Electrosprayed Polymer-Hybridized Multidoped ZnO Mesoscopic

Aug 21, 2018 - *E-mail: [email protected]. ACS AuthorChoice - This is an open access article published under an ACS AuthorChoice License, which permi...
0 downloads 0 Views 3MB 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 2018, 3, 9648−9657

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

Electrosprayed Polymer-Hybridized Multidoped ZnO Mesoscopic Nanocrystals Yield Highly Efficient and Stable Perovskite Solar Cells Khalid Mahmood,*,† Muhammad Taqi Mehran,‡ Faisal Rehman,† Muhammad Shahzad Zafar,§ Syed Waqas Ahmad,† and Rak-Hyun Song∥

ACS Omega 2018.3:9648-9657. Downloaded from pubs.acs.org by 185.223.161.87 on 08/21/18. For personal use only.



Department of Chemical & Polymer Engineering, University of Engineering & Technology Lahore, Faisalabad Campus, 31/2 Km. Khurrianwala, Makkuana By-Pass, Faisalabad 38000, Pakistan ‡ University of Science and Technology (UST), 217-Gajeong-ro, Yuseong-gu, Daejeon 34113, Republic of Korea § SKKU Advanced Institute of Nano-Technology (SAINT), Sungkyunkwan University (SKKU), Seobu-ro, Jangan-gu, Suwon-si, Gyeonggi-do 16419, Republic of Korea ∥ New and Renewable Energy Research Division, Korea Institute of Energy Research (KIER), 152-Gajeong ro, Yuseong gu, Daejeon 34129, South Korea S Supporting Information *

ABSTRACT: Solid-state perovskite solar cells have been expeditiously developed since the past few years. However, there are a number of open questions and issues related to the perovskite devices, such as their long-term ambient stability and hysteresis in current density−voltage curves. We developed highly efficient and hysteresis-less perovskite devices by changing the frequently used TiO2 mesoscopic layer with polymer-hybridized multidoped ZnO nanocrystals in a common n−i−p structure for the first time. The gradual adjustment of ZnO conduction band position using singleand multidopant atoms will likely enhance the power conversion efficiency (PCE) from 8.26 to 13.54%, with PCEmax = 15.09%. The highest PCEavg of 13.54% was demonstrated by 2 atom % boron and 6 atom % fluorine co-doped (B, F:ZnO) nanolayers (using optimized film thickness of 160 nm) owing to their highest conductivity, carrier concentration, optical transmittance, and band-gap energy compared to other doped films. We also successfully apply a fine polyethylenimine thin layer on the doped ZnO nanolayers, causing the reduction in work function and overall demonstrating the enhancement in PCE from ∼10.86% up to 16.20%. A polymer-mixed electron-transporting layer demonstrates the remarkable PCEmax of 20.74% by decreasing the trap sites in the oxide layer that probably reduces the chances of carrier interfacial recombination originated from traps and thus improves the device performance. Particularly, we produce these electron-rich multidoped ZnO nanolayers via electrospray technique, which is highly suitable for the future development of perovskite solar cells. severely hinders the fabrication of large flexible devices.20 Thus, in the near future, the development of perovskite devices at low temperature would be an encouraging way and movement. Nanostructures of ZnO are highly suitable alternative materials to other mesoscopic ETLs for perovskite devices owing to their suitable energy levels, ease of fabrication at low temperatures, and relatively high electron mobility.21−24 Only a handful of studies have evaluated the effect of ZnO-based ETLs of different morphologies.25,26 ZnO nanorods, for example, offer efficient charge collection compared to TiO2 nanorods, owing to their more rapid charge-carrying

1. INTRODUCTION The intensive interest in perovskite solar cells has led to a very rapid progress in cell performance compared to any other photovoltaic technology.1−3 The advances in power conversion efficiency (PCE) have been achieved by optimizing cell configuration and layer fabrication processes.4−9 PCE as high as 20.7% has been achieved using a planar structure, while a PCE of 22.1% has been obtained via a mesoscopic structure.10 For most of the device architectures, a thin film of TiO2 has been mostly used as an electron-transporting layer (ETL) in both planar and mesoscopic perovskite cells.11−16 New types of ETLs (e.g., WO3, Zn2SnO4, and SrTiO3) have also been explored to produce efficient perovskite cells.17−19 However, to develop up-to-date efficient perovskite solar cells using these mesoscopic ETLs along with thin hole-blocking layers of the same materials, high-temperature sintering is required. This © 2018 American Chemical Society

Received: June 21, 2018 Accepted: August 7, 2018 Published: August 21, 2018 9648

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

Figure 1. Schematic illustration of low-temperature and fully solution-processed electrosspray-deposited (a) pure ZnO and multidoped ZnO nanocrystals (the inset is the magnified view of mesoporous nanocrystals); (b) PEI-coated ZnO and multidoped ZnO nanolayers; and (c) PEImixed pure ZnO and multidoped ZnO nanocrystals (the inset is the magnified view of PEI-blended nanocrystals). (d (i)−(iii)) Schematic illustration of device architecture based on the above oxide nanolayers showing the electron trapping due to the voids between ZnO nanocrystals.

behavior.27 In addition, doping the ZnO by substituting the metallic atoms into its lattice not only improves its conductivity, but also facilitates the shift of the Fermi level toward conduction band direction, which will eventually enhance the performance of dye-sensitized solar cells (DSSCs), but the efficiency is not satisfactory yet.28−30 Moreover, the alternation in the surface morphology of oxide layer will greatly influence the device performance because the interface between the perovskite and the oxide layer plays an important role in charge separation and crystal growth.17,25 Reports are also available on the bulk modification and interfacial treatment of oxide layer to enhance the collection and injection of charges and also to reduce the carrier recombination at the same time.31−33 However, the efficiency of these liquid-based devices is not comparable to that of perovskite solar cells, and the main drawbacks of DSSCs (instability of liquid electrolyte, costly dye and catalyst, etc.) have not been solved yet. Recently, researchers6,22,34−36 have also presented a route of surface modification and improving the electron-transport properties to enhance the performance of perovskite devices by doping metallic elements (e.g., Al, Yt, Mg, Nb) into the nanostructures of ZnO and TiO2. However, compared to the advanced TiO2 ETL-based devices, lower device performances were achieved even in these devices. Motivated by the desire to evaluate the synergetic influence of dopant on the nanostructures surface of oxide layer and to tune the energy levels to obtain an efficient perovskite solar cell, we introduced first the substitution of multidopant into the ZnO lattice and studied their effect on the perovskite device performance. In our previous reports, we investigated the influence of single and co-dopants (with optimized dopant amounts) on the performance of transparent electrodes (transparent conductive oxides) by enhancing the electrical conductivity, optical transmission, stability, and possibility for future commercialization.37−39 We had found that the additional electrons could be generated using doping atoms, which will enhance the carrier density, mobility, conductivity, and surface stability because such findings of multidoped ZnO films have an excellent technological aspects and would also add to basic studies.

The use of conjugated polyelectrolytes offers the advantage of changing the work function of the oxide layer to boost the cell performance in an efficient way.40−43 In this method, the surface of oxide layer is finely coated using a material which physically or chemically adsorbed on it; the surface transformer is selected in a sense to generate strong surface dipoles that bring a shift in vacuum level and thus alter the work function of the oxide ETLs.44 Very recently, polyethylenimine (PEI) or polyethyleneimine-ethoxylated polymers have been investigated to modify indium tin oxide (ITO) work function by forming a dipole layer at the ITO−PEI interface and to yield efficient solar cells.45 These polymers can easily form solutions with 2-methoxyethanol or water (known as environmentally friendly solvents) and can be simply handled in ambient atmosphere. Moreover, their ease of handling and inexpensive nature make them well suited for printed electronics with rollto-roll large-area production.44 The key issue with this method is that the existence of trap sites in the bulk of the ZnO ETLs still forbid the transport of electrons caused by the empty spaces between ZnO nanocrystals (NCs). A different methodology is to form the composite of polymers and oxide layers (ZnO or TiO2) by mixing them together before deposition in the developing stage. A simple polymer-mixed composite single-layer ETL is more functional and easy to process compared to a complex bilayer film. However, only a few studies have been conducted where a composite ETL has been investigated rather than a bilayer ETL (oxide layer coated with polymer) in inverted polymer cells. For instance, only in polymer solar cells, composites of poly(ethylene oxide), 46 poly(ethylene glycol), 47 poly(vinylpyrrolidone),44 and PEI48 with ZnO films have been reported to construct polymer-mixed ETLs. However, such mixed polymer composite ETLs have not been investigated yet for perovskite solar cells. In this work, we have introduced a low-temperature and easy-to-operate full-solution-processed electrospraying technique for the development of mesoporous pure ZnO and multidoped ZnO nanocrystals. We constantly increase the position of conduction of the ZnO ETLs by doping with diverse combination of metallic and nonmetallic atoms. By 9649

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

substrate, solution concentration, substrate temperature, and applied voltage as summarized in Table S1) as reported elsewhere.37−39 As the film morphology is critically influenced by the dopant concentrations and process parameters. The grain size of the doped ETLs was gradually reduced as the dopant amount was increased in the precursor solution, which will eventually affect their optical and electrical characteristics and hence the device performance.37−39 It was also examined that the additions of dopant atoms have strongly influence the surface roughness of ETLs. It was also examined that when dopant atoms were substituted into pure ZnO films, a substantial reduction in surface roughness was observed.37−39 The surface modification and morphology refinement of the oxide ETLs is a better route to refine the morphology of light harvester and transfer of electrons in the mesostructured perovskite solar cells. The electrical and optical properties of the pure ZnO and doped ZnO films at the optimized dopant concentrations are plotted in Figure 2 and summarized in

changing the position of oxide’s conduction band and surface morphology, we are capable of significantly improving the open-circuit voltage (VOC) to achieve the maximum values of efficiency reported so far for doped ZnO-based perovskite devices. Previously, we have studied the transparent electrodes in the form of 2 atom % boron-doped ZnO (B:ZnO), 2 atom % tantalum-doped ZnO (Ta:ZnO), 2 atom % boron and 6 atom % fluorine co-doped (B, F:ZnO), and 2 atom % tantalum and 6 atom % nitrogen co-doped (Ta, N:ZnO) thin films.37−39 We have constructed the perovskite solar cells using these thin films and investigated their effect on the cell performance at optimized dopant concentrations reported previously. The B:ZnO films with highest conductivity, carrier concentration, optical transmittance, and band-gap energy will likely produce a PCEavg of 13.54% and a PCEmax of 15.09% among other doped films. We have substantially reduced the work functions of these films by coating them with a thin layer of PEI, which results in remarkable improvements of PCE of 16.2% in hysteresis-free mesoscopic perovskite solar cells. A remarkable PCEmax of 20.74% was achieved when a polymer-mixed ZnO ETL is used. To the best of the author’s knowledge, this is the first study that employs co-doped semiconducting nanostructured films in combination with polymer as an ETL to obtain hysteresis-less and very efficient perovskite solar cells. Furthermore, the obtained PCEmax of 20.74% exceeds the previously reported highest efficiency for the perovskite solar cells based on ZnO ETLs.49−53

2. RESULTS AND DISCUSSION 2.1. General Scheme for Tuning Nanolayer Properties and Device Architecture. Figure 1a shows the schematic illustration of low-temperature, easy-to-operate, and fully solution-processed electrostatic spray deposited pure ZnO and multidoped ZnO mesoporous ETLs composed of fine nanocrystals (as shown in the inset of Figure 1a). Highly mesoporous ZnO ETLs could be easily achieved by providing a constant low heating (150 °C) at the bottom of the substrate during the entire deposition process. The Brunauer−Emmett− Teller (BET) surface area increased from 7.3 m2/g for pure ZnO to 27.5 m2/g for B, F:ZnO ETL. However, these ETLs have numerous traps left in the bulk of ZnO, and doped ZnO nanocrystals (NCs) lead to electron trapping due to the voids between ZnO NCs (Figure 1d(i)). Coating a thin layer of conjugated polyelectrolyte, such as polymer (PEI), on top of the ETL (as illustrated in Figure 1b) can reduce the work function of ETL and will also reduce the surface traps, which enhances electronic pairing of ETL/perovskite layer and thus boost the device performance (as explained in Figure 1d(ii)). The functionality of these films was further enhanced by mixing the ZnO solution with PEI and the blended solution was electrosprayed directly onto the heated substrates, as illustrated in Figure 1c. The PEI is finely dispersed in the entire film (inset of Figure 1c), which decreased the surface roughness and trap states for ZnO film, thereby possibly improving the physical contact and stimulating the strong dipoles movements between the perovskite absorber layer and ITO conducting substrate, generating the cells with improved performance (Figure 1d(iii)). Figure S1 displays the surface nanostructures of pure ZnO and multidoped ZnO (such as B:ZnO, Ta:ZnO, B, F:ZnO, and Ta, N:ZnO) mesoscopic ETLs formed by electrospraying at the optimized dopant concentrations and the process parameters (flow rate, distance between the nozzle and the

Figure 2. Plots of (a) carrier concentration and electrical resistivity and (b) band-gap energy and optical transmittance (at 520 nm) for five different pure ZnO and doped ZnO nanolayers.

Table S2. The B, F:ZnO nanolayer demonstrates the lowest resistivity (9.70 × 10−5 Ω cm), highest optical transmittance (99.8%), maximum carrier concentration (3.41 × 1021 cm−3), and a band-gap energy of 3.42 eV compared to other pure and doped ZnO films. A side-view scanning electron microscopy (SEM) image of full cell based on B, F:ZnO nanolayer as an electron-transporting material is shown in Figure S1f. 2.2. Device Structure and Energy-Level Diagram. The schematic demonstration of the perovskite device architecture 9650

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

higher than other doped films as well. Actually, doping of ZnO enhances the electron density, which results in an increase in JSC and PCE as well.37−39 Moreover, the rise in VOC (from 860 to 960 mV) after doping originated from an increase in Fermi energy levels and electron density, and reduced recombination losses, which decreases the restriction to the transfer of electrons.37−39 Furthermore, continuous raising of the conduction band edge using various combinations of dopants causes a 0.22 eV shift in the conduction band compared to pure ZnO films, which will eventually improve the VOC and enhance the faster transfer of electrons.49 The devices fabricated with thicker pure ZnO and doped ZnO nanolayers (240 and 300 nm rather than 160 nm) showed lower VOC and JSC (in the reverse direction only) (Figure S3), showing that thinner nanolayers are better for highly efficient devices. The enhanced performance of the nanolayers with small thickness is attributed to the complete infiltration of the perovskite absorber into the oxide layer and to the improvement of charge-transfer characteristics.22 Figure S2 shows the EQE spectra of the perovskite devices with pure and doped ZnO nanolayers. A plateau of over 85% EQE across the visible region is clearly seen, proposing the full conversion due to the improved absorbance of perovskite absorber for the devices based on B, F:ZnO nanolayers. The solar cells exhibit a considerable hysteresis-less behavior when altering the scan direction, as shown in Figure 4a, and the related photovoltaic parameters are listed in Table S3. As exhibited in Figure 4b,c, device performance in terms of JSC, VOC, FF, and PCEavg was enhanced upon the addition of dopants into ZnO. The devices based on B, F:ZnO nanolayers exhibit the best performance compared to other nanolayers. The variation in the performances of 50 different perovskite devices based on pure ZnO and B, F:ZnO nanolayers is presented in Figure 4d, demonstrating a PCEavg of 8.26 and 13.54%, respectively. 2.4. Device Performance of Doped ZnO Films with Unoptimized Dopant Concentrations. We have also constructed perovskite devices using highly doped ZnO nanolayer to evaluate their performance with the devices based on optimized dopant concentration. It was observed that the devices based on highly doped ZnO films have low device performance (see Figure S4 and Table S4) compared to the optimized ones (as shown in Figure 4a). Actually, at higher dopant amounts, film morphology gets destroyed, which increases the surface roughness.37−39 Furthermore, the substitution of heavily doped atoms into the ZnO sites will result in lower carrier concentration and less conductive and transparent films,37−39 which will eventually affect the device performance in terms of low-efficient ones, as can be seen in Figure S4. 2.5. Device Performance for PEI-Coated Pure and Doped ZnO Films. We have now spin-coated the bestperforming pure ZnO and B, F:ZnO nanocrystals using a thin layer of PEI, keeping in mind the strong interaction between nonpolar surface of ZnO and PEI,45 to achieve a fine coating that reduces the work function of ZnO further.26 The schematic representation of the energy-level diagram based on the PEI-coated B, F:ZfunO nanolayer is shown in Figure 3b. We observe a 0.38 eV shift (measured by Kelvin probe) in work function between PEI-uncoated B, F:ZnO and PEIcoated B, F:ZnO nanolayer, caused by PEI coating. These findings also demonstrate that PEI fine coating reduces the work function of B, F:ZnO ETL from 4.16 to 3.78 eV.

(glass/ITO/ZnO or doped ZnO nanolayers/MAPbI(3−x)Clx/ spiro-OMeTAD/Ag) and the energy-level description of the corresponding materials (measured by Kelvin probe) are shown in Figure 3a,b, respectively. The incorporation of

Figure 3. (a) Illustration of device architecture with PEI-coated and PEI-mixed ZnO and doped ZnO nanolayers and (b) energy-level diagram of the corresponding devices with five various types of pure and doped ZnO nanolayers. Continuous raising of the conduction band edge using various combinations of dopants causes a 0.22 eV shift in the conduction band compared to the pure ZnO films, and also the conduction band edge of B, F:ZnO nanolayer decreased to 0.38 eV owing to dipole interaction by the PEI layer.

dopant atoms into ZnO lattice produced dopant-substituted crystal lattice in ZnO, allowing ZnO ETLs to have higher values of conduction band, as illustrated in Figure 3b. It is important to note that the electronic structure of ZnO can be tuned using doping, which results in increase of the band gap. The single-doped (B:ZnO and Ta:ZnO) ZnO nanolayers do not contribute significantly to the rise of conduction band edge. Therefore, co-doped ZnO nanolayers, such as Ta, N:ZnO and B, F:ZnO, are also employed to raise the conduction band prominently, and a significant shift of 0.22 eV is caused by B, F:ZnO nanolayer compared to pure ZnO. The continuous elevation of conduction band edge will increase the VOC for the perovskite cell and suppressed the recombination at the doped ZnO/perovskite interface.26 2.3. Device Performance of Pure and Doped ZnO Nanolayers with Optimized Dopant Concentrations. In Figures 4a and S2, we show the J−V plots and external quantum efficiency (EQE) of the perovskite devices with pure ZnO and doped ZnO nanolayers with film thickness of 160 nm. The photovoltaic parameters, such as open-circuit voltage (VOC), short-circuit current density (JSC), fill factor (FF), and power conversion efficiency (PCE) of the related cells, are outlined in Table 1. The cells with B, F:ZnO nanolayer show the best JSC of 20.75 mA/cm2, FF of 0.68, VOC of 960 mV, average PCE (PCEavg) of 13.54%, and maximum PCE (PCEmax) of 15.09%, which is much improved compared to the devices based on pure ZnO nanolayer with similar film thickness (PCEavg of 8.26% and PCEmax of 9.16%) and also 9651

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

Figure 4. (a) J−V plots of five various types of pure and doped ZnO nanolayers in different scanning directions at a rate of 0.01 V/s. Plots of (b) JSC and FF and (c) VOC and PCEavg for ZnO and four different dopant schemes. (d) Bar graphs demonstrating the changes in PCE for 50 separate devices with pure ZnO and B, F:ZnO nanolayers.

in Figure S5a, the representative peak (N 1s), which is collected from the nitrogen element in the PEI polymer (see Figure 3b), was clearly seen for PEI-coated B, F:ZnO nanolayer and PEI-coated ITO substrate samples around 400 eV. The presence of boron (B) and fluorine (F) peaks in Figure S5a confirms the successful substitution of B and F atoms into the ZnO lattice. These findings confirm that the PEI layer was essentially coated on the B, F:ZnO nanolayer surface. Especially, we observe that the positions of O1s and Zn2p peaks in the B, F:ZnO nanolayer were moved in the presence of PEI layer (Supporting Information, Figure S5c,d), which shows that the PEI layer definitely disturbs the ZnO electronic structure. The PEI coating covering the B, F:ZnO nanolayer surface results in the increase of JSC rise from 20.75 to 21.85 mA/cm2, FF from 68 to 70%, and VOC from 960 to 990 mV, as listed in Table 2 and exhibited in Figure S6, demonstrating a PCEavg of 15.14% (PCEmax = 16.20%). Thus,

Table 1. Device Parameters (in the Reverse Scan Direction Only) with Pure and Doped ZnO Nanolayers as ETLs under 1 sun Illumination (AM 1.5G, 100 mW/cm2) ETL type

JSC (mA/cm2)

VOC (mV)

FF (%)

pure ZnO B:ZnO Ta:ZnO Ta, N:ZnO B, F:ZnO

16.0 18.0 18.5 20.0 21.0

860 925 930 945 960

61 63 65 66 68

PCEavg (%)

PCEmax (%)

± ± ± ± ±

9.16 11.49 12.67 13.99 15.09

8.39 10.48 11.18 12.47 13.70

0.15 0.13 0.12 0.14 0.13

We have also confirmed whether the PEI thin layer was actually coated onto the ZnO surface or not since it has a thickness of only 2−3 nm. To prove the successful coverage of ZnO surface with PEI layer, X-ray photoelectron spectroscopy (XPS) was performed, as exhibited in Figure S5. As presented

Table 2. Device Parameters (in Both Forward and Reverse Scans) with PEI-Coated Pure ZnO and B, F:ZnO Nanolayers Having Optimal ETL Thickness of 160 nm film type

scan mode

JSC (mA/cm2)

VOC (mV)

FF (%)

pure ZnO

reverse forward average reverse forward average

17.0 16.8 16.9 22.0 21.70 21.85

895 895 895 990 990 990

63 63 63 70 70 70

B, F:ZnO

9652

PCEavg (%)

PCEmax (%)

± ± ± ± ± ±

10.95 10.77 10.86 16.28 16.13 16.20

9.58 9.47 9.52 15.24 15.03 15.14

0.12 0.14 0.13 0.11 0.10 0.10

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

Figure 5. J−V plots for the champion devices with (a) PEI-coated and (b) PEI-mixed ZnO and B, F:ZnO nanolayers using different sweep directions and bar graphs exhibiting the alteration in PCEavg for 50 individual devices with (c) PEI-coated and (d) PEI-mixed pure ZnO and B, F:ZnO nanolayers.

Table 3. Device Parameters (in Different Scanning Directions) with PEI-Mixed Pure ZnO and B, F:ZnO Nanolayers Having Optimal ETL Thickness of 160 nm film type

scan mode

JSC (mA/cm2)

VOC (mV)

FF (%)

pure ZnO

reverse forward average reverse forward average

17.5 17.3 17.4 23.0 22.50 22.75

923 923 923 1000 1000 1000

67 67 67 75 75 75

B, F:ZnO

PCEavg (%)

PCEmax (%)

± ± ± ± ± ±

13.04 12.78 12.91 20.83 20.65 20.74

10.82 10.69 10.76 17.25 16.80 17.06

0.10 0.14 0.12 0.12 0.14 0.13

and pure ZnO nanolayers) independently under the same experimental conditions. The histogram of the PCEavg (Figure 5b) exhibits that the cells (based on PEI-coated B, F:ZnO nanolayer) had an average efficiency of 15.14% and a PCEmax of 16.20%. 2.6. Device Performance for PEI-Mixed Pure and Doped ZnO Films. Earlier, we demonstrated that PEI coating covering the surface of pure and doped ZnO nanocrystals generated enhanced device efficiency compared to the only pure ZnO and doped ZnO ETLs. Next, we have electrosprayed PEI-mixed pure and doped ZnO nanocrystals to improve the device performance further thanks to the reduced traps found in the oxide nanocrystals filled by PEI and decreased surface roughness and reduced work function. Figure S7 displays the surface atomic force microscopy (AFM) monographs of the

PEI-coated B, F:ZnO nanolayer produced considerably better devices than uncoated nanolayers (PCEavg = 13.54%; PCEmax = 15.09%) and this enhancement is attributed to the reduced series resistance and rapid electron extraction.45 Furthermore, coating the surface of pure ZnO nanolayer using PEI polymer will also improve the device performance, but not much as observed for doped samples. The enhanced performance of doped films features the outstanding performance realized by the combined implementation of doping, surface modification of semiconducting surface, and PEI coating.23 The J−V plots for the best-performing perovskite devices after PEI coating also showed certainly improved efficiency, producing PCEmax values of 16.20 and 10.86% for the devices with B, F:ZnO and pure ZnO nanolayers, respectively (as plotted in Figure 5a). We have also prepared 50 cells (using PEI-coated B, F:ZnO 9653

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

Figure 6. (a) Steady-state PCEmax as a function of time for PEI-coated and PEI-mixed B, F:ZnO nanolayers, obtained by applying 835 mV bias voltage, which is identical to the maximum power output potential. Bar graphs of (b) PCEavg and (c) PCEmax summarizing the performance data for various types of ZnO and B, F:ZnO nanolayers and (d) hysteresis index for different types of ZnO and B, F:ZnO nanolayers.

committing the performance improvements of the perovskite devices. This is also proved by the EQE results exhibited in Figure S10 (Supporting Information), which are essentially in accordance with the trend of JSC in the J−V curves, as displayed in Figure 5b. The integrated JSC is 23.1 mA/cm2 (Figure S10, Supporting Information), which is satisfactory since the measured JSC (23.7 mA/cm2) from J−V curves is achieved in the reverse voltage scanning. In Figure 5d, the bar chart of the averaged PCE in the reverse and forward scanning directions for 50 separate cells based on PEI-mixed ZnO and B, F:ZnO nanolayers demonstrates good reproducibility with >85% of cells producing PCE > 10.76 and >17.06%, respectively. 2.7. Potential Key Summary Plots for Cells with PEICoated and PEI-Mixed ETLs. The champion devices with PEI-coated B, F:ZnO and PEI-mixed B, F:ZnO nanolayers also reveal the hysteresis-free current−voltage curves (Figure 5a,c), and the PCEmax (averaged) values calculated from those curves are in good agreement with that achieved from maximum power point tracking measurements (i.e., 0.835 V) at 1 sun illumination, as plotted in Figure 6a. It is also examined that the devices based on PEI-mixed B, F:ZnO nanolayers have a constant PCEmax of 20.74%, whereas the PCE of PEI-coated devices (16.2%) deteriorates a little with light soaking. The statistics of maximum and average PCEs obtained for the perovskite devices based on various ETLs (discussed in this study) are summarized in Figure 6b,c. We constantly demonstrate that multidoped nanolayer with highest conductivity and optical transmittance, high carrier concentration,

pure ZnO, PEI-coated ZnO, and ZnO/PEI composite layers. It is worth noting that for PEI-coated ETLs, surface roughness reduced from 6.78 nm for ZnO film to 5.21 nm and further decreased to 3.46 nm for polymer-mixed ZnO nanocrystals, possibly enhancing the physical contact and prompting robust molecular dipoles between the absorber layer and ITO substrate, yielding the enhanced PCEs.48 The PEI-mixed B, F:ZnO nanocrystals cause the JSC to increase from 21.85 to 22.75 mA/cm2, the FF to rise from 70 to 75%, and the VOC to increase from 990 to 1000 mV, as outlined in Table 3 and presented in Figures S8 and 5c, yielding a PCEavg of 17.06% (PCEmax = 20.74%). Thus, the PEI-mixed ETLs yielded substantially better devices than PEI-coated ETL-based devices (PCEavg = 15.14%; PCEmax = 16.2%) and is believed to benefit the devices by reducing the traps inside the oxide layer and having better physical contact between the polymer and oxide layer.48 Furthermore, a single layer existed between the perovskite layer and the PEI-mixed ZnO composite nanolayer interface; on the other hand, two interfaces are found between the PEI-coated ZnO bilayer ETL and perovskite layer, which will reduce the ability of the carrier to transport vertically due to surface traps. The best-performing cell revealed a PCEmax of 20.74%, the best efficiency ever stated for perovskite devices with ZnO ETLs. The PEI-mixed B, F:ZnO ETLs commit the improved absorption of perovskite absorber compared to the ETLs discussed above. This hypothesis is showed by the ultraviolet−visible absorption spectra presented in Figure S9 (Supporting Information). The perovskite layer based on PEImixed B, F:ZnO ETLs exhibits outstanding light absorption, 9654

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

nanolayers yielded marginally more stable data in terms of photovoltaic parameters (VOC, JSC, FF, and PCEavg) compared to devices based on PEI-coated nanolayers (decreased by only ∼6−8%) over a 15 day period (Figures S12 and S13, Supporting Information). Thus, the newly proposed polymer-mixed ETLs serve as a key addition and a scalable technique toward industrialization of perovskite solar cells.

and larger band-gap energy caused a maximum shift in the conduction band, surpassing pure ZnO nanolayers in all schemes and eventually producing devices with better PCEavg (2−3 points higher in each case). This highlights the potential benefits of multidopants on the electrical, optical, and morphological characteristics of electrosprayed deposited ZnO at low temperature. Improvements in device performance are further extended by applying conformal coating of PEI on the surface of ETLs (will reduce the work function further), resulting in highly efficient cells. The state-of-the-art devices (with PCEmax of 20.74%) are produced using a polymer-mixed oxide nanolayer thanks to the decrease in the trap sites of the ETLs, which would surely minimize the chances of trapassisted carrier interfacial recombination48 and eventually produce hysteresis-free, stable, and most efficient ZnO-based perovskite solar cells reported to date. Although the devices based on PEI-mixed nanolayers showed reduced hysteresis with almost no difference in the forward and backward scans, cells with lower PCEs presented more hysteresis for other ETL schemes. Interestingly, hysteresis index26 (used to evaluate the hysteresis response) varied significantly for the different ETL schemes (discuss in this study), as plotted in Figure 6d, and can be calculated from the following equation hysteresis index =

3. CONCLUSIONS In summary, an atmospheric pressure-based electrospray method is utilized for the formation of electron-rich multidoped ZnO nanolayers as ETLs for high-efficiency perovskite solar cells. The position of conduction band edge is continuously raised using various combinations of dopants, which causes a 0.22 eV shift in the conduction band compared to pure ZnO films, which will eventually improve the VOC, suppress the recombination, and enhance the faster transfer of electrons. The B, F:ZnO nanolayer with highest conductivity and optical transmittance, high carrier density, and larger bandgap energy and also benefited by the maximum shift in the conduction band demonstrates the highest PCEavg of 13.54% and PCEmax of 15.09%, which are much higher than those of pure ZnO film-based devices. Further improvements in the performance by reducing the work function of the oxide layer is achieved by coating the nanolayer surface with a thin layer of PEI, which results in remarkable highly efficient and hysteresisfree devices with a PCEmax of 16.20%. Eventually, a polymermixed ETL demonstrates a remarkable PCEmax of 20.74% by decreasing the bulk traps inside the oxide layer, which probably reduces the chances of trap-assisted surface recombination of charges and, subsequently, improves the device performance. The current work has notably confirmed the optimization of voltage by adjusting the energetic structure of the chargeremoving material together with the perovskite absorber, which will be one of the best reasonable approaches to boost the device efficiency.

Jscan −(VOC/2) − Jscan +(VOC/2) Jscan −(VOC/2)

(1)

where Jscan+(VOC/2) is the current density at VOC/2 for the forward scan and Jscan−(VOC/2) is the current density at VOC/2 for the backward scan. Hysteric devices demonstrate higher numbers of hysteresis index, whereas the lowest numbers represent the hysteresis-free devices. Particularly, in our case, the trend of hysteresis revealed that it is consistently lower for multidoped ZnO nanolayers with and without PEI coating on their surface and is further reduced to a minimum number when a polymer-mixed oxide nanolayer is used. It is important to mention that the decrease of hysteresis depends mainly on interfacial engineering of the perovskite devices by surface modification of ETLs surface.26 The lower hysteresis for devices based on multidoped nanolayers is mainly due to their better electrical, optical, and surface characteristics and the proven lower work function facilitating electron extraction and transfer through the ETL.26 A dipole layer (created on the ETLs surface by coating a thin layer of PEI) will further reduce the work function, resulting in the elimination of hysteresis significantly. Remarkably, the perovskite devices are also almost hysteresis-free by using a polymer-mixed nanolayer by reducing the traps in the oxide layer, which will definitely minimize trap-assisted interfacial recombination of carriers. Additionally, the hysteresis strongly relies on sweep directions of applied bias, voltage sweep rates, and light soaking.26 The devices based on PEI-mixed B, F:ZnO nanolayer demonstrate nonhysteric behavior compared to PEI-coated B, F:ZnO nanolayer (Figures S6 and S8), with negligible variation detected in the JSC when changing the voltage sweep rate or direction, as presented in Figure S11. The perovskite devices with PEI-coated and PEI-mixed B, F:ZnO nanolayers further undergo long-term stability test by light soaking in ambient environment (relative humidity = 40−45%) and storing the devices at 45 °C for 15 days continuously. In general, the perovskite solar cells are extremely susceptible to moisture and eventually rapidly deteriorate in humid environment because of their ionic nature.49 The devices based on PEI-mixed

4. EXPERIMENTAL SECTION 4.1. Electrosprayed Deposition of Pure and Multidoped ZnO Mesoscopic Nanolayers. Pure and doped ZnO (such as B:ZnO, Ta:ZnO, B, F:ZnO, and Ta, N:ZnO) mesoscopic nanolayers were coated onto the surface of ZnOblocking layer (deposited on indium tin oxide (ITO) substrates) using atmospheric pressure-based electrospraying method, as reported previously.37−39 Precursors such as zinc acetate dihydrate Zn(O2CCH3)2(H2O)2, boric acid (B(OH)3), tantalum (v) chloride (TaCl5), ammonium fluoride (NH4F), and ammonium acetate (C2H3O2NH4) were used as a source of zinc, boron, tantalum, fluorine, and nitrogen, respectively. The applied high voltages, precursor solution flow rates, and nozzle-to-substrate distances were maintained at 5.6 kV, 0.003 mL/min, and 4.0 cm, respectively. The precursor solution was sprayed for 1 h, while the hot plate temperature was retained at 150 °C (Table S1), and finally, the nanolayers were annealed at 410 °C in air atmosphere for 1 h. To synthesize the composite ETLs, the ZnO precursor was mixed using a fixed amount of PEI (7 wt %) and continuously stirred for 12 h, and the resultant solution was electrosprayed using the same conditions mentioned above. 4.2. Fabrication of Cells Using Pure and Doped ZnO Mesoscopic Nanolayers. First, a dense ZnO-blocking layer was spin-coated onto well-cleaned ITO substrates, followed by 9655

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

sintering at 500 °C for 30 min.22 After that, mesoporous pure and multidoped ZnO nanolayers were formed by electrospraying method, as explained above. A thin layer of PEI was formed over the surface of ZnO ETLs by spin-coating a PEI solution (prepared in deionized water with a concentration of 0.01 wt %) at 4500 rpm for 2 min and subsequent annealing at 150 °C for 5 min. After deposition, the perovskite absorber layer was formed over the surface of ZnO ETLs via the onestep spin-coating route, as explained below. The CH3NH3PbI(3−x)Clx solution (CH3NH3I/PbCl2 = 3:1, in N,N-dimethylformamide) was deposited by spin-coating at 2500 rpm for 30 s and subsequent drying at 95 °C for 1 h. Afterward, 30 μL of a hole-transport material was dropped on the substrate and spincoated at 3000 rpm for 30 s. Finally, 80 nm thick silver as a top electrode layer was formed by thermal evaporation method. 4.3. Characterization. The plane-view surface images of ZnO ETLs and cross-sectional view of complete devices were obtained using a scanning electron microscope (JSM-7600F, JEOL). The surface roughness was obtained by atomic force microscopy (AFM, SPA 400). X-ray photoelectron spectroscopy (XPS) was conducted to confirm the elemental composition of ETLs. The optical transmittance of the ETLs was measured using a UV−visible spectrometer (UV-3101PC). Electrical resistivity data were collected using four-point probe equipment (CMT-SR1000N). Hall effect measurement instrument (HMS-3000) was used to obtain the Hall mobility and carrier concentration data for various ETLs. The Brunauer− Emmett−Teller (BET) gas adsorption measurement route was used to measure the surface area of the powders using Quantochrome NOVA 1000 (Boynton Beach, FL). The current density vs voltage (J−V) plots were collected using a Keithley digital source meter by adjusting the intensity to 1 sun (100 mW/cm2). Both the reverse and forward scans were obtained at a rate of 10 mV/s with a delay time of 5 s. The active area of 0.25 cm2 was set using a metal mask aperture.



(2) Snaith, H. J. Perovskites: The Emergence of a New Era for LowCost, High-Efficiency Solar Cells. J. Phys. Chem. Lett. 2013, 4, 3623− 3630. (3) Green, M. A.; Hishikawa, Y.; Dunlop, E. D.; Levi, D. H.; HohlEbinger, J.; Ho-Baillie Anita, W. Y. Solar Cell Efficiency Tables (Version 51). Prog. Photovoltaics 2017, 26, 3−12. (4) Lee, M.; Teuscher, J.; Miyasaka, T.; Murakami, T.; Snaith, H. J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, No. 1228604. (5) Burschka, J.; Pellet, N.; Moon, S. J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Gratzel, M. Sequential Deposition As a Route to High-Performance Perovskite-Sensitized Solar Cells. Nature 2013, 499, No. 316. (6) Anaraki, E. H.; Kermanpur, A.; Steier, L.; Domanski, K.; Matsui, T.; Tress, W.; Saliba, M.; Abate, A.; Grätzel, M.; Hagfeldt, A.; CorreaBaena, J.-P. Highly Efficient and Stable Planar Perovskite Solar Cells by Solution-Processed Tin Oxide. Energy Environ. Sci. 2016, 9, 3128− 3134. (7) Jeon, N. J.; Noh, J. H.; Kim, Y. C.; Yang, W. S.; Ryu, S.; Seol, S. I. Sequential Deposition as a Route to High-Performance PerovskiteSensitized Solar Cells. Nat. Mater. 2014, 13, 897−903. (8) Im, J.-H.; Jang, I.-H.; Pellet, N.; Gratzel, M.; Park, N.-G. Growth of CH3NH3PbI3 Cuboids with Controlled Size for High-Efficiency Perovskite Solar Cells. Nat. Nanotechnol. 2014, 9, 927−932. (9) Mahmood, K.; Swain, B. S.; Amassian, A. Highly Efficient Hybrid Photovoltaics Based on Hyperbranched Three-Dimensional TiO2 Electron Transporting Materials. Adv. Mater. 2015, 27, 2859− 2865. (10) Yang, W. S.; Park, B. W.; Jung, E. H.; Jeon, N. J.; Kim, Y. C.; Lee, D. U.; Shin, S. S.; Seo, J.; Kim, E. K.; Noh, J. H.; Seok, S. I. Iodide Management in Formamidinium-Lead-Halide-Based Perovskite Layers for Efficient Solar Cells. Science 2017, 356, 1376−1379. (11) Han, G. S.; Chung, H. S.; Kim, B. J.; Kim, D. H.; Lee, J. W.; Swain, B. S.; Mahmood, K.; Yoo, J. S.; Park, N.-G.; Lee, J. H.; Jung, H. S. Retarding Charge Recombination in Perovskite Solar Cells Using Ultrathin MgO-Coated TiO2 Nanoparticulate Films. J. Mater. Chem. A 2015, 3, 9160−9164. (12) Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; Gratzel, M.; Park, N.-G. Lead Iodide Perovskite Sensitized AllSolid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, No. 591. (13) Kim, B. J.; Kim, D. H.; Lee, Y.-Y.; Shin, H.-W.; Han, G. S.; Hong, J. S.; Mahmood, K.; Ahn, T. K.; Joo, Y.-C.; Hong, K. S.; et al. Highly Efficient and Bending Durable Perovskite Solar Cells: Toward a Wearable Power Source. Energy Environ. Sci. 2015, 8, 916−921. (14) Li, Z. Q.; Chen, W. C.; Guo, F. L.; Mo, L. E.; Hu, L. H.; Dai, S. Y. Mesoporous TiO2 Yolk-Shell Microspheres for Dye-sensitized Solar Cells with a High Efficiency Exceeding 11%. Sci. Rep. 2015, 5, No. 14178. (15) Ren, Y. K.; Ding, X. H.; Wu, Y. H.; Zhu, J.; Hayat, T.; Alsaedi, A.; Xu, Y. F.; Li, Z. Q.; Yang, S. F.; Dai, S. Y. Temperature-Assisted Rapid Nucleation: A Facile Method to Optimize the Film Morphology for Perovskite Solar Cells. J. Mater. Chem. A 2017, 5, 20327−20333. (16) Ren, Y. K.; Shi, X. Q.; Ding, X. H.; Zhu, J.; Hayat, T.; Alsaedi, A.; Li, Z. Q.; Xu, X. X.; Yange, S. F.; Dai, S. Y. Facile Fabrication of Perovskite Layers With Large Grains Through a Solvent Exchange Approach. Inorg. Chem. Front. 2018, 5, 348−353. (17) Mahmood, K.; Swain, B. S.; Kirmani, A. R.; Amassian, A. Highly Efficient Perovskite Solar Cells Based on a Nanostructured WO3-TiO2 Core-Shell Electron Transporting Material. J. Mater. Chem. A 2015, 3, 9051−9057. (18) Oh, L. S.; Kim, D. H.; Lee, J. A.; Shin, S. S.; Lee, J.-W.; Park, I. J.; Ko, M. J.; Park, N.-G.; Pyo, S. G.; Hong, K. S.; Kim, J. Y. Zn2SnO4Based Photoelectrodes for Organolead Halide Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 22991−22994. (19) Bera, A.; Wu, K.; Sheikh, A.; Alarousu, E.; Mohammed, O. F.; Wu, T. Perovskite Oxide SrTiO3 as an Efficient Electron Transporter

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b01412.



SEM images of pure and doped ZnO ETLs, EQE spectra, J−V curves, XPS, and AFM images, UV−vis absorption spectra, and device stability (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Khalid Mahmood: 0000-0002-7611-3998 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are grateful to the Higher Education Commission (HEC) of Pakistan for financial support.



REFERENCES

(1) Best Research-Cell Efficiencies, NREL, 2016. http://www.nrel. gov/ncpv/images/effciency_chart.jpg. 9656

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657

ACS Omega

Article

for Hybrid Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 28494− 28501. (20) Shi, S.; Li, Y.; Li, X.; Wang, H. Advancements in All-Solid-State Hybrid Solar Cells Based on Organometal Halide Perovskites. Mater. Horiz. 2015, 2, 378−405. (21) Mahmood, K.; Park, S. B. Highly Efficient Dye-Sensitized Solar Cell With an Electrostatic Spray Deposited Upright-Standing BoronDoped ZnO (BZO) Nanoporous Nanosheet-Based Photoanode. J. Mater. Chem. A 2013, 1, 4826−4835. (22) Mahmood, K.; Swain, B. S.; Jung, H. S. Controlling the Surface Nanostructure of ZnO and Al-Doped ZnO Thin Films Using Electrostatic Spraying for Their Application in 12% Efficient Perovskite Solar Cells. Nanoscale 2014, 6, 9127−9138. (23) Liu, D. Y.; Kelly, T. L. Perovskite Solar Cells With a Planar Heterojunction Structure Prepared Using Room-Temperature Solution Processing Techniques. Nat. Photonics 2014, 8, 133−138. (24) Mahmood, K.; Sung, H. J. A Dye-Sensitized Solar Cell Based on a Boron-Doped ZnO (BZO) Film With Double Light-ScatteringLayers Structured Photoanode. J. Mater. Chem. A 2014, 2, 5408− 5417. (25) Mahmood, K.; Swain, B. S.; Amassian, A. Double-Layered ZnO Nanostructures for Efficient Perovskite Solar Cells. Nanoscale 2014, 6, 14674−14678. (26) Mahmood, K.; Swain, B. S.; Amassian, A. A. 16.1% Efficient Hysteresis-Free Mesostructured Perovskite Solar Cells Based on Synergistically Improved ZnO Nanorod Arrays. Adv. Energy Mater. 2015, 5, No. 1500568. (27) Son, D. Y.; Im, J.-H.; Kim, H.-S.; Park, N.-G. 11% Efficient Perovskite Solar Cell Based on ZnO Nanorods: An Effective Charge Collection System. J. Phys. Chem. C 2014, 118, 16567−16573. (28) Mahmood, K.; Kang, H. W.; Park, S. B.; Sung, H. J. Hydrothermally Grown Upright-Standing Nanoporous Nanosheets of Iodine-Doped ZnO (ZnO:I) Nanocrystallites for a High-Efficiency Dye-Sensitized Solar Cell. ACS Appl. Mater. Interfaces 2013, 5, 3075− 3084. (29) Mahmood, K.; Kang, H. W.; Munir, R.; Sung, H. J. A DualFunctional Double-Layer Film with Indium-Doped ZnO Nanosheets/ Nanoparticles Structured Photoanodes for Dye-Sensitized Solar Cells. RSC Adv. 2013, 3, 25136−25144. (30) Mahmood, K.; Munir, R.; Kang, H. W.; Sung, H. J. An Atmospheric Pressure-Based Electrospraying Route to Fabricate the Multi-Applications Bilayer (AZO/ITO) TCO Films. RSC Adv. 2013, 3, 25741−25751. (31) Chandiran, A. K.; Sauvage, F.; Cabanas, M. C.; Comte, P.; Zakeeruddin, S. M.; Graetzel, M. Doping a TiO2 Photoanode with Nb5+ to Enhance Transparency and Charge Collection Efficiency in Dye-Sensitized Solar Cells. J. Phys. Chem. C 2010, 114, 15849−15856. (32) Chandiran, A. K.; Tetreault, N.; Baker, R. H.; Kessler, F.; Baranoff, E.; Yi, C.; Zakeeruddin, S. M.; Graetzel, M. Subnanometer Ga2O3 Tunnelling Layer by Atomic Layer Deposition to Achieve 1.1 V Open-Circuit Potential in Dye-Sensitized Solar Cells. Nano Lett. 2012, 12, 3941−3947. (33) Li, T. C.; Góes, M. S.; Fabregat-Santiago, F.; Bisquert, J.; Bueno, P. R.; Prasittichai, C.; Hupp, J. T.; Marks, T. J. Surface Passivation of Nanoporous TiO2 via Atomic Layer Deposition of ZrO2 for Solid-State Dye-Sensitized Solar Cell Applications. J. Phys. Chem. C 2009, 113, 18385−18390. (34) Qin, P.; Domanski, A. L.; Chandiran, A. K.; Berger, R.; Butt, H.-J.; Dar, M. I.; Moehl, T.; Tetreault, N.; Gao, P.; Ahmad, S.; Nazeeruddin, M. K.; Grätzel, M. Yttrium-Substituted Nanocrystalline TiO2 Photoanodes for Perovskite Based Heterojunction Solar Cells. Nanoscale 2014, 6, 1508−1514. (35) Manseki, K.; Ikeya, T.; Tamura, A.; Ban, T.; Sugiura, T.; Yoshida, T. Mg-Doped TiO2 Nanorods Improving Open-Circuit Voltages of Ammonium Lead Halide Perovskite Solar Cells. RSC Adv. 2014, 4, 9652−9655. (36) Yang, M.; Guo, R.; Kadel, K.; Liu, Y.; O’Shea, K.; Bone, R.; Wang, X.; He, J.; Li, W. Improved Charge Transport of Nb-Doped

TiO2 Nanorods in Methylammonium Lead Iodide Bromide Perovskite Solar Cells. J. Mater. Chem. A 2014, 2, 19616−19622. (37) Mahmood, K.; Song, D.-S.; Park, S. B. Effects of Thermal Treatment on the Characteristics of Boron and Tantalum-doped ZnO Thin Films Deposited by the Electrospraying Method at Atmospheric Pressure. Surf. Coat. Technol. 2012, 206, 4730−4740. (38) Mahmood, K.; Park, S. B. Conductivity Enhancement by Fluorine Doping in Boron-Doped ZnO Thin Films Deposited by the Electrospraying Method. J. Cryst. Growth 2012, 361, 30−37. (39) Mahmood, K.; Park, S. B.; Sung, H. J. Fabrication of Tantalum and Nitrogen Codoped ZnO (Ta, N-ZnO) Thin Films Using the Electrospay: Twin Applications as an Excellent Transparent Electrode and a Field Emitter. ACS Appl. Mater. Interfaces 2013, 5, 3722−3730. (40) Xie, C.; Chen, L.; Chen, Y. Electrostatic Self-Assembled Metal Oxide/Conjugated Polyelectrolytes as Electron-Transporting Layers for Inverted Solar Cells with High Efficiency. J. Phys. Chem. C 2013, 117, 24804−24814. (41) He, Z.; Zhong, C.; Su, S.; Xu, M.; Wu, H.; Cao, Y. Enhanced Power-Conversion Efficiency in Polymer Solar Cells Using an Inverted Device Structure. Nat. Photonics 2012, 6, 591−595. (42) Sun, K.; Zhao, B.; Kumar, A.; Zeng, K.; Ouyang, J. Highly Efficient, Inverted Polymer Solar Cells with Indium Tin Oxide Modified with Solution-Processed Zwitterions as the Transparent Cathode. ACS Appl. Mater. Interfaces 2012, 4, 2009−2017. (43) Xu, X.; Zhu, Y.; Zhang, L.; Sun, J.; Huang, J.; Chen, J.; Cao, Y. Hydrophilic Poly(Triphenylamines) With Phosphonate Groups on the Side Chains:Synthesis and Photovoltaic Applications. J. Mater. Chem. 2012, 22, 4329−4336. (44) Zhou, Y.; Fuentes-Hernandez, C.; Shim, J.; Meyer, J.; Giordano, A. J.; Li, H.; Winget, P.; Papadopoulos, T.; Cheun, H.; Kim, J.; et al. A Universal Method to Produce Low-Work Function Electrodes for Organic Electronics. Science 2012, 336, 327−332. (45) Woo, S.; Kim, W. H.; Kim, H.; Yi, Y.; Lyu, H. K.; Kim, Y. 8.9% Single-Stack Inverted Polymer Solar Cells with Electron-Rich Polymer Nanolayer-Modified Inorganic Electron-Collecting Buffer Layers. Adv. Energy Mater. 2014, 4, No. 1301692. (46) Shao, S.; Zheng, K.; Pullerits, T.; Zhang, F. Enhanced Performance of Inverted Polymer Solar Cells by Using Poly(ethylene oxide)-Modified ZnO as an Electron Transport Layer. ACS Appl. Mater. Interfaces 2013, 5, 380−385. (47) Jo, S. B.; Lee, J. H.; Sim, M.; Kim, M.; Park, J. H.; Choi, Y. S.; Kim, Y.; Ihn, S.-G.; Cho, K. High Performance Organic Photovoltaic Cells Using Polymer-Hybridized ZnO Nanocrystals as a Cathode Interlayer. Adv. Energy Mater. 2011, 1, 690−698. (48) Chen, H.-C.; Lin, S.-W.; Jiang, J.-M.; Su, Y.-W.; Wei, K.-H. Solution-Processed Zinc Oxide/Polyethylenimine Nanocomposites as Tunable Electron Transport Layers for Highly Efficient Bulk Heterojunction Polymer Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 6273−6281. (49) Zheng, Y.-Z.; Zhao, E. F.; Meng, F. L.; Lai, X. S.; Dong, X. M.; Wua, J. J.; Tao, X. Iodine-Doped ZnO Nanopillar Arrays for Perovskite Solar Cells With High Efficiency Up to 18.24%. J. Mater. Chem. A 2017, 5, 12416−12425. (50) Luo, J.; Wang, Y.; Zhang, Q. Progress in Perovskite Solar Cells Based on ZnO Nanostructures. Solar Energy 2018, 163, 289−306. (51) Song, J.; Liu, L.; Wang, X. F.; Chen, G.; Tian, W.; Miyasaka, T. Highly Efficient and Stable Low-Temperature Processed ZnO Solar Cells With Triple Cation Perovskite Absorber. J. Mater. Chem. A 2017, 5, 13439−13447. (52) Mahmood, K.; Sarwar, S.; Mehran, M. T. Current Status of Electron Transport Layers in Perovskite Solar Cells: Materials and Properties. RSC Adv. 2017, 7, 17044−17062. (53) Cao, J.; Wu, B.; Chen, R.; Wu, Y.; Hui, Y.; Mao, B.-W.; Zheng, N. Efficient, Hysteresis-Free, and Stable Perovskite Solar Cells with ZnO as Electron-Transport Layer: Effect of Surface Passivation. Adv. Mater. 2018, 30, No. 1705596.

9657

DOI: 10.1021/acsomega.8b01412 ACS Omega 2018, 3, 9648−9657