Key parameters of two typical intercalation reactions to prepare hybrid inorganic–organic perovskite films
Shi Biao1, 2, 3, 4, Guo Sheng1, 2, 3, 4, Wei Changchun1, 2, 3, 4, Li Baozhang1, 2, 3, 4, Ding Yi1, 2, 3, 4, Li Yuelong1, 2, 3, 4, Wan Qing5, Zhao Ying1, 2, 3, 4, Zhang Xiaodan1, 2, 3, 4, †
Institute of Photoelectronic Thin Film Devices and Technology, Nankai University, Tianjin 300071, China
Key Laboratory of Photoelectronic Thin Film Devices and Technology of Tianjin, Tianjin 300071, China
Key Laboratory of Optical Information Science and Technology of Ministry of Education, Tianjin 300071, China
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China
College of Electronic Science and Engineering, Nanjing University, Nanjing 210093, China

 

† Corresponding author. E-mail: xdzhang@nankai.edu.cn

Abstract

A star hybrid inorganic–organic perovskite material selected as an outstanding absorbing layer in solar cells benefits from multiple preparation techniques and excellent photoelectric characteristics. Among numerous synthetic processes, uniform, compact, and multi-stack perovskite thin films can be manufactured using vacuum deposition. During sequential vacuum deposition, the penetration ability of the organic molecules cannot be effectively controlled. In addition, the relationship between the thickness of the inorganic seeding layer and the organic molecule concentration for optimized devices using an evaporation–solution method is unclear. In this work, we prepared high-quality perovskite films by effectively controlling the penetration ability and chemical quantity of organic methyl ammonium iodide by monitoring the evaporation pressure and time. Thus, a device efficiency of over 15% was achieved with an all-vacuum prepared perovskite film. For the evaporation–solution method, we reacted different thicknesses of inorganic lead iodine with various concentrations of the organic molecule solution. The inorganic layer thickness and organic molecule concentration showed a linear relationship to achieve an optimum perovskite film, and an empirical formula was obtained. This work noted the key parameters of two intercalation reactions to prepare perovskite films, which paves a way to deliver a device that enables multi-layered structures, such as tandem solar cells.

1. Introduction

Hybrid organic–inorganic perovskite materials have attracted tremendous attention due to their excellent properties, such as large charge carrier mobility,[1] long diffusion length,[2] high absorption coefficient,[3,4] tunable band gap,[58] and unique tolerance to structural defects.[9,10] The outstanding optoelectronic characteristics combined with multiple manufacturing methods[1119] have presented such materials with extensive applications in light emitting devices (LEDs),[20,21] thin film transistors (TFTs),[22] photodetectors,[23,24] solar cells,[2528] and others. During preparation, a distinct advantage of vacuum deposition over solution processing is the ability to prepare layered uniform, compact, and multi-stack thin films,[15] which enables multi-layered structures. However, the efficiencies are lower and successful cases of efficiencies over 15% are few for vacuum deposited solar cells[15,29] compared to solar cells prepared with solution processing,[30] which mainly results from the small density and high vapor pressure of organic methyl ammonium iodide (MAI) diffusing inside the vacuum chamber, making it very difficult to control the ratio of organic MAI to inorganic lead salt.[3134] To improve the controllability during vacuum deposition, inorganic lead halides and organic MAI have been prepared in separated processes to synthesize perovskite films, as shown in the scheme in Fig. 1(a). Vapor-assisted solution processing,[16,35,36] sequential vacuum deposition,[37,38] evaporation–solution processing,[3945] hybrid chemical vapor deposition,[46] layer-by-layer growth,[47] and alternating precursor layer deposition[48] have been developed and can enhance the experimental repeatability, as these preparation routes can provide strict control over the stoichiometry of organic MAI vapor. However, as common methods of control, monitoring the organic vapor rate with quartz monitor crystals[36,48] or evaporating organic halide at a fixed temperature[32] are invalid for controlling the penetration ability of the randomly diffused organic halide vapor. In addition, the relationship between the thickness of the inorganic lead halide layer and the concentration of MAI/isopropyl alcohol (IPA) for optimized device performance is unclear, which is vital for evaporation–solution methods widely used in semitransparent and tandem devices.[4042,49,50]

Fig. 1. (color online) Scheme illustration of the intercalation reaction to form the perovskite film and schematic structure of the PSC. (a) First, the inorganic lead halide was sublimed, and then, the as-deposited inorganic seeding layer was reacted with organic molecules through a penetration process. (b) Schematic structure of the PSC with the perovskite film prepared by sequential vacuum deposition and evaporation–solution methods.

In fact, an intercalation process is required to manufacture a perovskite film with a step-by-step deposition process.[5153] During the organic molecule penetration process into the inorganic skeleton, the thickness of the inorganic seeding layer and the penetration ability and chemical quantity of the organic molecules are core factors. Heating the substrate has been shown as a successful method to enhance the penetration ability of organic MAI molecules.[37] In our work, first, we prepared a high-quality perovskite film by controlling the penetration ability and chemical quantity of the MAI vapor through monitoring the evaporation pressure and time. The efficiency of the perovskite solar cell (PSC) reached over 15% with the all-vacuum prepared perovskite film. Moreover, the thickness of the inorganic lead halide and the concentration of the organic molecule solution show a linear relationship with the optimized perovskite film, and an empirical formula was proposed for the evaporation–solution method.

2. Experimental section
2.1. Substrate cleaning and HTL precursor preparation

Transparent conductive fluorine-doped tin oxide (FTO)/glass ( ) films were purchased from Lattice Solar Energy Technology Co. Ltd. in Wuhan. The FTO/glass substrates were first sequentially cleaned with a semiconductor industrial cleaner (Huaxing DZ-1, Jinan Xihua Technology Co. Ltd) and deionized water in an ultrasonic bath for 50 min and then dried using nitrogen gas. A chlorobenzene solution (1 mL) containing -tetrakis(N, -di-p-methoxyphenylamine)-9, -spirobifluorene (spiro-OMeTAD, 80 mg), 4-tert-butylpyridine (TBP, ) and a lithium bis(trifluoromethylsulfonyl)imide (LiTFSI, ) solution (520 mg LiTFSI in 1 mL of acetonitrile) were prepared for the organic hole transporting layer (HTL).

2.2. Cells with perovskite films manufactured by sequential vacuum deposition

A TiO2 film with a thickness of nm was deposited on top of the FTO/glass substrate by DC magnetron sputtering. Then, the substrate was transferred into a vacuum system (Kurt J. Lesker, Britain). After the base pressure of the system was pumped down to , PbCl2 evaporation was initiated by heating the crucible, and the evaporation rate of PbCl2 was precisely maintained at 2 Å/s under careful monitoring by a quartz crystal. The heating of the PbCl2 crucible was immediately stopped when the thickness on the quartz crystal reached 3.6 kÅ, which resulted in a PbCl2 film with an actual thickness of 100 nm. After cooling the PbCl2 crucible for 5 min, a crucible with MAI was slowly heated until the chamber pressure reached the required pressure, and the evaporation temperature was maintained over time. The as-deposited films were annealed at 100 °C for 30 min in a N2-filled glove box. Then, of a spiro-OMeTAD solution was spin coated on the perovskite films at 6000 rpm for 30 s. Finally, a 100 nm gold layer was thermally evaporated on top of the HTL to complete the device.

2.3. Cells with perovskite films manufactured by an evaporation–solution method

First, a 12 nm TiO2 film with titanium isopropoxide (TTIP) and oxygen plasma as precursors was deposited onto the FTO/glass by atomic layer deposition (ALD) (Model: SI ALD, Company: SENTECH). Then, of a 10 mg/mL [6,6]-phenyl C61-butyric acid methyl ester (PC61BM)/chlorobenzene (CB) solution was spin coated onto the TiO2/FTO/glass substrates at 3500 rpm for 30 s. Different thicknesses of PbI2 were subsequently deposited on the substrates by evaporation. The base pressure of the system was 3×10−6 Torr, and the evaporation rate was 2 Å/s. The substrates coated with PbI2 films were taken into a glove box without exposure to air. MAI/IPA solutions with various concentrations were loaded on the PbI2 film for 1 min before rotating the substrates at 2000 rpm for 40 s and then annealing the films at 140 °C for 25 min. Then, of a spiro-OMeTAD solution was spin coated on the perovskite films at 4000 rpm for 60 s. Finally, a 100 nm gold layer was thermally evaporated on top of the HTL to complete the device.

2.4. Film and device characterization

The PbI2 thickness was measured with a microfigure measuring instrument (Surfcorder ET200, Kosaka Laboratory Ltd). The morphologies of the perovskite films were characterized using scanning electron microscopy (SEM) (JEOL JSM-6700F). X-ray diffraction (XRD, Rigaku ATX-XRD) patterns of the perovskite films were obtained using Cu radiation as the radiation source ( ) across a range of 5° to 70°. Time-resolved photoluminescence (TRPL) spectroscopy was measured with a PL spectrometer (Edinburgh Instruments, FLS 920), and a pulsed laser with a wavelength of 635 nm and a repetition rate of 1 MHz was employed as the excitation source. A 655 nm filter was used to filter out the excitation light. Photocurrent density–voltage (JV) curves of the solar cells were measured from 1200 mV to −200 mV (forward bias to short circuit (FB-SC)) and returned (SC-FB) with a voltage step of 40 mV and delay time of 50 ms at 25 °C in a N2-filled glovebox. A metal mask with a window of 0.1 cm2 was coated on the light incident side to define the active area of the cell. A solar simulator (HAL-320, ASAHI SPECTRA Co. Ltd., Japan) with a compact xenon light source was used to produce the simulated AM 1.5G irradiation (100 mW/cm2), and the calibration of the light was carried out by a detector (CS-20, ASAHI SPECTRA Co. Ltd., Japan) with a silicon reference cell. The external quantum efficiency (EQE) spectral response was measured with a QEX10 (PV measurement).

3. Results and discussion

The perovskite film was manufactured with sequential vacuum deposition,[37,54] and the detailed process has been described in the experimental section. A 100 nm inorganic PbCl2 seeding layer was first sublimed onto the substrate with a controlled rate. In addition to the concentration and temperature, the pressure of the organic MAI vapor was manipulated to realize an effective diffusion reaction. The grain size at a low organic vapor pressure (Fig. 2(a), 5 × 10−5 Torr) was smaller than that at a high vapor pressure (Figs. 2(b) and 2(c), ). The three XRD peaks of CH3NH3PbI3 (MAPbI3) corresponding to the (110), (220), and (330)[15,55] lattice planes were stronger and shaper at an organic vapor pressure of , as shown in Fig. 2(d), indicating the formation of larger perovskite grains. The increased crystalline grain size of the perovskite at high pressure was attributed to the improved reaction rate of PbCl2 and MAI,[36] referred to as Ostwald ripening.[56,57] At high MAI vapor pressure, more and higher-reactive MAI molecules will react with PbCl2, which favors secondary nucleation at the primary sites. The nucleation sites of the large crystallites grow faster because of the larger contact area. As a result, the larger crystalline grains grow even larger and diminish the small perovskite nucleation sites. Therefore, a suitable high pressure of organic MAI is beneficial for the formation of high-quality perovskite films.

Fig. 2. (color online) Surface morphology, phase, and electrical properties of the perovskite film prepared by sequential vacuum deposition before annealing. SEM top-view images of the perovskite films prepared with various organic MAI sublimation pressures and duration time: (a) , 10 h; (b) , 5 h; (c) , 10 h. (d) XRD patterns of the perovskite films prepared with various sublimation conditions of organic MAI. (e) TRPL spectroscopy of the FTO/TiO2/perovskite film prepared with a MAI vapor pressure of or . The inorganic seeding layer is 100 nm PbCl2. The sublimation duration time of organic MAI reflects its chemical quality and is sufficient when the time is 10 h but insufficient at 5 h.

Transient TRPL decay measurements were conducted to investigate the electrical properties of the perovskite film (Fig. 2(e)). A single-exponential function was used to fit the PL decay time. The decay time of the sample with an organic vapor pressure of is ns, while the decay time of is ns. The rapid decay for the sample prepared at an organic vapor pressure of 10 Torr results from the quick recombination in the perovskite bulk caused by more grain boundaries.[58]

The chemical quantity of MAI is also important in the formation of MAPbI3. As shown in Fig. 2(d), the peaks at and 15.56° corresponding to PbI2 (001) and CH3NH3PbCl3 (MAPbCl3) (100)[15,59] appear at an MAI evaporation time of 5 h, indicating the formation of non-pure MAPbI3. However, pure MAPbI3 can be obtained when prolonging the evaporation time of MAI. Based on previous studies[5961] and the XRD patterns, the reaction steps are listed as follows: First, reactions (1) and (2) occur. Then, MAI reacts with PbI2 and MAPbCl3 through reactions (3) and (4), respectively. With an insufficient amount of MAI, PbI2 and MAPbCl3 cannot be consumed, leaving in MAPbI3. With a sufficient amount of MAI, pure MAPbI3 is formed.

The efficiency of a PSC (structural schematic shown in Fig. 1(b)) is optimal only when the perovskite film is formed with a high organic MAI vapor pressure and sufficient MAI content. The JV characteristics and corresponding EQE curves are shown in Fig. 3, and the device performance parameters are summarized in Table 1. The current density ( from the JV curves was larger than integral from EQE, resulting from lateral collection.[62] The PSC with the absorber formed at an MAI vapor pressure of Torr demonstrated higher , open-circuit voltage ( ), and fill factor ( ), which are attributed to the larger perovskite grain size, reducing recombination in the perovskite bulk. The PSC fabricated at an MAI vapor pressure of Torr but with an MAI evaporation time of 5 h displayed lower , FF and higher than the cell with an MAI evaporation time of 10 h, which may result from PbI2 and MAPbCl3 reducing the absorption[63] but enhancing the band gap.[64]

Fig. 3. (color online) Performance of PSCs prepared using sequential vacuum deposition. (a) JV characteristics with an effective area of 0.1 cm2 (FB-SC) under 1 sun illumination and (b) the corresponding EQE curves for PSCs with various thermal evaporation deposition conditions.
Table 1.

PSC parameters with different organic vapor pressures and duration time.

.

For the intercalation reaction of the forming perovskite film, the thickness of the inorganic seeding layer and the concentration of the organic molecules are vital.[18] However, the quantitative relationship between the thickness of the inorganic seeding layer and the concentration of the organic molecules for optimal device performance is unclear. We investigated this relationship via an evaporation–solution method[43] because thermally evaporated PbI2 is non-destructive to substrates and can be prepared with controlled morphology and thickness, and the concentration of the MAI/IPA solution can be easily changed. PbI2 layers with thicknesses of 85 nm, 150 nm, 200 nm, 270 nm were prepared, and MAI/IPA solutions with four concentrations were spin coated onto each PbI2 thickness. The performance parameters of PSCs with different PbI2 thicknesses and MAI/IPA concentrations are shown in Fig. 4. When PbI2 possessing each thickness reacts with a particular MAI/IPA solution concentration, the PSC can acquire an optimal performance. Maximum photoelectric conversion efficiencies (PCEs) of 11.88%, 13.99%, 14.91%, and 13.33% were obtained with 85 nm, 150 nm, 200 nm, 270 nm PbI2 reacting with 16 mg/mL, 27.5 mg/mL, 33 mg/mL, 46.5 mg/mL MAI/IPA, respectively.

Fig. 4. (color online) Performance parameters of the PSCs prepared using the evaporation–solution method. Performances of the PSCs with PbI2 thicknesses of (a) 85 nm, (b) 150 nm, (c) 200 nm, and (d) 270 nm reacting with various concentrations of the MAI/IPA solution.

The optimum matching of the MAI/IPA concentration with the PbI2 thickness is depicted in Fig. 5. Obviously, the optimum MAI/IPA concentration to acquire the maximum PSC efficiency linearly increases with the PbI2 thickness, and the linear formula fit to the MAI/IPA concentration versus PbI2 thickness data is where y is the MAI/IPA concentration, and x is the PbI2 thickness. To examine the feasibility of the above formula, we prepared 175 nm PbI2 and reacted the deposited film with 25 mg/mL, 28 mg/mL, 31 mg/mL, and 35 mg/mL MAI/IPA. According to the above formula, the optimum match is 175 nm PbI2 reacting with 30.5 mg/mL MAI/IPA. Within the margin of error, the optimum efficiency of 14.69% was achieved with the PSC with the film formed using 31 mg/mL MAI/IPA, and the performance curves showed negligible hysteresis, as shown in Fig. 6.

Fig. 5. (color online) Relationship between the PbI2 thickness and MAI/IPA concentration to achieve the highest PSC efficiency. The thickness of the PbI2 film has a linear correlation with the MAI/IPA concentration to achieve the optimum perovskite film.
Fig. 6. (color online) Performance of the PSC with 175 nm PbI2 and 30.5 mg/mL MAI/IPA. (a) JV curve with an effective area of 0.1 cm2 and (b) EQE curve of the cell under 1 sun illumination.

As seen from Fig. 5, the PSC efficiency increases with an increasing PbI2 thickness up to 200 nm and decreases with 270 nm PbI2. Figures 7(a)7(d) show the cross-sectional SEM images of the perovskite films with 85 nm, 150 nm, 200 nm, and 270 nm PbI2, respectively. These perovskite films are approximately twice as thick as the PbI2 films, agreeing with the previous reports.[65,66] The thickness of the perovskite film increases with an increasing PbI2 thickness, leading to obviously enhanced ultraviolet-visible absorption between 550 nm to 800 nm, as shown in Fig. 7(e). When the thickness of the perovskite film is less than 400 nm, as shown in Figs. 7(a)7(c), the, the crystalline grains are perpendicular to the substrates and larger than the film thickness, which are beneficial to transporting photogenerated carriers. For a thicker perovskite film (Fig. 7(d)), the film has a reduced quality with more grain boundaries, resulting in more loss of electricity. Balancing the optical absorption and carrier transport, the optimum thickness of the perovskite film is 400 nm manufactured with 200 nm PbI2 and 33 mg/mL MAI/IPA. The JV curves and performance parameters of the perovskite films of different thicknesses are shown in Fig. 7(f) and Table 2.

Fig. 7. (color online) SEM images and absorption properties of the perovskite films and performances of the PSCs. SEM cross-sectional images of perovskite films with thicknesses of (a) 160 nm, (b) 300 nm, (c) 400 nm, and (d) 540 nm manufactured by reacting 85 nm, 150 nm, 200 nm, and 270 nm PbI2 with 16 mg/mL, 27.5 mg/mL, 33 mg/mL, and 46.5 mg/mL MAI/IPA, respectively. (e) Absorption spectra of the perovskite films, and (f) JV curves (FB-SC) of the PSCs under 1 sun illumination with different PbI2 thicknesses matched with the optimal concentrations of MAI/IPA.
Table 2.

Performance parameters of the optimal PSCs using different PbI2 thicknesses reacting with the optimal concentrations of the MAI/IPA solution.

.
4. Conclusion

We obtained large-grain, pure CH3NH3PbI3 using a sequential vacuum deposition, and an efficiency of over 15% was achieved with the PSC fabricated using a sufficient and high pressure of organic MAI vapor. In addition, we investigated the relationship between the inorganic seeding layer thickness and organic molecule concentration to achieve optimum devices using the evaporation–solution method. For the optimum PSC at each PbI2 thickness, the thickness of the inorganic lead halide and concentration of organic molecule solution showed a linear relationship, and an empirical formula was determined.

Data availability

The data that support the findings of this study are available from the corresponding author on reasonable request.

Reference
[1] Ponseca C S Jr. Savenije T J Abdellah M Zheng K Yartsev A Pascher T Harlang T Chabera P Pullerits T Stepanov A Wolf J Sundström V 2014 J. Am. Chem. Soc. 136 5189
[2] Stranks S Eperon G Grancini G Menelaou C Alcocer M Leijtens T Herz L Petrozza A Snaith H J 2013 Science 342 341
[3] Gao P Grätzel M Nazeeruddin M K 2014 Energy Environ. Sci. 7 2448
[4] Green M A Ho-Baillie A Snaith H J 2014 Nat. Photon. 8 506
[5] Correa Baena J P Steier L Tress W Saliba M Neutzner S Matsui T Giordano F Jacobsson T J Srimath Kandada A R Zakeeruddin S M Petrozza A Abate A Nazeeruddin M K Grätzel M Hagfeldt A 2015 Energy Environ. Sci. 8 2928
[6] Jesper Jacobsson T Correa-Baena J P Pazoki M Saliba M Schenk K Grätzel M Hagfeldt A 2016 Energy Environ. Sci. 9 1706
[7] Yang Z Rajagopal A Chueh C C Jo S B Liu B Zhao T Jen A K 2016 Adv. Mater. 28 8990
[8] Eperon G E Leijtens T Bush K A Prasanna R Green T Wang J T McMeekin D P Volonakis G Milot R L May R Palmstrom A Slotcavage D J Belisle R A Patel J B Parrott E S Sutton R J Ma W Moghadam F Conings B Babayigit A Boyen H Bent S Giustino F Herz L M Johnston M B McGehee M D Snaith H J 2016 Science 354 861
[9] Yin W J Shi T Yan Y 2014 Appl. Phys. Lett. 104 063903
[10] Kim J Lee S H Lee J H Hong K H 2014 J. Phys. Chem. Lett. 5 1312
[11] Sutherland B R Hoogland S Adachi M M Kanjanaboos P Wong C T McDowell J J Xu J Voznyy O Ning Z Houtepen A J Sargent E H 2015 Adv. Mater. 27 53
[12] Malinkiewicz O Lenes M Brine H Bolink H J 2012 RSC Adv. 2 3335
[13] Ono L K Wang S Kato Y Raga S R Qi Y 2014 Energy Environ. Sci. 7 3989
[14] Burschka J Pellet N Moon S J Humphry-Baker R Gao P Nazeeruddin M K Grätzel M 2013 Nature 499 316
[15] Liu M Johnston M B Snaith H J 2013 Nature 501 395
[16] Chen Q Zhou H Hong Z Luo S Duan H S Wang H H Liu Y Li G Yang Y 2013 J. Am. Chem. Soc. 136 622
[17] Das S Yang B Gu G Joshi P C Ivanov I N Rouleau C M Aytug T Geohegan D B Xiao K 2015 ACS Photonics 2 680
[18] Xiao Z Bi C Shao Y Dong Q Wang Q Yuan Y Wang C Gao Y Huang J 2014 Energy Environ. Sci. 7 2619
[19] Chen B Bai Y Yu Z Li T Zheng X Dong Q Shen L Boccard M Gruverman A Holman Z Huang J 2016 Adv Energy. Mater. 6 1601128
[20] Chen Z Zhang C Jiang X F Liu M Xia R Shi T Chen D Xue Q Zhao Y Su S Yip H Cao Y 2017 Adv. Mater. 29 1603157
[21] Wang J Wang N Jin Y Si J Tan Z Du H Cheng L Dai X Bai S He H Ye Z Lai M L Friend R H Huang W 2015 Adv. Mater. 27 2311
[22] Wu Y Li J Xu J Du Y Huang L Ni J Cai H Zhang J 2016 RSC Adv. 6 16243
[23] Lu H Tian W Cao F Ma Y Gu B Li L 2016 Adv. Funct. Mater. 26 1296
[24] Shen L Fang Y Wang D Bai Y Deng Y Wang M Huang J 2016 Adv. Mater. 28 10794
[25] Bi D Yi C Luo J Décoppet J Zhang F Zakeeruddin Shaik M Li X Hagfeldt A Grätzel M 2016 Nat. Energy. 1 16142
[26] Kakavelakis G Maksudov T Konios D Paradisanos I Kioseoglou G Stratakis E Kymakis E 2017 Adv. Energy Mater. 7 1602120
[27] Tsai H Nie W Blancon J C Stoumpos C C Asadpour R Harutyunyan B Neukirch A J Verduzco R Crochet J J Tretiak S Pedesseau L Even J Alam M A Gupta G Lou J Ajayan P M Bedzyk M J Kanatzidis M G 2016 Nature 536 312
[28] You J Meng L Song T B Guo T F Yang Y M Chang W H Hong Z Chen H Zhou H Chen Q Liu Y De Marco N Yang Y 2016 Nat. Nanotech. 11 75
[29] Lin Q Armin A Nagiri R C R Burn P L Meredith P 2014 Nat. Photon. 9 106
[30] Yang W S Park B W Jung E H Jeon N J Kim Y C Lee D U Seok S I 2017 Science 356 1376
[31] Malinkiewicz O Yella A Lee Y H Espallargas G M Graetzel M Nazeeruddin M K Bolink H J 2013 Nat. Photon. 8 128
[32] Malinkiewicz O Roldán-Carmona C Soriano A Bandiello E Camacho L Nazeeruddin M K Bolink H J 2014 Adv Energy Mater 4
[33] Teuscher J Ulianov A Muntener O Gratzel M Tetreault N 2015 ChemSusChem 8 3847
[34] Gao C Liu J Liao C Ye Q Zhang Y He X Guo X Mei J Lau W 2015 RSC Adv. 5 26175
[35] Zhou H Chen Q Yang Y 2015 MRS Bull. 40 667
[36] Mao J Zhang H He H Lu H Xie F Zhang D Wong K S Choy W C H 2015 RSC Adv. 5 73760
[37] Chen C W Kang H W Hsiao S Y Yang P F Chiang K M Lin H W 2014 Adv. Mater. 26 6647
[38] Hu H Wang D Zhou Y Zhang J Lv S Pang S Chen X Liu Z Padture N P Cui G 2014 RSC Adv. 4 28964
[39] Bailie C D Christoforo M G Mailoa J P Bowring A R Unger E L Nguyen W H Burschka J Pellet N Lee J Z Grätzel M Noufi R Buonassisi T Salleo A McGehee M D 2015 Energy Environ. Sci. 8 956
[40] Fu F Feurer T Jager T Avancini E Bissig B Yoon S Buecheler S Tiwari A N 2015 Nat. Commun. 6
[41] Fu F Feurer T Weiss Thomas P Pisoni S Avancini E Andres C Buecheler S Tiwari Ayodhya N 2016 Nat. Energy 2 16190
[42] Werner J Weng C-H Walter A Fesquet L Seif J P Wolf S D Niesen B Ballif C 2015 J. Phys. Chem. Lett. 7 161
[43] Tao C Neutzner S Colella L Marras S Srimath Kandada A R Gandini M Bastiani M D Pace G Manna L Caironi M Bertarelli C Petrozza A 2015 Energy Environ. Sci. 8 2365
[44] Ioakeimidis A Christodoulou C Lux-Steiner M Fostiropoulos K 2016 J. Solid State Chem. 244 20
[45] Liu D Gangishetty M K Kelly T L 2014 J. Mater. Chem. 2 19873
[46] Leyden M R Ono L K Raga S R Kato Y Wang S Qi Y 2014 J. Mater. Chem. 2 18742
[47] Chen Y Chen T Dai L 2015 Adv. Mater. 27 1053
[48] Yang D Yang Z Qin W Zhang Y Liu S Li C 2015 J. Mater. Chem. 3 9401
[49] Werner J Barraud L Walter A Bräuninger M Sahli F Sacchetto D Tétreault N Paviet-Salomon B Moon S J Allebé C Despeisse M Nicolay S De Wolf S Niesen B Ballif C 2016 ACS Energy Lette ACS Energy Lett. 1 474
[50] Forgács D Gil-Escrig L Pérez-Del-Rey D Momblona C Werner J Niesen B Ballif C Sessolo M Bolink H J 2017 Adv. Energy. Mater. 7 1602121
[51] Ahmad S Kanaujia P K Niu W Baumberg J J Vijaya Prakash G 2014 ACS Appl. Mater. Interfaces 6 10238
[52] Era M Hattori T Taira T Tsutsui T 1997 Chem. Mater. 9 8
[53] Era M Maeda K Tsutsui T 1998 Thin Solid Films 331 285
[54] Shi B Liu B Luo J Li Y Zheng C Yao X Zhang X 2017 Sol. Energy Mater Sol. Cells 168 214
[55] Fu F Kranz L Yoon S Löckinger J Jäger T Perrenoud J Feurer T Gretener C Buecheler S Tiwari A N 2015 Physica Status Solidi (a) 212 2708
[56] Mahmud M A Elumalai N K Upama M B Wang D Puthen-Veettil B Haque F Wright M Xu C Pivrikas A Uddin A 2017 Sol. Energy Mater. Sol. Cells 167 87
[57] Yang M Zhang T Schulz P Li Z Li G Kim D H Guo N Berry J J Zhu K Zhao Y 2016 Nat. Commun. 7 12305
[58] Zhou H Chen Q Li G Luo S Song T b Duan H S Hong Z You J Liu Y Yang Y 2014 Science 345 542
[59] Song T B Chen Q Zhou H Luo S Yang Y You J Yang Y 2015 Nano Energy 12 494
[60] Ono L K Leyden M R Wang S Qi Y 2016 J. Mater. Chem. 4 6693
[61] Wang S Ono L K Leyden M R Kato Y Raga S R Lee M V Qi Y 2015 J. Mater. Chem. 3 14631
[62] Bryant D G P Troughton J et al. 2014 Adv. Mater. Interfaces 26 7499
[63] Dualeh A Tétreault N Moehl T Gao P Nazeeruddin M K Grätzel M 2014 Adv. Funct. Mater. 24 3250
[64] Jiang Q Chu Z Wang P Yang X Liu H Wang Y Yin Z Wu J Zhang X You J 2017 Adv. Mater. 1703852
[65] Zhang T Yang M Zhao Y Zhu K 2015 Nano Lett. 15 3959
[66] Im J H Jang I H Pellet N Grätzel M Park N G 2014 Nat. Nanotech. 9 927