Cubic structure of the mixed halide perovskite CH3NH3PbI3−xClx via thermal annealing

Deying Luo§ *a, Leiming Yua, Hai Wang*a, Taoyu Zoub, Li Luob, Zhu Liu*b and Zhenghong Lubc
aKey Laboratory of Yunnan Provincial Higher Education Institutions for Organic Optoelectronic Materials and Devices, Kunming University, Kunming 650214, China. E-mail: deyingluo123@outlook.com; hai.wang.ucl@gmail.com
bDepartment of Physics, Yunnan Key Laboratory of Micro/Nano-Materials and Technology, Yunnan University, Kunming, Yunnan 650091, China. E-mail: zhuliu@ynu.edu.cn
cDepartment of Materials Science and Engineering, University of Toronto, Toronto, Ontario M5S 3E4, Canada

Received 16th August 2015 , Accepted 21st September 2015

First published on 22nd September 2015


Abstract

A methodology has been developed to obtain a cubic structure of the mixed halide perovskite CH3NH3PbI3−xClx that involves thermal annealing of a vacuum-deposited perovskite layer. In this process, a tetragonal–cubic transition of the mixed halide perovskite CH3NH3PbI3−xClx has been observed using a variety of characterization techniques, e.g., X-ray diffraction (XRD), scanning electron microscopy (SEM), and optical and Hall effect measurements. According to the XRD, the tetragonal structure of CH3NH3PbI3−xClx changed into the cubic phase when the temperature increased to above the transition temperature, at which point all iodine atoms move towards the central axis while PbI6 rotates around the C–N bond. After annealing, the stable cubic structure of the mixed halide perovskite CH3NH3PbI3−xClx has a smaller band gap and a better optical absorption than the original tetragonal structure. Meanwhile, the microstructure has shown an increase of grain size after annealing, which could be given the fastest mobility of 13.5 cm2 V−1 S−1, yielded from the Hall effect measurements. The potential benefits of the stable cubic structure and large grain size in the mixed halide perovskite CH3NH3PbI3−xClx are to improve the device performance of solar cells from the viewpoint of bandgap engineering and crystalline quality.


1. Introduction

Recently, the highest record of power conversion efficiency (PCE) > 20% has been achieved in emerging solar cells by the use of absorbers of the lead halide perovskite CH3NH3PbX3 (X = Cl, Br, I and mixed).1 The reason for such high performance is due to the excellent properties of polycrystalline perovskites, e.g., direct-optical absorption2 and balanced bipolar long carrier diffusion length (∼1 μm),3 which is several orders longer than that for other emerging semiconductors in photovoltaics.4 Hence, in order to further enhance the device efficiency, larger grain sizes are one of the critical factors to ensure excellent optical absorption and low non-radiative recombination within the thicker perovskite absorbers.5,6 In addition, the unique hysteretic current–voltage behaviours of perovskite solar cells are also closely related to the hybrid organic–inorganic crystal structure and their surface defect states.7,8

In principle, the symmetry of the atomic structure within the crystal can have an influence on the band gap and optical absorption range of perovskite absorbers.9–12 CH3NH3PbI3 has three stable crystal structures, i.e., cubic, tetragonal and orthorhombic,13–15 of which the former provides the highest symmetry structure and stability at high temperatures.10,16 It was found that along with the crystal structure transforming from a lower symmetry structure (orthorhombic) to a higher symmetry structure (tetragonal) in CH3NH3PbI3, its band gap decreases and hence enables greater absorption for a much longer wavelength.9,12,15 Thus, it is expected that with further transformation into a cubic structure that has an increased symmetry, the optical band gap of CH3NH3PbI3 would further decrease, thereby matching the solar spectrum better and lead towards an enhanced optical absorption. However, for most CH3NH3PbI3 devices, only the use of a tetragonal crystal structure obtained by either solution processing17–19 or vacuum deposition20 has been reported. None of them have considered the approach of a thermal annealing-induced phase-transition to prepare the cubic structure of CH3NH3PbI3−xClx perovskite for the improvement of perovskite solar cells.

In general, doping a small amount of bromine (Br) or chlorine (Cl) within the CH3NH3PbI3 thin films effectively enables the phase transition from a tetragonal to a cubic structure.21,22 In addition, it is confirmed that the incorporation of Cl within CH3NH3PbI3 increases the carrier mobility23 and diffusion length3 as well as the carrier transport at the interface.24

Here, the cubic structure and the larger grain size have been successfully obtained for the vacuum-deposited CH3NH3PbI3−xClx layer through a proper thermal annealing. During annealing, the iodine atoms move towards the central axis while the PbI6 metal halide framework rotates around the C–N bond (i.e., the crystal axis) in the tetragonal structure, and finally reaches the stable cubic structure at the phase transition temperature. As the temperature increases further, the PbI2 phase separates from the bulk CH3NH3PbI3−xClx perovskite, owing to the process of bond breaking of organic cations (CH3NH3+) away from the PbI6 metal halide framework. Moreover, based on observations, an enhanced optical absorption with a red-shifted absorption wavelength has been achieved by transforming the crystal structure from tetragonal to cubic after annealing at 90 °C. Transport properties from Hall effect measurements suggested that the microstructure and grain sizes both have an impact on the carrier mobility, whereas the highest mobility (13.5 cm2 V−1 S−1) was achieved after annealing at 100 °C due to the reduced defect density.

2. Experimental

2.1 CH3NH3PbI3−xClx fabrication

CH3NH3I (99.999% purity) and PbCl2 (99.999% purity) were both received from commercial outlets. Before vacuum deposition, the substrates were ultrasonically cleaned in sequence with a detergent and methanol for 5 min, and subsequently these substrates were further treated with UV-ozone for 15 min. Then the substrates were fixed on a sample holder and transferred into a vacuum chamber, and two crucibles were heated as the base pressure (Mini-spectros, Kurt. J. Lesker) reached to <5.6 × 10−6 mbar. CH3NH3I (99.999% purity) and PbCl2 (99.999% purity) were heated in alumina crucibles, and 500 mg and 200 mg were employed, respectively, each time. During the deposition process the molar ratio of CH3NH3I to PbCl2 was fixed at 3[thin space (1/6-em)]:[thin space (1/6-em)]1 to ensure the formation of stable CH3NH3PbI3−xClx, and the thickness was monitored using a calibrated quartz crystal microbalance and further calibrated via cross-sectional measurements. After finishing the deposition, all thin films were thermally annealed in a nitrogen-filled glove-box. In our work, the as-grown thin films were annealed at 70 °C, 80 °C, 90 °C, 100 °C, 110 °C and 120 °C for 1 hour. After thermal annealing, most of the thin films change to reddish-brown colours from shallow oranges, while the film annealed at 120 °C did not show the same observation. Finally, as the annealed thin films cooled down to room temperature, some of the samples were transferred to a vacuum chamber again and metallic electrode deposition (Ag) was completed though a shallow mask to satisfy the Hall effect measurement, while the rest of the samples were characterized using various measuring methods. The thickness of the CH3NH3PbI3−xClx film is 530 nm.

2.2 Characterization

The films were characterized using X-ray diffraction (XRD), Scanning Electron Microscopy (SEM), UV-visible light absorption and Hall effect measurements. X-ray diffraction characterization was performed (XRD, Rigaku TTR III) using Cu Kα (λ = 1.5406 Å) radiation. The surface morphologies of the CH3NH3PbI3−xClx films were characterized using scanning electron microscopy (SEM, FEI Quanta 200), and their band structure was studied using photoluminescence (PL) spectroscopy (iHR320, Horiba JY). Moreover, UV-visible light absorption of CH3NH3PbI3−xClx was measured using AnalytikJena AG, and the Hall mobilities were obtained using a Nanometrics HL5500 Hall system (notice that a 3.8 nA DC mean current and a 200 mV applied voltage were used when performing the Hall effect measurements).

3. Results and discussion

Pure CH3NH3PbI3 has a tetragonal structure at room temperature,15,16 yet can undergo a tetragonal–cubic transition at around 54.4 °C.14 Typically, CH3NH3PbI3 obtained from solution processing after thermal annealing (>80 °C) shows a tetragonal structure.5 After doping with elements that have smaller atomic radii, such as bromine (Br) and chlorine (Cl), a halide perovskite with a cubic structure could be achieved without affecting the optical absorption and electrical properties.11,21,22,25 Details of the crystal structure are compared in Fig. 1(a) and (b) according to the reported atomic coordinates in the unit cells,15 in which the organic cations (CH3NH3+) and PbI6 are located in different sites, and the orientation of CH3NH3+ is distinct within the two phases. Theoretically calculated XRD results are plotted in Fig. S(1 and 2). It is noted that the calculated results are consistent with those experimentally observed for both the cubic and tetragonal crystal structures.21,26,27 The band gap of the cubic structure of CH3NH3PbI3 is around 1.55 eV,15 while the band gap of CH3NH3PbI3−xClx with a cubic structure is 1.57 eV.12
image file: c5ra16516d-f1.tif
Fig. 1 Crystal structure of the mixed halide perovskite with a cubic phase (a) and (b) a tetragonal phase, where black spheres represent lead, purple spheres are iodine and the others represent the organic cations (CH3NH3+).

Thermal treatment provides a route to control both the crystalline structure and grain size. Fig. 2(a) and (b) illustrate the XRD evolution of the CH3NH3PbI3−xClx perovskite on the as-grown thin films which are treated by post-annealing in a nitrogen-filled glove box at 80 °C, 90 °C, 100 °C and 110 °C for 1 hour. The as-grown CH3NH3PbI3−xClx perovskite shows the tetragonal-structured characteristic peaks of (002), (110), (004) and (220).14,21,28 Additionally, in the mixed halide perovskite films, it is found that the dominant component of the thin films is CH3NH3PbI3 not CH3NH3PbI3−xClx, which is consistent with the literature20,29–31 as only a small fraction of Cl atoms was observed from the EDS measurements, see the ESI (Table S1). After the thin films were annealed at 80 °C for 1 hour, the (110) and (220) peaks decreased while the (002) and (004) peaks became dominant.


image file: c5ra16516d-f2.tif
Fig. 2 XRD (a and b) patterns of the perovskite layer before and after thermal annealing at 80 °C, 90 °C, 100 °C, and 110 °C, the stars represent the peaks of the substrates.

As the annealing temperature increased to 90 °C, peaks corresponding to the tetragonal structure completely disappeared, instead, the (100) and (210) peaks of the cubic structure for CH3NH3PbI3−xClx emerged, which indicates a phase transition from the tetragonal to the cubic structure (i.e., space group Pm[3 with combining macron]m, lattice constant is 6.276 Å). Moreover, a strong (001) peak of PbI2 also appeared. As the annealing temperature increased much higher, to 100 °C and 110 °C, the peak intensity of PbI2 (001) become much higher, indicating that more of the CH3NH3PbI3−xClx perovskite thin film dissociated into the CH3NH3I gas and PbI2 crystals due to the chemical bonds between the organic cations and the PbI6 cages being broken at a high temperature. The color of the CH3NH3PbI3−xClx films (see in Fig. S3) changed from a reddish-dark-brown to a yellow colour as the annealing temperature increased from 90 °C to 110 °C, further confirming that the optical absorption of the films is reduced. Some pioneering work predicted that PbI2 existing at grain boundaries can function as a passivator32 and facilitates carrier transport.29,33 The bi-phasic materials of the perovskite phase and PbI2 within the films have been observed from low magnification SEM images in Fig. S4(b and c) as the annealing temperature was over 90 °C.

For the lead halide perovskite film, there are three major dynamic mechanisms for the phase transition. (i) Organic cations (CH3NH3+) move towards the C–N bonds or with respect to the crystal axis with the temperature and rotation often enhanced during the structural phase transition from a lower symmetry to a higher symmetry;10 (ii) the orientation of highly ordered CH3NH3+ along the different directions changes34 and dynamic disorder appears in the cubic phase;16 (iii) it is often accompanied by the rotation of PbI6 cages around the crystal axis.14

In our case, the phase transition of the structure is majorly attributed to the temperature-induced rotation of slightly distorted PbI6 octahedra along the crystal-axis of (001).16 Fig. 3(a)–(c) show the atomic arrangements on the (100), (002) and (110) lattice planes of the tetragonal and cubic structures. It is found that atoms on the (002) and (110) lattice planes of the tetragonal structure are iodines, while the (100) plane is composed of Pb atoms and organic cations (CH3NH3+) for the cubic structure. But as the as-grown thin film of the tetragonal-structured CH3NH3PbI3−xClx perovskite was heated, all the iodides on the (110) lattice planes started to move towards the crystal-axis, and led to the symmetric structure of the PbI6 cages and a reduction in the chemical bond lengths between the atoms (see in Fig. 3(a)), resulting in the dominant diffraction peaks changing from (110) to (002) as the annealing temperature increased to 80 °C. As the temperature further increased from 80 °C to 90 °C, all iodine atoms aligned with the central axis owing to the PbI6 cages rotating around the crystal-axis, which resulted in the formation of perfect PbI6 octahedra in such crystals, i.e., a cubic structure.


image file: c5ra16516d-f3.tif
Fig. 3 Atomic arrangements (a–c) on the (110) and (002) lattice planes of the tetragonal structure and the (100) lattice plane of the cubic structure.

Fig. 4 shows the surface morphology of the CH3NH3PbI3−xClx perovskite thin film characterized after annealing at the different temperatures. For the as-grown thin films, the surface shows a crack-like morphology with a few isolated CH3NH3PbI3−xClx islands, and the grain size is around less than 500 nm. After the annealing temperature increased to 70 °C and 80 °C, some isolated islands started to connect into a continuous film. After the annealing temperature reached 80 °C, the isolated islands merged together to form uniform thin films.


image file: c5ra16516d-f4.tif
Fig. 4 SEM images of the vacuum-deposited MAPbI3−xClx perovskite layer before (a) and after thermal annealing (b–f) at 70 °C, 80 °C, 90 °C, 100 °C and 110 °C.

As the annealing temperature reached 90 °C, not only did the structure change from tetragonal to cubic, but also the grain size became larger. As the temperature reached 100 °C, the largest grain size of around 5 microns was obtained but a void was also seen in Fig. 4(e). This void could have originated from the dissociation of CH3NH3PbI3−xClx, as discussed with respect to the X-ray diffraction. As the temperature increased to 110 °C, smaller grains emerged again which could be due to the large amount of PbI2 crystals dissociated from the bulk sample. Furthermore, the coexistence of PbI2 and CH3NH3PbI3−xClx within the thin films after annealing at temperatures of over 90 °C has also been confirmed with high contrast SEM images at 100 °C and 110 °C (Fig. S4(c and d)), as bright crystals of PbI2 were observed.

The band gaps of the CH3NH3PbI3−xClx films were determined using the photoluminescence (PL) spectrum as shown in Fig. 5(a). Compared to the as-grown film and the film after annealing at 80 °C, the red-shift of the PL peak is due to the increased grain size.9,35 As the annealing temperature changed from 80 °C to 90 °C, which corresponds to the phase transition from tetragonal to cubic, the PL peak position red-shifted from 768 nm to 778 nm, the red-shift is around 10 nm. The change in the PL peak position is due to variation in the band-edge emission, not the different grain-induced defect states36,37 and the onset of the light absorption demonstrated an enhanced absorption at the long wavelengths in Fig. 5(b) for the cubic CH3NH3PbI3−xClx. This confirmed that the higher symmetry cubic structure has a smaller band gap compared to the structure of CH3NH3PbI3−xClx. Similar observations have been reported for orthorhombic–tetragonal transitions.9,13


image file: c5ra16516d-f5.tif
Fig. 5 Photoluminescence (PL) spectra (a) and light absorption spectra (b) of the vacuum-deposited CH3NH3PbI3−xClx perovskite layer (530 nm) on the quartz substrates before and after thermal annealing at 80 °C, 90 °C, 100 °C and 110 °C, with an excitation wavelength of 467 nm, where the dashed line indicates the PL peaks.

As the annealing temperature increased to 100 °C, the PL peak position exhibits almost no change at 778 nm, but its intensity enhanced and the width of the PL peak reduced, which means that thin films with annealing temperatures at 100 °C have a much lower density of defect states. For the optical absorption in Fig. 5(b) and Fig. S(5 and 6) in the ESI, a second absorption peak, which corresponds to PbI2, also emerged after annealing at 100 °C. As the temperature further increased to 110 °C, the PL peak blue shifts to 774 nm. This phenomenon is largely due to the smaller grain size37 and the emerging phase of PbI2.4 Such a change also results in colour variation of the samples, photographed in Fig. S3. Hence, the corresponding PL and absorption of the films are consistent with the change of grain size and the phase transition after annealing at the different temperatures.

The optical absorption of the CH3NH3PbI3−xClx film benefits from the higher symmetry cubic structure and the large grain size after annealing. The corresponding change in the carrier mobility as the annealing temperature increased was further characterized using Hall effect measurements. A Hall effect measurements setup is plotted in Fig. 6(a) and during the measurements an applied magnetic field was rotated to eliminate the effect of the irregular metallic Ag electrode on the measured results.


image file: c5ra16516d-f6.tif
Fig. 6 Hall effect measurement setup (a), and (b) the mobility and resistivity both plotted against the thermal annealing temperature.

Fig. 6(b) shows CH3NH3PbI3−xClx as a p-type semiconductor and the mobility/resistivity as a function of the annealing temperature. The mobility reaches a maximum value of 13.5 cm2 V−1 S−1 after annealing at 100 °C and this value is even larger than the mobility from other solution processes.23 The reason is that the CH3NH3PbI3−xClx films have different grain sizes and structures as well as compositions as the thin films were annealed at different temperatures. Obviously, whether the thin films undergo annealing at 70–80 °C or not, the surface morphologies of the thin films have changed considerably but the changes are negligible for the grain size and crystal structure. Therefore, the observed enhancement of mobility (or decrease in resistivity) is primarily contributed by the microstructure. When the temperature rose to 90–100 °C and even 110 °C, all the perovskite in the thin films transformed into a cubic structure, along with a gradual increase in the amount of PbI2. The largest mobility was achieved when the thin films were annealed at 100 °C, indicating that the large grain size within the CH3NH3PbI3−xClx film strongly contributes to the mobility, while an appropriate amount of PbI2 and a low density of defect states are pre-requisites for high mobility. In short, for the cubic-structured CH3NH3PbI3−xClx perovskite, controllable mobility can be achieved by tuning the microstructure and grain size within the thin films after annealing at an appropriate temperature.

4. Conclusions

In summary, the cubic phase for CH3NH3PbI3−xClx perovskite has been achieved after thermal annealing at 90 °C, and it is found that the dominant diffraction peaks are different from the as-grown thin films annealed at 80 °C. The fundamental reason for such changes is due to the rotation of the octahedral cages of PbI6 around the (001) c-axes with increasing temperature which leads to iodine atoms on the (110) plane moving towards the central axis. The results confirmed that the phase transition was accompanied by changes in the microstructure and grain size post-annealing, which in turn govern its optical absorption and non-radiative recombination as well as the defect states. Finally, the highest mobility of 13.5 cm2 V−1 S−1 and a longer absorption wavelength of 778 nm have been achieved for the cubic phase after annealing at 100 °C. As such, the cubic-structured CH3NH3PbI3−xClx thin films with large crystalline structures and a small amount of PbI2 at room temperature ensured a low density of defect states and non-radiative recombination. As a result, this paves the way to yielding excellent optoelectronic properties for vacuum-deposited CH3NH3PbI3−xClx mixed halide perovskites by adjusting post-annealing conditions.

Acknowledgements

We acknowledge the funding support for this work from the National Natural Science Foundation of China (No. 61166006, 61366002, 61166007 and U1137602) and the Introduction Program of High end scientific and technological talents in Yunnan Province (No. 2013HA019). Also, we wish to acknowledge funding from the Applied Basic Research Programs of Science and the Technology Commission Foundation of Yunnan Province (No. 2010ZC164).

Notes and references

  1. W. S. Yang, J. H. Noh, N. J. Jeon, Y. C. Kim, S. Ryu, J. Seo and S. I. Seok, Science, 2015 DOI:10.1126/science.aaa8765.
  2. J. S. Manser and P. V. Kamat, Nat. Photonics, 2014, 8, 737–743 CrossRef CAS.
  3. S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. Alcocer, T. Leijtens, L. M. Herz, A. Petrozza and H. J. Snaith, Science, 2013, 342, 341–344 CrossRef CAS PubMed.
  4. M. Zhang, H. Yu, M. Lyu, Q. Wang, J.-H. Yun and L. Wang, Chem. Commun., 2014, 50, 11727–11730 RSC.
  5. D. Shi, V. Adinolfi, R. Comin, M. Yuan, E. Alarousu, A. Buin, Y. Chen, S. Hoogland, A. Rothenberger and K. Katsiev, Science, 2015, 347, 519–522 CrossRef CAS PubMed.
  6. Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao and J. Huang, Science, 2015, 347, 967–970 CrossRef CAS PubMed.
  7. H. J. Snaith, A. Abate, J. M. Ball, G. E. Eperon, T. Leijtens, N. K. Noel, S. D. Stranks, J. T.-W. Wang, K. Wojciechowski and W. Zhang, J. Phys. Chem. Lett., 2014, 5, 1511–1515 CrossRef CAS PubMed.
  8. Y. Shao, Z. Xiao, C. Bi, Y. Yuan and J. Huang, Nat. Commun., 2014, 5 DOI:10.1038/ncomms6784.
  9. Y. Yamada, T. Nakamura, M. Endo, A. Wakamiya and Y. Kanemitsu, Appl. Phys. Express, 2014, 7, 032302 CrossRef.
  10. R. E. Wasylishen, O. Knop and J. B. Macdonald, Solid State Commun., 1985, 56, 581–582 CrossRef CAS.
  11. E. Mosconi, A. Amat, M. K. Nazeeruddin, M. Grätzel and F. De Angelis, J. Phys. Chem. C, 2013, 117, 13902–13913 CrossRef CAS.
  12. W. Kong, Z. Ye, Z. Qi, B. Zhang, M. Wang, A. Rahimi-Iman and H. Wu, Phys. Chem. Chem. Phys., 2015, 17, 16405–16411 RSC.
  13. J. Feng and B. Xiao, J. Phys. Chem. Lett., 2014, 5, 1278–1282 CrossRef CAS PubMed.
  14. Y. Kawamura, H. Mashiyama and K. Hasebe, J. Phys. Soc. Jpn., 2002, 71, 1694–1697 CrossRef CAS.
  15. C. C. Stoumpos, C. D. Malliakas and M. G. Kanatzidis, Inorg. Chem., 2013, 52, 9019–9038 CrossRef CAS PubMed.
  16. A. Poglitsch and D. Weber, J. Chem. Phys., 1987, 87, 6373–6378 CrossRef CAS.
  17. J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M. Grätzel, Nature, 2013, 499, 316–319 CrossRef CAS PubMed.
  18. J. H. Heo, D. H. Song, H. J. Han, S. Y. Kim, J. H. Kim, D. Kim, H. W. Shin, T. K. Ahn, C. Wolf and T. W. Lee, Adv. Mater., 2015 DOI:10.1002/adma.201570152.
  19. 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.
  20. M. Liu, M. B. Johnston and H. J. Snaith, Nature, 2013, 501, 395–398 CrossRef CAS PubMed.
  21. P. Pistor, J. Borchert, W. Fränzel, R. Csuk and R. Scheer, J. Phys. Chem. Lett., 2014, 5, 3308–3312 CrossRef CAS PubMed.
  22. J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal and S. I. Seok, Nano letters, 2013, 13, 1764–1769 CrossRef CAS PubMed.
  23. C. Wehrenfennig, G. E. Eperon, M. B. Johnston, H. J. Snaith and L. M. Herz, Adv. Mater., 2014, 26, 1584–1589 CrossRef CAS PubMed.
  24. Q. Chen, H. Zhou, Y. Fang, A. Z. Stieg, T.-B. Song, H.-H. Wang, X. Xu, Y. Liu, S. Lu, J. You, P. Sun, J. McKay, M. S. Goorsky and Y. Yang, Nat. Commun., 2015, 6 DOI:10.1038/ncomms8269.
  25. W. Geng, L. Zhang, Y.-N. Zhang, W.-M. Lau and L.-M. Liu, J. Phys. Chem. C, 2014, 118, 19565–19571 CrossRef CAS.
  26. L. K. Ono, S. Wang, Y. Kato, S. R. Raga and Y. Qi, Energy Environ. Sci., 2014, 7, 3989–3993 Search PubMed.
  27. S. Wang, L. K. Ono, M. R. Leyden, Y. Kato, S. R. Raga, M. V. Lee and Y. Qi, J. Mater. Chem. A, 2015, 3, 14631–14641 RSC.
  28. H. Zhu, Y. Fu, F. Meng, X. Wu, Z. Gong, Q. Ding, M. V. Gustafsson, M. T. Trinh, S. Jin and X. Zhu, Nat. Mater., 2015 DOI:10.1038/nmat4271.
  29. S. Colella, E. Mosconi, P. Fedeli, A. Listorti, F. Gazza, F. Orlandi, P. Ferro, T. Besagni, A. Rizzo and G. Calestani, Chem. Mater., 2013, 25, 4613–4618 CrossRef CAS.
  30. D. Liu and T. L. Kelly, Nat. Photonics, 2014, 8, 133–138 CrossRef CAS.
  31. H. Yu, F. Wang, F. Xie, W. Li, J. Chen and N. Zhao, Adv. Funct. Mater., 2014, 24, 7102–7108 CAS.
  32. W. J. Yin, H. Chen, T. Shi, S. H. Wei and Y. Yan, Advanced Electronic Materials, 2015 DOI:10.1002/aelm.201500044.
  33. Q. Chen, H. Zhou, T.-B. Song, S. Luo, Z. Hong, H.-S. Duan, L. Dou, Y. Liu and Y. Yang, Nano Lett., 2014, 14, 4158–4163 CrossRef CAS PubMed.
  34. M. T. Weller, O. J. Weber, P. F. Henry, A. M. Di Pumpo and T. C. Hansen, Chem. Commun., 2015, 51, 4180–4183 RSC.
  35. W. Nie, H. Tsai, R. Asadpour, J.-C. Blancon, A. J. Neukirch, G. Gupta, J. J. Crochet, M. Chhowalla, S. Tretiak and M. A. Alam, Science, 2015, 347, 522–525 CrossRef CAS PubMed.
  36. Y. Yamada, T. Yamada, L. Q. Phuong, N. Maruyama, H. Nishimura, A. Wakamiya, Y. Murata and Y. Kanemitsu, J. Am. Chem. Soc., 2015, 150814091520002 Search PubMed.
  37. S. M. Vorpahl, S. D. Stranks, H. Nagaoka, G. E. Eperon, M. E. Ziffer, H. J. Snaith and D. S. Ginger, Science, 2015, 348, 683–686 CrossRef PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra16516d
Present address: Institute of Modern Optics, School of Physics, Peking University, Beijing 100871, China.
§ Authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.