b, from which the integrated current density is calculated to be 7.68 mA/cm2, agreeing well with the J–V result and confirming the accuracy of our measurement.
As known, improving cell efficiency and reducing costs without sacrificing long-term stability are necessary for future commercial applications [12]. To study the storage stability, the PSC doped with PEAI at 10 mg/mL stored in ambient air (temperature: 10~20 °C, relative humidity: 20~30%) without encapsulation was measured. As shown in Figure 6c, for the device doped with PEAI, the PCE not only did not decline but increased after a few months of storage. According to reports, the PCE enhancement of CsPbBr3 solar cells after storage may be attributed to the improved crystal quality of CsPbBr3 film under the effect of water molecules. Furthermore, it should be noted that the efficiency of PEAI-doped PSC remained nearly unchanged for more than 11 months (340 days). For comparison, the storage stability of the non-doped PSC was measured under the same conditions and shown in Figure S3 in the Supplementary Materials. Obviously, the storage stability is remarkably improved after PEAI doping. In addition to storage stability in ambient air, stability under harsh conditions, such as high humidity and temperature, was also tested, as shown in Figure 6d. We can see that the PCE of PEAI-doped PSCs has almost no reduction under 80% humidity or at 80 °C without encapsulation within 30 or 14 days, while the devices without PEAI doping declined rapidly. The improvement of device stability after PEAI doping could also be attributed to the reduced traps and defects, which can lead to the suppression of charge recombination and accumulation.

4. Conclusions

In conclusion, to reduce interface defects, suppress non-radiative recombination, and improve the morphology of the CsPbBr3 films to achieve highly efficient and stable CsPbBr3 PSCs, we reported a facile but highly efficient approach for additive engineering. CsPbBr3 films were synthesized by a multi-step spin-coating method, and PEAI was introduced into the CsPbBr3 films for the first time via doping into the PbBr2 precursor solution. We found that PEAI additive can passivate interface defects, suppress non-radiative recombination, and increase the grain sizes of CsPbBr3 films by optimizing crystal quality and interface contact. By virtue of PEAI doping, a carbon-based CsPbBr3 PSC with a PCE > 8.51%, storage stability > 340 days, and excellent stability under high temperature and humidity, has been achieved. Our work contributes to improving the performance of CsPbBr3 PSCs.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en15207740/s1, Figure S1: Schematic diagram of multi-step spin-coating process to prepare CsPbBr3 films; Figure S2: J–V characteristics of various CsPbBr3 PSCs under illumination (AM 1.5 G, 100 mW cm−2 ) undoped and doped with 5, 10 and 15 mg/mL PEAI, measured at the maximum power point; Figure S3: Storage stability (in the ambient air, without encapsulation) of CsPbBr3 PSCs without PEAI-doping; Figure S4: J–V characteristics and storage stability (in the ambient air, without encapsulation) of CsPbBr3 PSCs with different electrode.

Author Contributions

Conceptualization, S.H. and S.W.; methodology, X.L.; software, J.X.; validation, S.H., S.W. and X.L.; formal analysis, S.H.; investigation, H.H. (Haochong Huang); resources, J.D.; data curation, J.Y.; writing—original draft preparation, S.H.; writing—review and editing, S.W.; visualization, H.L.; supervision, H.H. (Huiying Hao); project administration, J.D.; funding acquisition, J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Fundamental Research Funds for the Central Universities” (Grant No. 2652019121) and “the National Natural Science Foundation of China” (Grant No. 11404293).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, F.; Guo, Y.; Wang, X.; Tao, J.; Li, Z.; Zheng, D.; Jiang, J.; Hu, Z.; Chu, J. Efficient carbon-based planar CsPbBr3 perovskite solar cells with Li-doped amorphous Nb2O5 layer. J. Alloy. Compd. 2020, 842, 1559840. [Google Scholar] [CrossRef]
  2. Xu, W.; Gao, Y.; Ming, W.; He, F.; Li, J.; Zhu, X.; Kang, F.; Li, J.; Wei, G. Suppressing Defects-Induced Nonradiative Recombination for Efficient Perovskite Solar Cells through Green Antisolvent Engineering. Adv. Mater. 2020, 32, 1–10. [Google Scholar] [CrossRef] [PubMed]
  3. Miyasaka, T.; Kulkarni, A.; Kim, G.M.; Öz, S.; Jena, A. Perovskite Solar Cells: Can We Go Organic-Free, Lead-Free, and Dopant-Free? Adv. Energy Mater. 2019, 10, 1–20. [Google Scholar] [CrossRef]
  4. Wang, Y.; Hu, Z.; Yao, W.; Yang, C.; Zhang, H.; Zhang, J.; Zhu, Y. Heterogeneous photoresponse of individual grain in all-inorganic perovskite solar cells. Appl. Phys. Lett. 2020, 117, 083902–083906. [Google Scholar] [CrossRef]
  5. NREL. Best Research-Cell Efficiency Chart. 2022. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 25 September 2022).
  6. Duan, J.; Zhao, Y.; Yang, X.; Wang, Y.; He, B.; Tang, Q. Lanthanide Ions Doped CsPbBr3 Halides for HTM-Free 10.14%-Efficiency Inorganic Perovskite Solar Cell with an Ultrahigh Open-Circuit Voltage of 1.594 V. Adv. Energy Mater. 2018, 8, 1–9. [Google Scholar] [CrossRef]
  7. Ren, Y.; Zhang, N.; Arain, Z.; Mateen, M.; Chen, J.; Sun, Y.; Li, Z. Polymer-induced lattice expansion leads to all-inorganic CsPbBr3 perovskite solar cells with reduced trap density. J. Power Sources 2020, 475, 228676. [Google Scholar] [CrossRef]
  8. Liu, X.; Liu, Z.; Tan, X.; Ye, H.; Sun, B.; Xi, S.; Shi, T.; Tang, Z.; Liao, G. Novel antisolvent-washing strategy for highly efficient carbon-based planar CsPbBr3 perovskite solar cells. J. Power Sources 2019, 439, 227092. [Google Scholar] [CrossRef]
  9. Liu, J.; Zhu, L.; Xiang, S.; Wei, Y.; Xie, M.; Liu, H.; Li, W.; Chen, H. Growing high-quality CsPbBr3 by using porous CsPb2Br5 as an intermediate: A promising light absorber in carbon-based perovskite solar cells. Sustain. Energy Fuels 2019, 3, 184–194. [Google Scholar] [CrossRef]
  10. Xu, H.; Duan, J.; Zhao, Y.; Jiao, Z.; He, B.; Tang, Q. 9.13%-Efficiency and stable inorganic CsPbBr3 solar cells. Lead-free CsSnBr3-xIx quantum dots promote charge extraction. J. Power Sources 2018, 399, 76–82. [Google Scholar] [CrossRef]
  11. Tong, G.; Chen, T.; Li, H.; Qiu, L.; Liu, Z.; Dang, Y.; Song, W.; Ono, L.K.; Jiang, Y.; Qi, Y. Phase transition induced recrystallization and low surface potential barrier leading to 10.91%-efficient CsPbBr3 perovskite solar cells. Nano Energy 2019, 65, 104015. [Google Scholar] [CrossRef]
  12. 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]
  13. Duan, J.; Zhao, Y.; He, B.; Tang, Q. High-Purity Inorganic Perovskite Films for Solar Cells with 9.72 % Efficiency. Angew. Chem. 2018, 130, 3849–3853. [Google Scholar] [CrossRef]
  14. Wang, K.; Jin, Z.; Liang, L.; Bian, H.; Bai, D.; Wang, H.; Zhang, J.; Wang, Q.; Liu, S. All-inorganic cesium lead iodide perovskite solar cells with stabilized efficiency beyond 15%. Nat. Commun. 2018, 9, 1–8. [Google Scholar] [CrossRef] [Green Version]
  15. Zhu, J.; Tang, M.; He, B.; Shen, K.; Zhang, W.; Sun, X.; Sun, M.; Chen, H.; Duan, Y.; Tang, Q. Ultraviolet filtration and defect passivation for efficient and photostable CsPbBr3 perovskite solar cells by interface engineering with ultraviolet absorber. Chem. Eng. J. 2020, 404, 126548. [Google Scholar] [CrossRef]
  16. Jiang, J.; Jin, Z.; Gao, F.; Sun, J.; Wang, Q.; Liu, S. CsPbCl3-Driven Low-Trap-Density Perovskite Grain Growth for >20% Solar Cell Efficiency. Adv. Sci. 2018, 5, 1800474. [Google Scholar] [CrossRef]
  17. Wang, Z.; Lin, Q.; Chmiel, F.P.; Sakai, N.; Herz, L.; Snaith, H. Efficient ambient-air-stable solar cells with 2D–3D heterostructured butylammonium-caesium-formamidinium lead halide perovskites. Nat. Energy 2017, 2, 17135. [Google Scholar] [CrossRef]
  18. Liu, M.; Dahlström, S.; Ahläng, C.; Wilken, S.; Degterev, A.; Matuhina, A.; Hadadian, M.; Markkanen, M.; Aitola, K.; Kamppinen, A.; et al. Beyond hydrophobicity: How F4-TCNQ doping of the hole transport material improves stability of mesoporous triple-cation perovskite solar cells. J. Mater. Chem. A 2022, 10, 11721–11731. [Google Scholar] [CrossRef]
  19. Saliba, M.; Stolterfoht, M.; Wolff, C.M.; Neher, D.; Abate, A. Measuring Aging Stability of Perovskite Solar Cells. Joule 2018, 2, 1019–1024. [Google Scholar] [CrossRef] [Green Version]
  20. Zhou, Q.; Duan, J.; Yang, X.; Duan, Y.; Tang, Q. Interfacial Strain Release from the WS 2 /CsPbBr 3 van der Waals Heterostructure for 1.7 V Voltage All-Inorganic Perovskite Solar Cells. Angew. Chem. Int. Ed. 2020, 59, 21997–22001. [Google Scholar] [CrossRef]
  21. Li, N.; Zhu, Z.; Chueh, C.-C.; Liu, H.; Peng, B.; Petrone, A.; Li, X.; Wang, L.; Jen, A.K.-Y. Mixed Cation FAxPEA1-xPbI3with Enhanced Phase and Ambient Stability toward High-Performance Perovskite Solar Cells. Adv. Energy Mater. 2016, 7, 1–9. [Google Scholar] [CrossRef]
  22. Zhang, F.; Kim, D.H.; Lu, H.; Park, J.-S.; Larson, B.W.; Hu, J.; Gao, L.; Xiao, C.; Reid, O.G.; Chen, X.; et al. Enhanced Charge Transport in 2D Perovskites via Fluorination of Organic Cation. J. Am. Chem. Soc. 2019, 141, 5972–5979. [Google Scholar] [CrossRef] [PubMed]
  23. Jiang, Q.; Zhao, Y.; Zhang, X.; Yang, X.; Chen, Y.; Chu, Z.; Ye, Q.; Li, X.; Yin, Z.; You, J. Surface passivation of perovskite film for efficient solar cells. Nat. Photon- 2019, 13, 460–466. [Google Scholar] [CrossRef]
  24. Wu, J.; Dong, J.-J.; Chen, S.-X.; Hao, H.-Y.; Xing, J.; Liu, H. Fabrication of Efficient Organic-Inorganic Perovskite Solar Cells in Ambient Air. Nanoscale Res. Lett. 2018, 13, 293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Yan, J.; Hou, S.; Li, X.; Dong, J.; Zou, L.; Yang, M.; Xing, J.; Liu, H.; Hao, H. Preparation of highly efficient and stable CsPbBr3 perovskite solar cells based on an anti-solvent rinsing strategy. Sol. Energy Mater. Sol. Cells 2022, 234, 111420–111428. [Google Scholar] [CrossRef]
  26. Fu, W.; Liu, H.; Shi, X.; Zuo, L.; Li, X.; Jen, A.K. Tailoring the Functionality of Organic Spacer Cations for Efficient and Stable Quasi-2D Perovskite Solar Cells. Adv. Funct. Mater. 2019, 29, 1–8. [Google Scholar] [CrossRef]
  27. Yun, Y.; Wang, F.; Huang, H.; Fang, Y.; Liu, S.; Huang, W.; Cheng, Z.; Liu, Y.; Cao, Y.; Gao, M.; et al. A Nontoxic Bifunctional (Anti)Solvent as Digestive-Ripening Agent for High-Performance Perovskite Solar Cells. Adv. Mater. 2020, 32, e1907123. [Google Scholar] [CrossRef] [PubMed]
  28. Lee, K.; Kim, J.; Yu, H.; Lee, J.W.; Yoon, C.-M.; Kim, S.K.; Jang, J. A highly stable and efficient carbon electrode-based perovskite solar cell achieved via interfacial growth of 2D PEA2PbI4 perovskite. J. Mater. Chem. A 2018, 6, 24560–24568. [Google Scholar] [CrossRef]
  29. Canil, L.; Salunke, J.; Wang, Q.; Liu, M.; Köbler, H.; Flatken, M.; Gregori, L.; Meggiolaro, D.; Ricciarelli, D.; De Angelis, F.; et al. Halogen-Bonded Hole-Transport Material Suppresses Charge Recombination and Enhances Stability of Perovskite Solar Cells. Adv. Energy Mater. 2021, 11, 2101553. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic structure of CsPbBr3 PSCs; (b) XRD patterns of CsPbBr3 films without and with PEAI doping in different concentrations.
Figure 1. (a) Schematic structure of CsPbBr3 PSCs; (b) XRD patterns of CsPbBr3 films without and with PEAI doping in different concentrations.
Energies 15 07740 g001
Figure 2. Top-view SEM images of CsPbBr3 films (a) undoped and doped with PEAI at (b) 5, (c) 10, and (d) 15 mg/mL. Cross-sectional SEM images of CsPbBr3 films (e) undoped and (f) doped with 10 mg/mL PEAI.
Figure 2. Top-view SEM images of CsPbBr3 films (a) undoped and doped with PEAI at (b) 5, (c) 10, and (d) 15 mg/mL. Cross-sectional SEM images of CsPbBr3 films (e) undoped and (f) doped with 10 mg/mL PEAI.
Energies 15 07740 g002
Figure 3. High-resolution XPS spectra of (a) Cs 3d, (b) Pb 4f, and (c) Br 3d. Pb 4f and Br 3d core-energy-level spectra are compared in (d).
Figure 3. High-resolution XPS spectra of (a) Cs 3d, (b) Pb 4f, and (c) Br 3d. Pb 4f and Br 3d core-energy-level spectra are compared in (d).
Energies 15 07740 g003
Figure 4. (a) UV–vis absorbance, steady PL, and (b) TRPL spectra of CsPbBr3 films undoped and doped with 10 mg/mL PEAI without ETL or HTL.
Figure 4. (a) UV–vis absorbance, steady PL, and (b) TRPL spectra of CsPbBr3 films undoped and doped with 10 mg/mL PEAI without ETL or HTL.
Energies 15 07740 g004
Figure 5. (a) Dark J-V and (b) SCLC versus voltage curves of undoped and PEAI-doped devices (VTFL are obtained to be 1.061 and 0.776 V for the undoped and PEAI-doped devices, respectively).
Figure 5. (a) Dark J-V and (b) SCLC versus voltage curves of undoped and PEAI-doped devices (VTFL are obtained to be 1.061 and 0.776 V for the undoped and PEAI-doped devices, respectively).
Energies 15 07740 g005
Figure 6. Characteristic J–V curves of PSCs based on CsPbBr3 films without and with PEAI doping at 10 mg/mL, and (a) the inset shows corresponding photovoltaic parameters; (b) PCE spectrum of the champion device; (c) storage stability in the ambient air without encapsulation; (d) stability under harsh conditions.
Figure 6. Characteristic J–V curves of PSCs based on CsPbBr3 films without and with PEAI doping at 10 mg/mL, and (a) the inset shows corresponding photovoltaic parameters; (b) PCE spectrum of the champion device; (c) storage stability in the ambient air without encapsulation; (d) stability under harsh conditions.
Energies 15 07740 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hou, S.; Wu, S.; Li, X.; Yan, J.; Xing, J.; Liu, H.; Hao, H.; Dong, J.; Huang, H. Efficient CsPbBr3 Perovskite Solar Cells with Storage Stability > 340 Days. Energies 2022, 15, 7740. https://doi.org/10.3390/en15207740

AMA Style

Hou S, Wu S, Li X, Yan J, Xing J, Liu H, Hao H, Dong J, Huang H. Efficient CsPbBr3 Perovskite Solar Cells with Storage Stability > 340 Days. Energies. 2022; 15(20):7740. https://doi.org/10.3390/en15207740

Chicago/Turabian Style

Hou, Shaochuan, Siheng Wu, Xiaoyan Li, Jiahao Yan, Jie Xing, Hao Liu, Huiying Hao, Jingjing Dong, and Haochong Huang. 2022. "Efficient CsPbBr3 Perovskite Solar Cells with Storage Stability > 340 Days" Energies 15, no. 20: 7740. https://doi.org/10.3390/en15207740

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