Mesoporous SnO2 nanoparticle films as electron-transporting material in perovskite solar cells

Yi Liab, Jun Zhu*a, Yang Huanga, Feng Liua, Mei Lva, Shuanghong Chena, Linhua Hua, Junwang Tang*c, Jianxi Yaod and Songyuan Dai*ad
aKey Laboratory of Novel Thin Film Solar Cells, Institute of Applied Technology, Hefei Institutes of Physical Science, Chinese Academy of Sciences, Hefei, 230031, P. R. China. E-mail: zhujzhu@gmail.com; sydai@ipp.ac.cn
bDepartment of Modern Physics, University of Science and Technology of China, Hefei 230026, P. R. China
cDepartment of Chemical Engineering, University College London, London WC1E 7JE, UK. E-mail: junwangtang@ucl.ac.uk
dBeijing Key Laboratory of Novel Thin Film Solar Cells, State Key Laboratory of Alternate Electrical Power System with Renewable Energy Sources, North China Electric Power University, Beijing, 102206, P. R. China

Received 26th January 2015 , Accepted 9th March 2015

First published on 9th March 2015


Abstract

Perovskite solar cells with mesoporous metal oxide films as scaffold layers have demonstrated very impressive advances in performance recently. Here, we present an investigation into mesoporous perovskite solar cells incorporating mesoporous SnO2 nanoparticle films as electron-transporting materials and scaffold layers, to replace traditional mesoporous TiO2 films. We have optimized the SnO2 film thickness and treated the surface of the SnO2 film with an aqueous solution of TiCl4. Due to the TiCl4 treatment the recombination process was significantly retarded. The short-circuit current density (Jsc) and open-circuit voltage (Voc) reached nearly 18 mA cm−2 and 1 V, respectively. Consequently, the power conversion efficiency of the device with the SnO2 film exceeded 10%.


1. Introduction

Organic–inorganic halide lead perovskite (CH3NH3PbX3, X = Cl, Br, I) materials have shown great potential in thin-film photovoltaic devices owing to their excellent light absorption coefficient, appropriate band gap, long diffusion length, and the ease by which these materials can be fabricated with a solution process.1–11 Recently, the power conversion efficiencies (PCEs) of mesoporous perovskite solar cells have skyrocketed from 3.8% to more than 20% within 5 years.1,5,10,11 In the perovskite solar cells based on the mesoporous configuration, mesoporous metal oxide thin films are used as either electron-transporting materials (ETMs), to accept photogenerated electrons from perovskites and then transmit them to the conductive substrates, or mesoporous scaffold layers for perovskite material formation. Mesoporous TiO2 nanoparticle films are usually used as ETMs and scaffold layers for high-efficiency perovskite solar cells. In addition to mesoporous TiO2 films, a variety of materials are also used in perovskite solar cells as ETMs and/or the scaffold. For example, one-dimensional TiO2 nanowires, rutile TiO2 nanorods, ZnO nanorods, [6,6]-phenyl-C61-butyric acid methyl ester (PCBM), Al2O3 and ZrO2, etc. have been successfully introduced into mesoporous solar cells with different PCEs.12–18 Despite the remarkable achievements in the field of perovskite solar cells, the development of alternative ETMs/scaffolds to TiO2 is a promising avenue to further improve the performance of perovskite solar cells.

Previously, mesoporous tin oxide (SnO2) thin film electrodes were widely applied in dye-sensitized solar cells (DSCs) as ETMs.19,20 SnO2 has a deeper conduction band than TiO2 and in principle should facilitate a more efficient transfer of photo-generated electrons from the perovskite light absorber to the SnO2 conduction band.21 Furthermore, bulk SnO2 has an electron mobility of up to 240 cm2 V−1 s−1, which is 100 times higher than that of TiO2,22 making it conceptually a more likely candidate for highly efficient solar cells. The DSCs based on SnO2 electrodes usually give a lower PCE compared to those based on TiO2 electrodes, due to the degree of recombination between the “bare” SnO2 and the hole-transporting materials (HTMs).19 However, by treating the surface of the SnO2 nanoparticles with an aqueous solution of TiCl4 or wide band gap “insulating” oxides, such as MgO or Al2O3, suppression of the backward reaction enables a significant enhancement in the PCE of the DSCs by over 7%.23 There are still no reports about mesoporous SnO2 being used as an ETM in perovskite solar cells despite the distinguished advantages. In this study, we present perovskite solar cells utilizing mesoporous SnO2 electrodes as the ETM. By optimizing the thickness of the SnO2 film and treating the surface of the SnO2 film with an aqueous solution of TiCl4, the short-circuit current density and open-circuit voltage have been increased to nearly 18 mA cm−2 and 1 V, respectively, resulting in devices with a PCE of over 10%. Furthermore, we employed impedance spectroscopy (IS) to investigate the recombination kinetics of devices based on mesoporous SnO2 electrodes before and after treatment of the SnO2 with an aqueous solution of TiCl4.

2. Results and discussion

We present a solid state mesoporous perovskite solar cell incorporating a mesoporous SnO2 film as the ETM, CH3NH3PbI3 as the light absorber and spiro-MeOTAD as the hole-transporting layer. A schematic structure of the device is depicted in Fig. 1. It is imperative that the mesoporous perovskite solar cells contain a compact blocking layer to retard the combination of the FTO substrate and the hole conductor.24 In this perovskite solar cell, mesoporous SnO2 thin films are used as both the ETMs, to accept photogenerated electrons from the CH3NH3PbI3 perovskites and then transmit them to the conductive substrate, and the mesoporous scaffold layer for perovskite material formation. We fabricated the CH3NH3PbI3 perovskite on the mesoporous SnO2 films via a two-step spin-coating procedure,25 as shown in Fig. 2. Firstly, the mesoporous SnO2 film was infiltrated with a PbI2-dissolved DMF solution by spin coating; the mesoporous SnO2 layer was completely covered by the PbI2 overlayer with a thickness of about 200 nm, as can be seen in Fig. 2b. Secondly, a 100 μL CH3NH3I-dissolved isopropyl alcohol solution was spin-coated onto the SnO2/PbI2 film. The reaction of PbI2 with CH3NH3I forms void-free CH3NH3PbI3 cuboids on the mesoporous SnO2 film, as shown in Fig. 2c.
image file: c5ra01540e-f1.tif
Fig. 1 Schematic device structure of the CH3NH3PbI3 perovskite solar cells using mesoporous SnO2 as the ETM.

image file: c5ra01540e-f2.tif
Fig. 2 SEM images of (a) the mesoporous SnO2 film; (b) the SnO2/PbI2 film; and (c) the SnO2/CH3NH3PbI3 film. The mesoporous SnO2 film thickness is about 200 nm.

In order to investigate the effect of the SnO2 film thickness on the photovoltaic performance of perovskite solar cells, we diluted the spin-coating SnO2 paste with ethanol in different weight ratios: 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]3 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5, to obtain the SnO2 electrodes with different thickness. The cross section SEM images of these SnO2 electrodes are shown in Fig. 3. As can be seen, the thicknesses of these SnO2 electrodes were 100 nm, 200 nm and 300 nm, respectively. Fig. 4 shows the absorption spectra of the SnO2/CH3NH3PbI3 films with SnO2 thickness of 100 nm, 200 nm and 300 nm, respectively. From the results shown in Fig. 4, it can be seen that the absorption spectra of the CH3NH3PbI3 goes up to 800 nm wavelength owing to its band gap of 1.57 eV. In addition, light absorption increases with increasing the thickness of the mesoporous SnO2 films from 100 nm to 300 nm, which is mainly due to the fact that increasing the SnO2 thickness increases the amount of CH3NH3PbI3 deposited into the SnO2 film.


image file: c5ra01540e-f3.tif
Fig. 3 Cross-sectional SEM images of the mesoporous perovskite solar cells using porous SnO2 films with different thicknesses as the ETMs. SnO2 films were obtained by spin-coating SnO2 paste diluted with ethanol to different weight ratios: (a) 1[thin space (1/6-em)]:[thin space (1/6-em)]1; (b) 1[thin space (1/6-em)]:[thin space (1/6-em)]3; (c) 1[thin space (1/6-em)]:[thin space (1/6-em)]5. The structures of the devices are FTO/SnO2/CH3NH3PbI3/spiro-MeOTAD/Au.

image file: c5ra01540e-f4.tif
Fig. 4 Effect of different mesoporous SnO2 films thicknesses on the absorption spectra of SnO2/CH3NH3PbI3 films.

The photovoltaic performance of the perovskite solar cells using a mesoporous SnO2 electrode as the ETM was evaluated. Firstly, the dependence of the photovoltaic performance on the thickness of the SnO2 electrode was investigated. The PCE of the FTO/SnO2/CH3NH3PbI3/spiro-MeOTAD/Au solar cells is slightly improved from 6.42% (100 nm) to 6.50% (200 nm), which is mainly due to the remarkable increase in Jsc from 16.78 mA cm−2 to 18.69 mA cm−2, though Voc and FF are slightly decreased from 0.753 V and 55.4% (100 nm) to 0.701 V and 53.4% (200 nm), respectively, as shown in Fig. 5a and Table 1. The increased Jsc and PCE might arise from the higher light absorption caused by the increased loading of CH3NH3PbI3 (Fig. 4). However, with a further increase in the thickness of the SnO2 electrode from 200 nm to 300 nm, the PCE of the perovskite solar cells is substantially decreased from 6.50% to 0.90%, which mainly results from the significant decrease in Voc, Jsc and FF from 0.701 V, 17.39 mA cm−2 and 53.4% to 0.272 V, 8.85 mA cm−2 and 37.3%, respectively (as shown in Fig. 5a and Table 1). The decreased efficiency may be explained by the fact that a thicker SnO2 film means a longer distance for charge carriers to travel which also increases the probability of electron–hole recombination. As has been known in the field, the morphology of the perovskite capping layer significantly affects the photovoltaic performance.25–28 The main reason for effecting the PCE of the perovskite solar cell is that the thinner CH3NH3PbI3 capping layer is prepared on the 300 nm-thick SnO2 electrode, as can be seen in Fig. 3. In Fig. 5b, the incident photon-to-electron conversion efficiency (IPCE) spectra follow the same trend observed for the short-circuit photocurrent with the SnO2 based perovskite solar cells. Specially, the perovskite solar cells based on a SnO2 thin film with 200 nm thickness exhibit IPCE values of over 80% between 400 and 600 nm. Integrating this IPCE spectrum over the AM 1.5 solar spectra at 100 mW cm−2 estimates a Jsc of 17.59 mA cm−2, which is in close agreement with our measured maximum value of 17.39 mA cm−2 under simulated solar conditions.


image file: c5ra01540e-f5.tif
Fig. 5 (a) Current density–voltage (JV) characteristics of the perovskite solar cells depending on the thickness of the SnO2 electrodes. (b) The corresponding IPCE spectra.
Table 1 Photovoltaic performance parameters of the perovskite solar cells depending on the thickness of the SnO2 electrodes
Thickness of SnO2 (nm) Voca (V) Jscb (mA cm−2) FFc PCEd (%)
a Open-circuit photovoltage.b Short-circuit photocurrent.c Fill factor.d Power conversion efficiency.
100 0.753 15.39 0.554 6.42
200 0.701 17.39 0.534 6.50
300 0.272 8.85 0.373 0.90


For the spiro-MeOTAD-based perovskite solar cell composed of a “bare” SnO2 electrode as the ETM, we found that the performance of this device is not satisfactory, which may be caused by the degree of electron recombination occurring between the SnO2 and HTMs or perovskite. In order to inhibit the charge recombination, we optimized the interface between the SnO2 and perovskite layer via treatment with an aqueous solution of TiCl4.24 The photovoltaic performance of the perovskite solar cells with the mesoporous SnO2 electrode, with and without treating the SnO2 film surface with an aqueous solution of TiCl4, was evaluated, as can be seen in Fig. 6 and Table 2. Upon treating the SnO2 film surface with an aqueous solution of TiCl4, the PCE of the FTO/SnO2/CH3NH3PbI3/spiro-MeOTAD/Au solar cell is substantially improved from 6.50% to 10.18% (33% increment), which is mainly due to the significant increase in Voc from 0.701 V to 0.933 V, with a slight increase in FF from 53.4% to 62.8% and a nearly equal Jsc value.


image file: c5ra01540e-f6.tif
Fig. 6 Current–voltage curves for the mesoporous perovskite solar cells with and without treating the SnO2 film surface with an aqueous solution of TiCl4. The thickness of the SnO2 electrodes is 200 nm.
Table 2 Photovoltaic performance parameter values, extracted from the current–voltage curves presented in Fig. 6
Surface treatment of SnO2 Voc (V) Jsc (mA cm−2) FF PCE (%)
None 0.701 17.39 0.534 6.50
TiCl4 0.933 17.38 0.628 10.18


We employed impedance spectroscopy (IS) to investigate the effect of the TiCl4 treatment on the dramatic improvement of the performance of the SnO2-based perovskite solar cells. Fig. 7a shows representative IS spectra (Nyquist plots) for the SnO2-based perovskite solar cells before and after TiCl4 treatment at low applied forward bias (0.7 V) under dark conditions. The obtained IS spectra includes two arcs, where the first arc in the high frequency region is related to the charge transfer behavior at the counter electrode and the second arc is due to a combination of the recombination resistance (Rrec) and the chemical capacitance of the film (Cμ).24,29,30 Excellent fitting results (Fig. 7a) were obtained using a simplified equivalent circuit (Fig. 7b).31 The Rrec for a perovskite solar cell incorporating a “bare” SnO2 electrode shows a lower value than for that incorporating a SnO2 electrode coated with an ultrathin TiO2 layer at the same applied bias voltage, which indicates that the recombination process in the device based on a TiO2-coated SnO2 electrode is remarkably retarded via TiCl4 treatment. Thus, the perovskite solar cells with a “bare” SnO2 electrode represent a lower open-circuit voltage due to the higher recombination. Likewise, by treating the SnO2 electrode surface with an aqueous solution of TiCl4, the device shows lower recombination kinetics due to the ultrathin TiO2 between the SnO2 and CH3NH3PbI3 layers enabling ease of electron transfer from the TiO2 to the SnO2 and also avoiding any extra internal trap sites,24 and thus decreasing the charge recombination, leading to a 230 mV higher Voc and a 10% higher FF than that exhibited by the perovskite solar cell incorporating a “bare” SnO2 electrode.


image file: c5ra01540e-f7.tif
Fig. 7 (a) Nyquist plots for the perovskite solar cells with mesoporous SnO2 as the ETM before and after treating the SnO2 electrode surface with an aqueous solution of TiCl4, measured under dark conditions at 0.7 V bias. The thickness of the SnO2 electrodes is 200 nm. (b) A simplified equivalent circuit model employed for impedance analysis of the perovskite solar cells.

3. Conclusions

In summary, a mesoporous SnO2 film has been successfully utilized in the CH3NH3PbI3 perovskite solar cell as an electron-transporting material and scaffold layer. By optimizing the SnO2 thickness and treating the surface of the SnO2 film with an aqueous solution of TiCl4, the perovskite solar cell exhibits an increased Voc of around 1 V and a PCE of 10.18%. Impedance spectroscopy results indicated that the ultrathin TiO2 coating the SnO2 film as a result of TiCl4 treatment significantly retards the charge recombination process via avoiding any extra internal trap sites and facilitating electron transfer from the CH3NH3PbI3 perovskite to the SnO2 conduction band. These results make the mesoporous SnO2 based perovskite solar cells competitive with the mesoporous TiO2 based devices, and a much more promising concept for future optimization of perovskite solar cells.

4. Experimental section

4.1. Materials

Lead(II) iodide (PbI2, 99%), lithium bis(trifluoromethylsulphonyl) imide (Li-TFSI) (99.95%), 4-tert-butylpyridine (t-BP, 96%), butyl-(tin chloride), N,N-dimethylformamide (DMF, 99.9%) and hydriodic acid (57 wt% in water) were purchased from Sigma-Aldrich. Methylamine solution (40% in methanol) was purchased from TCI. Tin(IV) oxide (SnO2) nanoparticle powders (NanoArc) were purchased from Alfa Aesar and ranged in size from 22 to 43 nm. Spiro-MeOTAD was purchased from Merck KGaA. All chemicals were used as received.
Synthesis of CH3NH3I. CH3NH3I was synthesized according to the reported method.6 Hydroiodic acid (30 mL, 57 wt% in water) was reacted with methylamine (27.86 mL, 40% in methanol) in a 250 mL round-bottom flask at 0 °C for 2 h. The precipitate was recovered by putting the solution on a rotary evaporator and carefully removing the solvent at 50 °C for 1 h. The generated yellowish powder was dissolved in ethanol, recrystallized from diethyl ether, and finally dried at 60 °C in a vacuum oven for 24 h.

4.2. Solar cell fabrication

Substrate preparation. Fluorine-doped tin oxide (FTO, 15 Ω square−1) glass substrates with dimensions of 2 cm × 1.5 cm were partially etched with Zn powder and 2 M HCl to reveal the electrode pattern. The etched substrates were cleaned with detergent, followed by ultrasonication in pure water and ethanol for 30 min, and were then rinsed with deionized water and ethanol and dried by air. Finally, the substrates were annealed at 500 °C for 30 min. A thin (60 nm) blocking layer of SnO2 (bl-SnO2) was deposited via spray pyrolysis deposition at 450 °C from a solution of butyl-(tin chloride) in anhydrous ethanol at a 1[thin space (1/6-em)]:[thin space (1/6-em)]10 volume ratio and then annealed at 500 °C for 30 min.

SnO2 nanoparticle powders were used to synthesize the SnO2 paste by following the method for making TiO2 paste.32 Briefly, 5 g SnO2 nanoparticles was redispersed in 150 mL of anhydrous ethanol and mixed with 20 g of terpineol and 15 g of ethyl cellulose to prepare the SnO2 paste. The mesoporous SnO2 electrodes were obtained by spin-coating at 5000 rpm for 30 s onto the bl-SnO2 substrates using the SnO2 paste diluted in ethanol (1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]3 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5, weight ratio). After drying at 100 °C, the resulting mesoporous SnO2 films were annealed at 500 °C for 30 min to remove the organic part. Some of SnO2 films resulting from the SnO2 paste diluted in ethanol (1[thin space (1/6-em)]:[thin space (1/6-em)]2, weight ratio) were treated with a 0.04 M aqueous solution of TiCl4 at 60 °C for 1 h, rinsed with deionized water and annealed at 500 °C for 30 min. Prior to their use, the SnO2 films were again dried at 500 °C for 30 min.

462 mg of PbI2 was dissolved in 1 mL DMF under stirring at 70 °C overnight, followed by filtering with a 0.22 μm pore PVDF syringe filter. The solution was kept at 70 °C during the whole procedure. A 25 μL PbI2 solution was spin-coated on the mesoporous SnO2 films at 3000 rpm for 20 s, and dried at 50 °C for 3 min and 100 °C for 5 min, consecutively. After cooling to room temperature, a 100 μL CH3NH3I solution in 2-propanol (10 mg mL−1) was loaded onto the PbI2-coated SnO2 films for 20 s, which was spun at 4000 rpm for 30 s and then dried at 100 °C for 5 min.

The spiro-MeOTAD HTM was deposited on the SnO2/CH3NH3PbI3 film by spin coating at 4000 rpm for 30 s. The composition of the spiro-MeOTAD HTM was 72.3 mg spiro-MeOTAD, 28.8 μL TBP, and 17.5 μL of a solution of 520 mg mL−1 LiTFSI in acetonitrile in 1 mL chlorobenzene.

Finally, 60 nm-thick Au was thermally evaporated on top of the device to form the back contact. The active area of the devices was 9 mm2 determined by a black mask with dimensions of 3 mm × 3 mm.

4.3. Characterization

The surface morphology of the film was observed with a field emission scanning electron microscope (FE-SEM, sirion200, FEI Corp., Holland). The UV-vis spectrum of the films was obtained using a UV-vis spectrophotometer (U-3900H, HITACHI, Japan). The incident-photon-to-electron conversion efficiency (IPCE) measurement was conducted using a QE/IPCE measurement kit (Newport Corporation, CA). The current–voltage characteristics (JV curves) were measured with a Keithley model 2420 digital source meter (Keithley Instruments, Inc., USA) under the illumination of 100 mW cm−2 (AM 1.5) provided by a solar simulator (solar AAA simulator, Oriel USA). The impedance spectra were measured with an electrochemical analyzer (Autolab 320, Metrohm, Switzerland) with a bias potential at 0.7 V. AC 20 mV perturbation was applied with a frequency from 1 MHz to 1 Hz. The obtained impedance spectra were fitted with ZView software (v2.8b, Scribner Associates, USA).

Acknowledgements

This work is financially supported by the National Basic Research Program of China under Grant no. 2011CBA00700 and the National Natural Science Foundation of China under Grant no. 21403247, 21173228, 61204075 and 61404142.

Notes and references

  1. A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Soc., 2009, 131, 6050–6051 CrossRef CAS PubMed .
  2. H.-S. Kim, C.-R. Lee, J.-H. Im, K.-B. Lee, T. Moehl, A. Marchioro, S.-J. Moon, R. Humphry-Baker, J.-H. Yum, J. E. Moser, M. Gratzel and N.-G. Park, Sci. Rep., 2012, 2, 591 Search PubMed .
  3. J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M. Gratzel, Nature, 2013, 499, 316–319 CrossRef CAS PubMed .
  4. S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. P. Alcocer, T. Leijtens, L. M. Herz, A. Petrozza and H. J. Snaith, Science, 2013, 342, 341–344 CrossRef CAS PubMed .
  5. H. Zhou, Q. Chen, G. Li, S. Luo, T.-b. Song, H.-S. Duan, Z. Hong, J. You, Y. Liu and Y. Yang, Science, 2014, 345, 542–546 CrossRef CAS PubMed .
  6. J. H. Im, C. R. Lee, J. W. Lee, S. W. Park and N. G. Park, Nanoscale, 2011, 3, 4088–4093 RSC .
  7. J. M. Ball, M. M. Lee, A. Hey and H. J. Snaith, Energy Environ. Sci., 2013, 6, 1739–1743 CAS .
  8. H.-S. Kim, I. Mora-Sero, V. Gonzalez-Pedro, F. Fabregat-Santiago, E. J. Juarez-Perez, N.-G. Park and J. Bisquert, Nat. Commun., 2013, 4, 2242 Search PubMed .
  9. G. Xing, N. Mathews, S. Sun, S. S. Lim, Y. M. Lam, M. Grätzel, S. Mhaisalkar and T. C. Sum, Science, 2013, 342, 344–347 CrossRef CAS PubMed .
  10. http://www.nrel.gov/ncpv/images/efficiency_chart.jpg, The National Renewable Energy Laboratory (NREL), 2015.
  11. W. Nie, H. Tsai, R. Asadpour, J.-C. Blancon, A. J. Neukirch, G. Gupta, J. J. Crochet, M. Chhowalla, S. Tretiak, M. A. Alam, H.-L. Wang and A. D. Mohite, Science, 2015, 347, 522–525 CrossRef CAS PubMed .
  12. H.-S. Kim, J.-W. Lee, N. Yantara, P. P. Boix, S. A. Kulkarni, S. Mhaisalkar, M. Grätzel and N.-G. Park, Nano Lett., 2013, 13, 2412–2417 CrossRef CAS PubMed .
  13. M. H. Kumar, N. Yantara, S. Dharani, M. Graetzel, S. Mhaisalkar, P. P. Boix and N. Mathews, Chem. Commun., 2013, 49, 11089–11091 RSC .
  14. J. You, Z. Hong, Y. Yang, Q. Chen, M. Cai, T.-B. Song, C.-C. Chen, S. Lu, Y. Liu and H. Zhou, ACS Nano, 2014, 8, 1674–1680 CrossRef CAS PubMed .
  15. K. Wojciechowski, M. Saliba, T. Leijtens, A. Abate and H. J. Snaith, Energy Environ. Sci., 2014, 7, 1142–1147 CAS .
  16. D. Bi, S.-J. Moon, L. Haggman, G. Boschloo, L. Yang, E. M. J. Johansson, M. K. Nazeeruddin, M. Gratzel and A. Hagfeldt, RSC Adv., 2013, 3, 18762–18766 RSC .
  17. X. Zhang, Z. Bao, X. Tao, H. Sun, W. Chen and X. Zhou, RSC Adv., 2014, 4, 64001–64005 RSC .
  18. K. Manseki, T. Ikeya, A. Tamura, T. Ban, T. Sugiura and T. Yoshida, RSC Adv., 2014, 4, 9652–9655 RSC .
  19. A. N. Green, E. Palomares, S. A. Haque, J. M. Kroon and J. R. Durrant, J. Phys. Chem. B, 2005, 109, 12525–12533 CrossRef CAS PubMed .
  20. A. Kay and M. Grätzel, Chem. Mater., 2002, 14, 2930–2935 CrossRef CAS .
  21. M. Gratzel, Nature, 2001, 414, 338–344 CrossRef CAS PubMed .
  22. S. Gubbala, V. Chakrapani, V. Kumar and M. K. Sunkara, Adv. Funct. Mater., 2008, 18, 2411–2418 CrossRef CAS .
  23. M. K. I. Senevirathna, P. K. D. D. P. Pitigala, E. V. A. Premalal, K. Tennakone, G. R. A. Kumara and A. Konno, Sol. Energy Mater. Sol. Cells, 2007, 91, 544–547 CrossRef CAS PubMed .
  24. H. J. Snaith and C. Ducati, Nano Lett., 2010, 10, 1259–1265 CrossRef CAS PubMed .
  25. J.-H. Im, I.-H. Jang, N. Pellet, M. Grätzel and N.-G. Park, Nat. Nanotechnol., 2014, 9, 927–932 CrossRef CAS PubMed .
  26. F. Huang, Y. Dkhissi, W. Huang, M. Xiao, I. Benesperi, S. Rubanov, Y. Zhu, X. Lin, L. Jiang, Y. Zhou, A. Gray-Weale, J. Etheridge, C. R. McNeill, R. A. Caruso, U. Bach, L. Spiccia and Y.-B. Cheng, Nano Energy, 2014, 10, 10–18 CrossRef CAS PubMed .
  27. G. E. Eperon, V. M. Burlakov, P. Docampo, A. Goriely and H. J. Snaith, Adv. Funct. Mater., 2014, 24, 151–157 CrossRef CAS .
  28. J.-H. Im, H.-S. Kim and N.-G. Park, APL Mater., 2014, 2, 081510 CrossRef PubMed .
  29. E. M. Barea, M. Shalom, S. Giménez, I. Hod, I. Mora-Seró, A. Zaban and J. Bisquert, J. Am. Chem. Soc., 2010, 132, 6834–6839 CrossRef CAS PubMed .
  30. V. González-Pedro, X. Xu, I. Mora-Seró and J. Bisquert, ACS Nano, 2010, 4, 5783–5790 CrossRef PubMed .
  31. M. A. Hossain, J. R. Jennings, Z. Y. Koh and Q. Wang, ACS Nano, 2011, 5, 3172–3181 CrossRef CAS PubMed .
  32. L. H. Hu, S. Y. Dai, J. Weng, S. F. Xiao, Y. F. Sui, Y. Huang, S. H. Chen, F. T. Kong, X. Pan and L. Y. Liang, J. Phys. Chem. B, 2007, 111, 358–362 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry 2015