Stability Improvement of Perovskite Solar Cells for Application of

Feb 15, 2019 - In this work, we developed a CH3NH3PbI3 perovskite solar cell with CuInS2 quantum dot-modified TiO2 nanoarrays (TiO2–CuInS2 QD-NAs) ...
0 downloads 0 Views 5MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 3432−3438

http://pubs.acs.org/journal/acsodf

Stability Improvement of Perovskite Solar Cells for Application of CuInS2 Quantum Dot-Modified TiO2 Nanoarrays Feng Gao,* Qing Zheng, and Ying Zhang School of Food and Chemical Engineering, Shaoyang University, Shaoyang 422000, P. R. China

Downloaded via 193.9.158.112 on February 16, 2019 at 00:38:31 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: In this work, we developed a CH3NH3PbI3 perovskite solar cell with CuInS2 quantum dot-modified TiO2 nanoarrays (TiO2−CuInS2 QD-NAs) as a scaffold layer. Based on the suitable device configuration, we achieved improved power conversion efficiency (PCE) of 13.3%, which was 38.3% higher than that of the device without QD modification (8.2%). After exposure to air for 30 days, the TiO2−CuInS2 QD-NA-based device possessed a PCE of 5.4%, being 41% of the original performance, which was far superior to that of TiO2 nanoarray-based solar cells with a PCE of 1.1%. Our results showed that the crystallinity of perovskite, surface state, and interface for charge transport of TiO2−CuInS2 QD-NAbased perovskites all remarkably improved, indicating the improved air stability for TiO2−CuInS2 QD-NA-based solar cells.



devices,34 and perovskite-based solar cells.35,36 In our recent research, TiO2 nanoarrays decorated with CuInS2 QDs (TiO2−CuInS2 QD-NAs) were fabricated to be applied in MAPbI3 solar cells. These TiO2−CuInS2 QD-NAs are found to be advanced interfacial ETMs, which can effectively improve the stability of hybrid MAPbI3-based solar cells. The results show that the crystallinity of perovskite, surface state, and interface for charge transport of TiO2−CuInS2 QD-NA-based perovskites all remarkably improved in fluorine doped tin oxide (FTO)/TiO 2−CuInS2 QDs/MAPbI3/spiro-MeOTAD/Au solar cells. Additionally, we observed that the TiO2−CuInS2 QD-NA-based devices greatly improved air stability compared to that of the pure TiO2-NA-based MAPbI3 solar cells.

INTRODUCTION Recently, metal halide perovskites have been recognized to be most outstanding photoelectric materials due to their special performances1−5 including the well-matched optical band gap,1 prominent charge diffusion lengths,6 low fabrication property,7 very low temperature processability, 8 and high light absorbance.9 This aroused great enthusiasm among scientific researchers for the rapid enhancement of power conversion efficiency (PCE). Nowadays, tremendous efforts have been made by various engineering techniques such as a crystal culture method,10−12 element modification,13 interfacial modification,14−16 and the application of effective carriertransport materials,8,17,18 propelling the PCE to reach a high value up to 20%.19−25 However, the poor stability remains to be the major obstacle in the potential commercialization of perovskite photovoltaic devices. The hybrid perovskite material is found to be easy to decompose into PbI2 and volatile organic groups under moisture corrosion, photo-oxidation, and a high-temperature environmental due to the instability of the weak bonding effect.26−28 Interfacial modification and controlling have been known as one of the key methods for enhancing stability in hybrid perovskite solar cells.29−31 Herein, we improved the iodine-based hybrid perovskite CH3NH3PbI3 (MAPbI3)/electron transport material (ETM) interface using TiO2 nanoarrays (TiO2-NAs) embellished with chalcopyrite CuInS2 quantum dots (CuInS2 QDs), which have been considered as excellent photovoltaic materials due to high light absorbance (ca. 105 cm−1), suitable band gap (ca. 1.6 eV), and low toxicity.32 These have been applied in some fields such as polymer-based photovoltaic devices,33 ternary hybrid © 2019 American Chemical Society



RESULTS AND DISCUSSION Figure 1a shows the absorption spectrum of CuInS2 QDs, which shows wide absorption in the region from 300 to 800 nm, indicating their excellent light absorption properties in the whole ultraviolet−visible region. Coincided with chalcopyrite CuInS2, the absorption band gap of as-synthesized CuInS2 QDs was calculated to be about 1.75 eV, which matches well with the energy band gap requirements.37 The TiO2-NAs show a wide shoulder in the region of 300−360 nm, which is in good agreement with the previous work (Figure 1b).38 The strong peak at the wavelength of 310 nm is based on Ti (3d)−O (2p) energy transition of TiO2 nanoarrays.38 Figure 1c shows the absorption spectra of MAPbI3 films coated on TiO2 nanoarrays Received: December 26, 2018 Accepted: February 7, 2019 Published: February 15, 2019 3432

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega

Article

Figure 1. Absorption spectra of CuInS2 QDs (a), TiO2 nanoarrays and CuInS2 QD-modified TiO2 nanoarrays (b), MAPbI3 thin films coated on TiO2 nanoarrays and QD-modified TiO2 nanoarrays (c), and XRD patterns of various thin films (d). The solid circle in (d) indicates the XRD signals of FTO. The inset in (a) is the optical energy band gap curve of CuInS2 QDs.

and TiO2−CuInS2 QD array films. Both TiO2-NAs/MAPbI3 and TiO2−CuInS2 QD-NAs/MAPbI3 exhibit two bands at about 480 and 760 nm (Figure S1), for which the optical transition of MAPbI3 should be responsible.39 Obviously, the absorption intensity of the TiO2−CuInS2 QD-NAs/MAPbI3 is enhanced compared to that of the TiO2-NAs/MAPbI3 sample, which can be explained by two aspects. on the one hand, the wide absorption response of CuInS2 QDs can add to the absorption intensity of the films. On the other hand, the fine crystal changes of the MAPbI3 on TiO2−CuInS2 QD-NAs (Xray diffraction (XRD) data discussed in Figure 6) should be responsible for the higher absorption intensity. The XRD patterns of chalcopyrite CuInS2 QDs, TiO2 nanoarrays, and QD-modified TiO2 nanoarrays are shown in Figure 1d. From the XRD patterns, we can see that the chalcopyrite CuInS2 QD peaks recorded at 2θ = 28, 47, and 55° are attributed to the (112), (220), and (312) diffraction crystal planes, respectively (JCPDS card #86-0147).40 To further study the CuInS2 QDs, we measured the size of as-synthesized CuInS2 QDs by Scherrer’s formula. The average size of the CuInS2 QDs was estimated to be 1−3 nm, consistent with that of the pure particles in transmission electron microscopy (TEM) data (Figure S2). It can be clearly seen in the TiO2-NA patterns that three diffraction peaks attributed to (101), (211), and (002), respectively, appeared, which is in good agreement with the rutile TiO2 crystal marked by JCPDS card #21-1276.40 Different from TiO2-NAs, the new (112) peak originated from the chalcopyrite CuInS2 QDs in TiO2−CuInS2 QD-NA samples, suggesting the successful crystallization of CuInS2 QDs on the TiO2 nanoarray substrate. The crystal growth direction is easily implied to be preferential along the [112] direction. Transmission electron microscopy (TEM) images were obtained to study the nanoscale surface morphologies and the interfacial state of the TiO2-NA and QD-modified TiO2-NA samples. The pure TiO2-NAs (Figures 2a, S3a, and S3b) are highly unique nanorods with average diameters of 40−50 nm

Figure 2. TEM (a) and SEDP (b) images of TiO2-NAs. TEM (c) and HR-TEM (d) images of TiO2−CuInS2 QD-NAs.

and corresponding lengths of 500 nm. The selected electron diffraction patterns (SEDPs) of the as-synthesized TiO2-NAs are displayed in Figure 2b, which are in quite agreement with those of the rutile TiO2 samples. Figures 2c, S3c, and S3d show the SEM and TEM images of TiO2−CuInS2 QD-NAs. Clearly, the CuInS2 QDs evenly dispersed on the surface of TiO2-NAs with a size of 3−5 nm, which was slightly larger than that of the pure CuInS2 QDs, for which the effect of the surface energy of TiO2-NAs should be responsible. The high-resolution transmission electron microscopy (HR-TEM) image of TiO2− CuInS2 QD-NAs is displayed in Figure 2d. CuInS2 QDs are clearly visible with an interplanar spacing of 0.25 nm, matching well with the (101) plane of the rutile TiO2.40 The thickness of the CuInS2 QD layer in the QD-modified TiO2-NAs was calculated to be 6.5 nm, which is almost 2 QDs in length. The photoluminescence (PL) spectra of the MAPbI3 thin films on FTO, TiO2-NAs, and TiO2−CuInS2 QD-NAs were investigated to confirm charge transfer behavior of CuInS2 QDs (Figure 3a). The pure MAPbI3 thin films on FTO showed a wide band located at 775 nm, which agrees well with 3433

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega

Article

between the MAPbI3 layer and TiO2-NAs. Time-resolved photoluminescence spectrum was further recorded to test the charge transfer time (τCT), which can be calculated by the formula 1/τfilm = 1/τMAPbI3 + 1/τCT, where τfilm is the decay lifetime of MAPbI3 films on the ETM substrate and 1/τMAPbI3 is the decay lifetime of pure MAPbI3 films. As shown in Table S1, the τMAPbI3 of the pure MAPbI3 films was tested and estimated to be 42 ns, by which the τfilm and τCT for TiO2-NAs/MAPbI3 thin films were calculated to be 11.0 and 14.9 ns, respectively. In comparison, the τfilm and τCT in TiO2−CuInS2 QD-NAs/ MAPbI3 were calculated to be 6.2 and 7.3 ns, which are much shorter than those of TiO2-NAs/MAPbI3 thin films, indicating that an improved charge transfer process occurred at the TiO2−CuInS2 QD-NA and MAPbI3 interface. Figure S4 shows the cross section of solar cells. The corresponding device structure and energy-level diagram are displayed in Figure 4a,b, respectively. The J−V curve data of TiO2-NA-based and TiO2−CuInS2 QD-NA-based solar cells are measured in Figure 4c. For the TiO2-NA-based device, the performance with an open-circuit voltage (Voc) of 0.93 V, a short-circuit current (Jsc) of 13.4 mA cm−2, and a PCE of 8.2% was obtained. Compared to those of the solar cells based on TiO2-NAs, the TiO2−CuInS2 QD-NA-based device showed remarkably improved performance (Table 1), where the Voc and Jsc were 0.98 V and 19.2 mA cm−2, respectively. The highest PCE of such a device was as high as 13.3%, which was 38.3% higher than that of the TiO2-NA-based device. The external quantum efficiency (EQE) spectra of the solar cells are recorded in Figure 4d. All of the devices showed a broad band in the region from 400 to 800 nm. The response peak at 730 nm can be attributed to the characteristic absorption of MAPbI3. The TiO2-NA-based device showed the highest EQE of only 62%. In contrast, the most efficient EQE of the TiO2−CuInS2 QD-NA-based solar cell was as high as

Figure 3. PL spectra (λex = 475 nm) (a) and time-resolved photoluminescence spectra (recorded at 775 nm, λex = 560 nm) of MAPbI3, TiO2-NAs/MAPbI3, and TiO2−CuInS2 QD-NAs/MAPbI3 thin films (b).

the previous work.40 The emission band was significantly quenched after coating of MAPbI3 on the TiO2-NA substrate, suggesting that an efficient charge transport channel existed between the MAPbI3 and TiO2-NA interface. What is more, the quenching intensity was remarkably enhanced when MAPbI3 contacted with TiO2−CuInS2 QD-NAs, implying the facilitated electron injection efficiency at the interface

Figure 4. (a) Device structure and (b) energy-level diagram of the TiO2−CuInS2-NA-based solar cells; (c) J−V characteristics of TiO2-NA-based and TiO2−CuInS2 QD-NA-based solar cells and (d) corresponding external quantum efficiency (EQE) spectra. 3434

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega

Article

Table 1. Device Performance of MAPbI3 Solar Cells Based on TiO2-NAs and TiO2−CuInS2 QD-NAs devices

Voc (V)

Jsc (mA cm−2)

FF

PCE (%)

TiO2-NAs/MAPbI3 TiO2−CuInS2 QD-NAs/MAPbI3

0.93 0.98

13.4 19.2

0.66 0.71

8.2 13.3

95%, much higher than that of the TiO2-NA-based solar cell. The results indicated that the QD-modified TiO2 nanoarrays can efficiently facilitate the charge transport process. To confirm this, we calculated the series resistance (Rs) using the typical eq 1,41 where the ideal diode behavior (n = 1) is assumed with IL and I0 representing photocurrent and dark saturation current (inverse polarization), respectively. Besides, q, T, K, and V correspond to the charge, absolute temperature, Boltzmann constant, and bias potential, respectively. i q(V + IR s) yz zz I = IL − I0 expjjjj z nkT (1) k { 2 Rs was estimated to be 2.81 Ω cm for the TiO2−CuInS2 QDNA-based solar cell and 3.74 Ω cm2 for the TiO2-NA-based solar cell, indicating that the charge transport behavior is remarkably improved. Thus, the enhancement of the performance, especially the photocurrent in the TiO2−CuInS2 QDNA-based solar cell, can be easily understood. To investigate the stability of the solar cells, we studied the large-scale surface state of MAPbI3 films under an air environment for 30 days (Figures 5 and S6). MAPbI3 coated

Figure 6. (a) XRD patterns of MAPbI3 films on TiO2-NA and TiO2− CuInS2 QD-NA substrates and (b) corresponding XRD patterns after 30 days. The XRD signals of TiO2 and FTO are indicated by different symbols.

(224), and (314) planes of the tetragonal MAPbI3 crystal, respectively.42 Notably, a new diffraction peak at ca. 36° appeared on TiO2−CuInS2 QD-NA samples that could be assigned to the (312) crystal plane of MAPbI3.43 This implied that the MAPbI3 crystal was more inclined to grow along the [312] direction. It should be explained that there was almost no XRD signal of CuInS2 QDs in the films because of the very low content of QDs in TiO2−CuInS2-NAs/MAPbI3 films. To study the phase change of MAPbI3 films on TiO2-NA and QDmodified TiO2-NA substrates in air for 30 days, the corresponding XRD patterns were recorded (Figure 6b). For further comparison, we obtained the XRD pattern of PbI2. The diffraction peaks at 2θ = 12, 26, 34, and 39° can be assigned to the (001), (101), (102), and (110) planes, respectively, of the PbI2 nanocrystal.44 The XRD pattern of MAPbI3 films on the TiO2-NA substrate after 30 days showed a sharp peak at 2θ = 12°, which could be assigned to the (001) plane of the PbI2 nanocrystal, indicating the decomposition of the MAPbI3 thin film. Interestingly, the PbI2 diffraction peak intensity of the (001) plane in MAPbI3 films coated on TiO2−CuInS2 QD-NA substrates was greatly reduced, indicating a more stable interface between MAPbI3 and CuInS2 QD-modified TiO2NAs. Our results suggest that the CuInS2 QD-modified TiO2NAs can effectively increase the stability of MAPbI3 films, which was in good agreement with the SEM data. The performances of TiO2-NA-based and TiO2−CuInS2 QD-NA-based solar cells in air for 30 days were characterized (Figure S7 and Table 2). After 30 days in air, the device based on TiO2-NAs showed a Voc of 0.73 V, a Jsc of 5.7 mA cm−2, and a PCE of 1.1%. In contrast, the corresponding TiO2−CuInS2

Figure 5. SEM images of MApbI3 films coated on (a) TiO2-NAs and (b) TiO2−CuInS2 QD-NAs; the corresponding MApbI3 film SEM images on (c) TiO2-NAs and (d) QD-modified TiO2-NAs under an air environment for 30 days.

on both TiO2-NA and QD-modified TiO2 nanoarray substrates showed smooth and compact surfaces, indicating a good crystallization process during annealing. After exposure to air for 30 days, the MAPbI3 on TiO2-NAs broken down into small particles with cracks on their surfaces. Compared to that on TiO2-NAs, MAPbI3 on the TiO2−CuInS2 QD-NAs substrate showed better air stability. The decomposition degree was greatly reduced with a more regular and smooth surface, indicating that TiO2−CuInS2 QD-NAs can effectively prevent MAPbI3 from decomposition. XRD patterns of MAPbI3 films on TiO2-NA and TiO2− CuInS2 QD-NA substrates were recorded to further corroborate the SEM results. Figure 6a shows the XRD patterns of MAPbI3 films on different substrates. For films on the TiO2NA substrate, the diffraction peaks at 2θ = 14, 24, 28, 32, 40, and 43° can be assigned to the (110), (202), (220), (310), 3435

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega

Article

Table 2. Device Performance of MAPbI3 Perovskite Solar Cells in Air for 30 days devices

Voc (V)

Jsc (mA cm−2)

FF

PCE (%)

TiO2-NAs/MAPbI3 TiO2−CuInS2 QD-NAs/MAPbI3

0.73 0.93

5.7 10.1

0.27 0.57

1.1 5.4

QD-NA-based device exhibited a Voc of 0.93 V, a Jsc of 10.1 mA cm−2, and a PCE of 5.4%. The PCE of the TiO2-NA-based device was only 12% of the original efficiency with a decreased fill factor of 0.27, for which the low stability of MAPbI3 should be responsible. Fortunately, the corresponding TiO2−CuInS2 QD-NA-based device showed better performance, which possessed 41% PCE of the original device. The results demonstrated that the CuInS2 QD-modified TiO2-NAs can effectively increase the stability of the solar cells.

mixture solution consisting of 60 mL of ethanol solution with Cu(Ac)2 (0.1 mmol), In(Ac)3 (0.1 mmol), CH 3 (CH 2 ) 16 CH 2 NH 2 (1.2 mmol), and CH 4 N 2 S (0.4 mmol).32 The reaction mixture was further heated at 160 °C with a growth time of 9 h to form TiO2−CuInS2 QD-NAs/ FTO thin film. The as-synthesized TiO2−CuInS2 QD-NAs/ FTO thin film was finally annealed at 450 °C for 30 min, forming the chalcopyrite CuInS2 QD-coated TiO2-NAs/FTO thin film. Fabrication and Characterization of Solar Cells. CH3NH3I was prepared by following a method reported previously.46 Methylamine (ethanol solution, 24 mL, 33%) together with 10 mL of 57% HI (water solution) was dissolved in 100 mL of absolute ethanol for at least 2 h in a nitrogen atmosphere to achieve crystallization of CH3NH3I. After that, 0.463 g of PbI2 and 0.447 g of CH3NH3I (mole ratio 1:3) were added into 3 mL of dimethylformamide solution to form the CH3NH3PbI3 precursor solution. The precursor solution was then spin-coated onto the FTO/TiO2−CuInS2 QD-NA thin film (1500 rpm, 40 s). After annealing at 100 °C for 30 min, the MAPbI3 active layer was formed. Afterward, a spiroMeOTAD (68 mM) chlorobenzene solution contains tertbutylpyridine (55 mM), and lithium bis(trifluoromethylsulfonyl)imide salt (9 mM) was prepared and spin-coated onto MAPbI3 at 2000 rpm for 40 s. Finally, a device (area of 0.12 cm2) was prepared by a 200 nm gold depositing onto the spiro-MeOTAD surface. The current−voltage (J−V) curves were tested under the AM 1.5G spectrum at 100 mW cm−2 (Newport Oriel). The EQE measurement was carried out to test the insight light−current conversion of devices using a Qtest station 500 instrument.



CONCLUSIONS In summary, solar cells based on CuInS2 QD-modified TiO2 nanoarrays were fabricated in this work. The solar cells based on TiO2−CuInS2 QD-NAs exhibited an improved PCE of 13.3%, much higher than that of solar cells using pure nanoarrays (8.2%). After 30 days in an air environment, the device based on TiO2−CuInS2 QD-NAs still possessed a PCE of 5.4%, which was 41% of the original device. The SEM and XRD results showed that the crystallinity, surface morphology, and interface for charge transfer of TiO2−CuInS2 QD-NAbased perovskites all remarkably improved, indicating the improved air stability of MAPbI3 films on TiO2−CuInS2 QDNAs. On the basis of this, the CuInS2 QD-modified nanoarrays provide a novel channel for the preparation of highperformance and high-stability perovskite devices.



EXPERIMENTAL SECTION Characterizations and Methods. UV−vis absorption spectra were measured to investigate the absorption properties (Agilent Cary 300). Time-resolved photoluminescence spectroscopic measurements (the obtained decay data was fitted by exponential functions with χ2 equal to 0.9−1.2) and PL spectra of the prepared films were measured to test the lifetime of the samples (Edinburgh FLS980). The X-ray diffraction patterns were recorded to measure the crystal characteristics (D/MAX 2500, Cu Kα radiation, λ = 1.5405 Å). Transmission electron microscopy (TEM) measurements (Tecnai G2 20 S-TWIN) and scanning electron microscopy (SEM) measurements (JEOL S-4800) were applied to test the morphologies of the samples. The MAPbI3 films were obtained by spin-coating the precursor solution on corresponding substrates at a speed of 1500 rpm. The films were then annealed at 100 °C for 30 min for further use. Preparation of TiO2−CuInS2 QD-NAs. The TiO2-NAs were prepared by following the method in previous literatures.45 First, a homogeneous solution was prepared in a 250 mL flask with 37% HCl and deionized water added (VHCl/Vwater = 1:1). The homogeneous solution was then shifted into a Teflon-lined chemical reaction vessel with impaction of FTO glass. Afterward, isopropyl titanate (1 mL) was added and the solution was sonication for 5 min. After heating at 180 °C for 90 min, the TiO2-NAs/FTO film was obtained, which was further annealed at 450 °C for 30 min to obtain TiO2-NAs. The TiO2-NAs/FTO film was modified under room temperature by soaking into a cysteine solution (3 × 10−4 M) for at least 3 days, which was then added into a reaction



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b03629.



Sample characterizations and absorption spectrum of MApbI3, TEM of pure CuInS2 QDs and QD-modified TiO2-NA, and size distribution; and cross-sectional images of solar cells and films and J−V curves of TiO2-NA-based and TiO2−CuInS2 QD-NA-based solar cells in air for 30 days (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Feng Gao: 0000-0001-5273-7627 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are grateful to the Shaoyang University for financial support. 3436

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega



Article

C under Continuous 1 Sun Illumination in Ambient Air. Adv. Mater. 2018, 30, No. 1801010. (18) Zhang, J.; Tong, T.; Zhang, L.; Li, X.; Zou, H.; Yu, J. Enhanced Performance of Planar Perovskite Solar Cell by Graphene Quantum Dot Modification. ACS Sustainable Chem. Eng. 2018, 6, 8631−8640. (19) Donaldson, L. New technique improves perovskite solar cells. Mater. Today 2017, 20, 485. (20) Tanaka, Y.; Lay, S.; et al. Fabrication of Mesoporous Titania Nanoparticles with Controlled Porosity and Connectivity for Studying the Photovoltaic Properties in Perovskite Solar Cells. ChemNanoMat 2018, 4, 394−400. (21) Liang, X.; Cheng, Y.; Xu, X.; Dong, R.; Li, D.; Zhou, Z.; Wei, R.; Dong, G.; Tsang, S.-W.; Ho, J. C. Enhanced performance of perovskite solar cells based on vertical TiO2 nanotube arrays with full filling of CH3NH3PbI3. Appl. Surf. Sci. 2018, 451, 250−257. (22) Aeineh, N.; Castro-Mendez, A. F.; Rodriguez-Canto, P. J.; Abargues, R.; Hassanabadi, E.; Suarez, I.; Behjat, A.; Ortiz, P.; Martinez-Pastor, J. P.; Mora-Sero, I. Optical Optimization of the TiO2 Mesoporous Layer in Perovskite Solar Cells by the Addition of SiO2 Nanoparticles. ACS Omega 2018, 3, 9798−9804. (23) Wang, Z.; Fang, J.; Mi, Y.; Zhu, X.; Ren, H.; Liu, X.; Yan, Y. Enhanced performance of perovskite solar cells by ultraviolet-ozone treatment of mesoporous TiO2. Appl. Surf. Sci. 2018, 436, 596−602. (24) Bush, K. A.; Palmstrom, A. F.; Yu, Z. J.; Boccard, M.; Cheacharoen, R.; Mailoa, J. P.; McMeekin, D. P.; Hoye, R. L. Z.; Bailie, C. D.; Leijtens, T.; Peters, I. M.; Minichetti, M. C.; Rolston, N.; Prasanna, R.; Sofia, S.; Harwood, D.; Ma, W.; Moghadam, F.; Snaith, H. J.; Buonassisi, T.; Holman, Z. C.; Bent, S. F.; McGehee, M. D. 23.6%-efficient monolithic perovskite/silicon tandem solar cells with improved stability. Nat. Energy 2017, 2, No. 17009. (25) Bi, D.; Yi, C.; Luo, J.; Décoppet, J.-D.; Zhang, F.; Zakeeruddin, S. M.; Li, X.; Hagfeldt, A.; Grätzel, M. Polymer-templated nucleation and crystal growth of perovskite films for solar cells with efficiency greater than 21%. Nat. Energy 2016, 1, No. 16142. (26) Wang, Q.; Jin, Z.; Chen, D.; Bai, D.; Bian, H.; Sun, J.; Zhu, G.; Wang, G.; Liu, S. F. μ-Graphene Crosslinked CsPbI3 Quantum Dots for High Efficiency Solar Cells with Much Improved Stability. Adv. Energy Mater. 2018, 8, No. 1800007. (27) Zheng, H.; Liu, G.; Zhu, L.; Ye, J.; Zhang, X.; Alsaedi, A.; Hayat, T.; Pan, X.; Dai, S. The Effect of Hydrophobicity of Ammonium Salts on Stability of Quasi-2D Perovskite Materials in Moist Condition. Adv. Energy Mater. 2018, 8, No. 1800051. (28) Song, Z.; Abate, A.; Watthage, S. C.; Liyanage, G. K.; Phillips, A. B.; Steiner, U.; Graetzel, M.; Heben, M. J. Perovskite Solar Cell Stability in Humid Air: Partially Reversible Phase Transitions in the PbI2-CH3NH3I-H2O System. Adv. Energy Mater. 2016, 6, No. 1600846. (29) Hu, Z.; Miao, J.; Li, T.; Liu, M.; Murtaza, I.; Meng, H. Reduced interface losses in inverted perovskite solar cells by using a simple dual-functional phenanthroline derivative. Nano Energy 2018, 43, 72− 80. (30) Shi, J.; Xu, X.; Li, D.; Meng, Q. Interfaces in perovskite solar cells. Small 2015, 11, 2472−2486. (31) Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T. B.; Duan, H. S.; Hong, Z.; You, J.; Liu, Y.; Yang, Y. Photovoltaics. Interface engineering of highly efficient perovskite solar cells. Science 2014, 345, 542−546. (32) Yue, W.; Wu, F.; Liu, C.; Qiu, Z.; Cui, Q.; Zhang, H.; Gao, F.; Shen, W.; Qiao, Q.; Wang, M. Incorporating CuInS2 quantum dots into polymer/oxide-nanoarray system for efficient hybrid solar cells. Sol. Energy Mater. Sol. Cells 2013, 114, 43−53. (33) Maier, E.; Rath, T.; Haas, W.; Werzer, O.; Saf, R.; Hofer, F.; Meissner, D.; Volobujeva, O.; Bereznev, S.; Mellikov, E.; Amenitsch, H.; Resel, R.; Trimmel, G. CuInS 2Poly(3-(ethyl-4-butanoate)thiophene) nanocomposite solar cells: Preparation by an in situ formation route, performance and stability issues. Sol. Energy Mater. Sol. Cells 2011, 95, 1354−1361. (34) Lefrançois, A.; Luszczynska, B.; Pepin-Donat, B.; Lombard, C.; Bouthinon, B.; Verilhac, J. M.; Gromova, M.; Faure-Vincent, J.;

REFERENCES

(1) Zhao, Y.; Zhu, K. Efficient planar perovskite solar cells based on 1.8 eV band gap CH3NH3PbI2Br nanosheets via thermal decomposition. J. Am. Chem. Soc. 2014, 136, 12241−12244. (2) Zhao, Y.; Tan, H.; Yuan, H.; Yang, Z.; Fan, J. Z.; Kim, J.; Voznyy, O.; Gong, X.; Quan, L. N.; Tan, C. S.; Hofkens, J.; Yu, D.; Zhao, Q.; Sargent, E. H. Perovskite seeding growth of formamidinium-lead-iodide-based perovskites for efficient and stable solar cells. Nat. Commun. 2018, 9, No. 1607. (3) Liu, M.; Johnston, M. B.; Snaith, H. J. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature 2013, 501, 395−398. (4) Chen, Q.; Zhou, H.; Hong, Z.; Luo, S.; Duan, H. S.; Wang, H. H.; Liu, Y.; Li, G.; Yang, Y. Planar heterojunction perovskite solar cells via vapor-assisted solution process. J. Am. Chem. Soc. 2014, 136, 622−625. (5) Qin, P.; Paek, S.; Dar, M. I.; Pellet, N.; Ko, J.; Gratzel, M.; Nazeeruddin, M. K. Perovskite solar cells with 12.8% efficiency by using conjugated quinolizino acridine based hole transporting material. J. Am. Chem. Soc. 2014, 136, 8516−8519. (6) 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. (7) Kohnehpoushi, S.; Nazari, P.; Nejand, B. A.; Eskandari, M. MoS2: a two-dimensional hole-transporting material for highefficiency, low-cost perovskite solar cells. Nanotechnology 2018, 29, No. 205201. (8) Song, J.; Xu, X.; Wu, J.; Lan, Z. Low-temperature solutionprocessing high quality Nb-doped SnO2 nanocrystals-based electron transport layers for efficient planar perovskite solar cells. Funct. Mater. Lett. 2018, 12, No. 1850091. (9) Febriansyah, B.; Koh, T. M.; John, R. A.; Ganguly, R.; Li, Y.; Bruno, A.; Mhaisalkar, S. G.; England, J. Inducing Panchromatic Absorption and Photoconductivity in Polycrystalline Molecular 1D Lead-Iodide Perovskites through π-Stacked Viologens. Chem. Mater. 2018, 30, 5827−5830. (10) Burwig, T.; Franzel, W.; Pistor, P. Crystal Phases and Thermal Stability of Co-evaporated CsPbX3 (X = I, Br) Thin Films. J. Phys. Chem. Lett. 2018, 9, 4808−4813. (11) Meng, F. L.; Wu, J. J.; Zhao, E. F.; Zheng, Y. Z.; Huang, M. L.; Dai, L. M.; Tao, X.; Chen, J. F. High-efficiency near-infrared enabled planar perovskite solar cells by embedding upconversion nanocrystals. Nanoscale 2017, 9, 18535−18545. (12) Rakita, Y.; Cohen, S. R.; Kedem, N. K.; Hodes, G.; Cahen, D. Mechanical properties of APbX 3 (A = Cs or CH3NH3; X = I or Br) perovskite single crystals. MRS Commun. 2015, 5, 623−629. (13) Day, J.; Senthilarasu, S.; Mallick, T. K. Improving spectral modification for applications in solar cells: A review. Renewable Energy 2019, 132, 186−205. ̇ (14) Kırbıyık, Ç .; Kara, K.; Kara, D. A.; Yiğit, M. Z.; Istanbullu, B.; Can, M.; Sariciftci, N. S.; Scharber, M.; Kuş, M. Enhancing the cTiO2 based perovskite solar cell performance via modification by a serial of boronic acid derivative self-assembled monolayers. Appl. Surf. Sci. 2017, 423, 521−527. (15) Xing, G.; Wu, B.; Chen, S.; Chua, J.; Yantara, N.; Mhaisalkar, S.; Mathews, N.; Sum, T. C. Interfacial Electron Transfer Barrier at Compact TiO2/CH3NH3PbI3 Heterojunction. Small 2015, 11, 3606−3613. (16) Stolterfoht, M.; Wolff, C. M.; Márquez, J. A.; Zhang, S.; Hages, C. J.; Rothhardt, D.; Albrecht, S.; Burn, P. L.; Meredith, P.; Unold, T.; Neher, D. Visualization and suppression of interfacial recombination for high-efficiency large-area pin perovskite solar cells. Nat. Energy 2018, 3, 847−854. (17) Seo, S.; Jeong, S.; Bae, C.; Park, N. G.; Shin, H. Perovskite Solar Cells with Inorganic Electron- and Hole-Transport Layers Exhibiting Long-Term (approximately 500 h) Stability at 85 degrees 3437

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438

ACS Omega

Article

Pouget, S.; Chandezon, F.; Sadki, S.; Reiss, P. Enhanced charge separation in ternary P3HT/PCBM/CuInS2 nanocrystals hybrid solar cells. Sci. Rep. 2015, 5, No. 7768. (35) Chen, C.; Li, C.; Li, F.; Wu, F.; Tan, F.; Zhai, Y.; Zhang, W. Efficient perovskite solar cells based on low-temperature solutionprocessed (CH3NH3)PbI3 perovskite/CuInS2 planar heterojunctions. Nanoscale Res. Lett. 2014, 9, 457. (36) Lv, M.; Zhu, J.; Huang, Y.; Li, Y.; Shao, Z.; Xu, Y.; Dai, S. Colloidal CuInS2 Quantum Dots as Inorganic Hole-Transporting Material in Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 17482−17488. (37) Sun, C.; Cevher, Z.; Zhang, J.; Gao, B.; Shum, K.; Ren, Y. Onepot synthesis and characterization of chalcopyrite CuInS2 nanoparticles. J. Mater. Chem. A 2014, 2, 10629−10633. (38) Vásquez, G. C.; Karazhanov, S. Z.; Maestre, D.; Cremades, A.; Piqueras, J.; Foss, S. E. Oxygen vacancy related distortions in rutile $/mathrm{Ti}{/mathrm{O}}_{2}$ nanoparticles: A combined experimental and theoretical study. Phys. Rev. B 2016, 94, No. 235209. (39) Im, J.-H.; Kim, H.-S.; Park, N.-G. Morphology-photovoltaic property correlation in perovskite solar cells: One-step versus twostep deposition of CH3NH3PbI3. APL Mater. 2014, 2, No. 081510. (40) Leijtens, T.; Lauber, B.; Eperon, G. E.; Stranks, S. D.; Snaith, H. J. The Importance of Perovskite Pore Filling in Organometal Mixed Halide Sensitized TiO2-Based Solar Cells. J. Phys. Chem. Lett. 2014, 5, 1096−1102. (41) Im, J.-H.; In-Hyuk, J.; Norman, P.; Michael, G. T.; Nam-Gyu, P. Growth of CH3NH3PbI3 cuboids with controlled size for highefficiency perovskite solar cells. Nat. Nanotechnol. 2014, 9, 927−932. (42) Ahmed, M. I.; Hussain, Z.; Khalid, A.; Amin, H. M. N.; Habib, A. Absorption enhancement in CH3NH3PbI3solar cell using a TiO2/ MoS2nanocomposite electron selective contact. Mater. Res. Express 2016, 3, No. 045022. (43) Liu, Y.; Collins, L.; Proksch, R.; Kim, S.; Watson, B. R.; Doughty, B.; Calhoun, T. R.; Ahmadi, M.; Ievlev, A. V.; Jesse, S.; Retterer, S. T.; Belianinov, A.; Xiao, K.; Huang, J.; Sumpter, B. G.; Kalinin, S. V.; Hu, B.; Ovchinnikova, O. S. Chemical nature of ferroelastic twin domains in CH3NH3PbI3 perovskite. Nat. Mater. 2018, 17, 1013−1019. (44) Kim, Y. C.; Jeon, N. J.; Noh, J. H.; Yang, W. S.; Seo, J.; Yun, J. S.; Ho-Baillie, A.; Huang, S.; Green, M. A.; Seidel, J.; Ahn, T. K.; Seok, S. I. Beneficial Effects of PbI2Incorporated in Organo-Lead Halide Perovskite Solar Cells. Adv. Energy Mater. 2016, 6, No. 1502104. (45) Gao, F.; Dai, H.; Pan, H.; Chen, Y.; Wang, J.; Chen, Z. Performance enhancement of perovskite solar cells by employing TiO2 nanorod arrays decorated with CuInS2 quantum dots. J. Colloid Interface Sci. 2018, 513, 693−699. (46) Chen, Q.; Zhou, H.; Song, T. B.; Luo, S.; Hong, Z.; Duan, H. S.; Dou, L.; Liu, Y.; Yang, Y. Controllable self-induced passivation of hybrid lead iodide perovskites toward high performance solar cells. Nano Lett. 2014, 14, 4158−4163.

3438

DOI: 10.1021/acsomega.8b03629 ACS Omega 2019, 4, 3432−3438