Shaped Perovskite Solar Cells with High Power Conversion Efficiency

19 downloads 0 Views 2MB Size Report
Feb 6, 2018 - Longbin Qiu , Sisi He , Jiahua Yang , Jue Deng , and Huisheng Peng ..... [3] G. Zhu , B. Peng , J. Chen , Q. Jing , Z. L. Wang , Nano Energy 2015 ...
www.MaterialsViews.com

Photovoltaics

Fiber-Shaped Perovskite Solar Cells with High Power Conversion Efficiency Longbin Qiu, Sisi He, Jiahua Yang, Jue Deng, and Huisheng Peng* Flexible energy harvesting textiles that can work at ambient conditions are attracting increasing attentions in recent years. They have been converted from solar, thermal, or mechanical energy to power electronic devices.[1–3] Among them, solar energy has been mostly explored as a general and promising strategy in the development of flexible energy harvesting textiles as it is clean, safe, and inexhaustible. As a result, a lot of efforts are made to fabricate fiber-shaped solar cells that can be then woven into energy harvesting textiles.[4–6] These fiber-shaped photovoltaic devices have been mainly made from dye-sensitized solar cells where liquid electrolytes are required to realize relatively high power conversion efficiencies (PCEs),[7] and the use of solid-state electrolytes generally produces much lower PCEs.[8–10] Some attempts are also made to develop fiber-shaped all-solid-state polymer solar cells but with even lower PCEs.[11–13] It remains challenging to fabricate fiber-shaped solar cells at a solid state with acceptable PCEs. Perovskite solar cells appear in a solid state and have also been widely studied for high PCEs.[14,15] However, unlike traditional active materials which are chemically bonded to electrodes[16,17] or possess high film-forming capability,[18] it is challenging to deposit the perovskite layers with full coverage and high quality due to their poor film-forming ability and sensitivity to humidity.[19] The poor coverage of light harvesting layer would cause low-resistance shunting pathways and poor light absorption that lead to both low open-circuit voltage (VOC) and short-circuit current density (JSC).[20] It becomes even worse when the deposition has to be made on a curved surface such as on a microfiber substrate. For instance, several technologies like vacuum deposition,[21] two-step deposition,[22] solid-state film formation,[23] and vapor-assisted deposition have been developed to prepare high-quality perovskite layers on planar substrates,[24] but they are not applicable for the fiber substrates. Currently, a fiber-shaped device may be fabricated through a dip-coating process with low-quality active layers. As a result, low PCEs L. Qiu, S. He, J. Yang, J. Deng, Prof. H. Peng State Key Laboratory of Molecular Engineering of Polymers Department of Macromolecular Science and Laboratory of Advanced Materials Fudan University Shanghai 200438, China E-mail: [email protected] DOI: 10.1002/smll.201600326 small 2016, 12, No. 18, 2419–2424

had been observed from the perovskite solar cell with a curved structure.[25–28] In this Communication, a cathodic deposition solution process has been developed to prepare perovskite layers with high coverage and uniformity on curve surfaces such as titanium (Ti) wires. When the modified Ti wire and transparent aligned carbon nanotube (CNT) sheet are used as two electrodes to produce a coaxial perovskite solar cell fiber (PSCF), a high PCE of 7.1% also with a high open-circuit voltage of 0.85 V has been generated. Figure 1 schematically shows the fabrication and structure of the PSCF. In a typical solar cell, the perovskite layer is sandwiched between electron and hole transport materials.[29] The most investigated electron extraction and transport materials are made from nanoparticles such as TiO2 and ZnO.[10–12,27] After the electron injection, the transport through these nanoparticle boundaries might increase the resistance and slower electron transport. The use of aligned 1D TiO2 materials has been demonstrated to produce highly efficient solar cells.[5,6] Here the TiO2 nanotube array had been incorporated for rapid electron extraction and transport with controlled thickness and uniformity. A Ti wire is used as a substrate (Figure S1, Supporting Information). Through an anodization process, aligned TiO2 nanotube array was radially grown as a mesoporous layer on the surface for electron extraction and transport (Figure 2a,b). There were three stages in the formation of TiO2 nanotubes from a fluorinecontaining solution, i.e., formation of a compact oxide layer, formation of pores by fluorine-ion etching and steady growth of TiO2 nanotubes.[30] No TiO2 nanotubes were observed at the first stage during anodization, and they appeared at the second stage and were perfected at the third stage (Figure S2, Supporting Information). The diameters of TiO2 nanotubes were controlled by varying the used voltage in anodization (Figure S3, Supporting Information). For instance, their diameters had been increased from 45 to 100 nm when the anodization voltages were enhanced from 10 to 20 V. With the further increase to 30 V, no complete nanotubes were produced, and a honeycomb structure had been generated as a result. As larger nanotubes favored the infiltration of active materials to fully cover them both inside and outside, the diameter of 100 nm was mainly investigated below. Another important parameter that may influence the photovoltaic performance of TiO2 nanotubes was related to their lengths that can be controlled by varying the growth time. The lengths were increased with the increasing growth time and reached a platform after ≈15 min (Figure S4, Supporting

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.small-journal.com

2419

communications www.MaterialsViews.com

Figure 1. Schematic illustration to the fabrication and structure of the perovskite solar cell fiber (PSCF).

Information), and they could be controlled at a range of 350–1100 nm in 15 min. Note that a compact TiO2 layer with the thickness of ≈50 nm was formed between the TiO2 nanotube array and Ti wire. This compact layer could prevent a direct contact between the perovskite layer and Ti wire electrode.[31] The direct contact was found to cause a charge recombination with low photovoltaic performance.[29,32] A porous, sponge-structured PbO layer was then coated on the TiO2 nanotubes through a cathodic deposition (Figure S5, Supporting Information). PbO reacted with hydroiodic acid to produce PbI2 that displayed a perfect crystal structure (Figure S6, Supporting Information), and the porous structure was beneficial for the solution infiltration and formation of PbI2 crystal plates. These PbI2 crystals had been infiltrated into the voids among aligned TiO2 nanotubes, and widths of crystal plates were ranged from 500 to 1000 nm. The thickness of the PbI2 layer had been increased by increasing the thickness of the PbO layer. For instance, a thick PbO layer (e.g., 1 µm) produced thick PbI2 crystal plates (thickness of 500 nm) that were perpendicularly aligned due to a steric hindrance (Figure 2c,d); a thin PbO layer (e.g., 100 nm) resulted in the formation of sparsely and randomly distributed PbI2 crystal plates with a thickness of 40 nm (Figure S7, Supporting Information). The PbI2 crystal plate reacted with CH3NH3I to form perovskite CH3NH3PbI3 nanocrystals. Compared with the conventional PbI2 layer formed by spin-coating or dip-coating

2420 www.small-journal.com

processes, the designed porous PbI2 plate array provided more voids for the infiltration of CH3NH3I solution and benefited a complete formation of perovskite. The dense PbI2 layer (Figure S8, Supporting Information) deposited from the previous spin-coating process would prevent the complete reaction with CH3NH3I, i.e., a certain percentage of PbI2 remained without reaction. The evolution of the perovskite layers from two different PbI2 layers based on this electrochemical deposition and spin-coating process was also compared by X-ray diffraction (Figure S9, Supporting Information). The thickness of PbI2 layer is around 500 nm. For the PbI2 crystal, the peaks at 12.6°, 25.5°, 38.5°, and 52.2° correspond to the (006), (0012), (0018), and (0024) planes, respectively. For the perovskite CH3NH3PbI3 material, the peaks at 14.08°, 28.41°, 31.85°, and 43.19° correspond to the (110), (220), (310), and (314) planes, respectively.[33] For the perovskite layer derived from a dense and compact PbI2 layer, the peaks that correspond to PbI2 could be clearly observed at the angle of 12.6° with a similar intensity to 14.08° that corresponded to CH3NH3PbI3. In a strong contrast, less crystalline PbI2 remained in the perovskite layer derived from the aligned plate array.[7,34] It was found that the remaining PbI2 was detrimental to the performance of photovoltaic device due to a slow charge transport.[34,35] X-ray photoelectron spectroscopy was further carried out for the surface composition of this material (Figure S10, Supporting Information). C, N, Pb, and I were mainly detected with trace O and Ti. The ratios among N, Pb, and I are 1/0.90/3.65. Compared with the composition of CH3NH3PbI3, there remained excess CH3NH3I on the surface. The perovskite materials appeared as cubic crystals with an average size of 300 nm (Figure 2e,f), and the coverage and thickness of perovskite layer could be easily tuned with the PbI2 layer. For a thickness of 40 nm, only sparsely distributed CH3NH3PbI3 nanocrystals were observed (Figure S11, Supporting Information). With the thickness of PbI2 layer increase to 500 nm, the perovskite layer fully covered on the TiO2 nanotube array (Figure S12a, Supporting Information). In addition, the voids among the aligned TiO2 nanotubes were also filled with perovskite materials (Figure S12b, Supporting Information). The high coverage and uniformity may increase the light absorption and decrease the charge recombination due to the direct contact between the CNT sheet and electron extraction materials. An aligned CNT sheet was dry-drawn from a spinnable array and then wound around the perovskite layer. The asprepared CNT sheets were a kind of aerogels that were coated around the perovskite layer without a close contact (Figure S13, Supporting Information). After an isopropanol treatment, the CNT sheet aerogel was condensed into a thin film (Figure 2g,h). The CNT sheet was transparent with a transmittance over 80% in the spectrum range of 400–800 nm (Figure S14, Supporting Information) and electrically conductive with a conductivity of 500 S cm−1, so it may serve as an effective electrode.[36,37] With a large specific surface area that provided high charge collection and transport, CNT sheet also functioned as an effective hole extraction layer.[36–38] Stable photoluminescence measurement was made to investigate the hole extraction performance of CNT sheet (Figure 3a). Due to its efficient hole extraction,

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

small 2016, 12, No. 18, 2419–2424

www.MaterialsViews.com

Figure 2. Scanning electronic microscopy images for the structure of the PSCF during fabrication. a,b) TiO2 nanotube array by top and side views, respectively. c,d) PbI2 nanoplate array by top and side views, respectively. e,f) Perovskite nanocrystal layer by top and side views, respectively. g) Aligned CNT sheet. h) A completed PSCF containing TiO2, CH3NH3PbI3, and CNT by a side view.

the intensity of fluorescence decreased significantly, so no hole transport layer was required to simplify the fabrication process and device structure aiming at high performances.[38] The wound CNTs remained highly aligned and intensely adhered to the perovskite layer, which was important for both high PCEs and high stability under deformation. By replacing the organic hole transport material with aligned CNTs, the stability of solar cells could be improved.[39] The combination of aligned TiO2 nanotube array, highcoverage CH3NH3PbI3 layer, and aligned CNT sheet electrode contributed to a high performance in PSCF. For instance, small 2016, 12, No. 18, 2419–2424

the TiO2 was critical to the injection and transport of photogenerated charge carriers.[5] To investigate the influence of TiO2 layer in the photovoltaic performance, a high-coverage CH3NH3PbI3 layer with a thickness around 900 nm was first studied. When the anodization was stopped at an early stage with the formation of a porous TiO2 layer rather than aligned nanotubes (Figure 3b), the resulting PSCF showed a poor performance. After the formation of aligned TiO2 nanotubes, higher JSC values were observed as the aligned nanotubes enhanced the charge transport in the radial direction. When the length of the TiO2 nanotubes was around 400 nm, the best

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.small-journal.com

2421

communications www.MaterialsViews.com

Figure 3. a) Steady-state photoluminescence spectra of CH3NH3PbI3 layer and CH3NH3PbI3/CNT sheet layer. b) Voltage–current density curves of PSCFs without and with TiO2 nanotube arrays at increasing thicknesses. c) Voltage–current density curves of PSCFs with increasing perovskite CH3NH3PbI3 thicknesses. d) Voltage–current density curve of the PSCF fabricated from CNT sheet/silver composite electrode with the silver thickness of 5 nm.

photovoltaic performance was achieved, i.e., VOC of 0.852 V, JSC of 8.9 mA cm−2, fill factor (FF) of 0.482, and PCE of 3.6%. VOC, JSC, and FF were decreased simultaneously with the increasing length of TiO2 nanotube. The increasing thickness of electron transport layer would increase the resistance and charge recombination with a lower performance.[29] This is consistent to the previous report that TiO2 layer with length of hundreds of nanometers were more effective as shorter ones could not incorporate enough active materials while longer ones increased the transport pathways of charges with lower efficiencies.[30] Therefore, the length of about 400 nm had been mainly used for the fabrication of the PSCF. The capping perovskite layer was also a key factor for the photovoltaic performance of the PSCF.[30] The optimization of PSCF was investigated with a fixed thickness of the TiO2 nanotube array at ≈400 nm. For a sparse PbI2 layer, the coverage of perovskite layer was low (Figure S11, Supporting Information). The electron transport material was directly contacted with the aligned CNT sheet electrode, producing a shunt pathway that significantly lowered the performance of the PSCF, e.g., VOC of 0.45 V, JSC of 2.66 mA cm−2, and PCE of 0.70% (Figure 3c). When the perovskite layer was fully infiltrated among and covered on the aligned TiO2 nanotube with a thickness of ≈350 nm, a high VOC of ≈0.85 V and JSC of 16.1 mA cm−2 were obtained, and the PCE was calculated as 6.8% that exceeds the current solid-state fiber-shaped solar cells. Note that the VOC was significantly higher than the previous fiber-shaped perovskite solar cells and similar to its planar counterparts.[25–28,38] Compared with the previous dip-coating process in the fabrication of fibrous solar cells,

2422 www.small-journal.com

the power conversion efficiency was increased significantly. In particular, the open-circuit voltage had been largely enhanced (Figure S15, Supporting Information). The high coverage and uniformity of the perovskite layer contributed to the high VOC due to less shunt pathways between TiO2 nanotube array and CNT sheet thus with a lower recombination. The incident photon conversion efficiency spectra also confirmed the wide range light absorption of the PSCF from 400 to 800 nm (Figure S16, Supporting Information). With a further increase in the CH3NH3PbI3 thickness, the photovoltaic performance was decreased with a decreasing JSC as more crystalline PbI2 remained at the bottom part connected to TiO2. This remaining precursor would also decrease the performance of solar cell. Furthermore, a thicker perovskite formed a stacking layered structure with more grain boundaries in the perpendicular direction (Figure S17, Supporting Information). The increased grain boundaries would also decrease the charge transport. Here a deposition of PbI2 array at a thickness around 500 nm was sufficient for a full coverage and high-quality perovskite layer with a thickness of 350 nm. Figure S18 (Supporting Information) has further compared the J–V curves of a PSCF under forward and reverse scan directions, and it showed a hysteresis with a slightly decreased fill factor. The PSCF has been compared with the other solid-state fiber-shaped solar cells in Figure 4. The PCE are much higher than the other photovoltaic devices. Conventional solid state solar cell fiber applied polymer solar cell or dye sensitized solar cell structure, which would have lost a great part energy in interface charge transport. A moderate band gap

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

small 2016, 12, No. 18, 2419–2424

www.MaterialsViews.com

Figure 4. Comparison on the power conversion efficiency between the PSCF and the other solid-state fiber-shaped solar cells.

of CH3NH3PbI3 was promising for both high light absorption and high voltage.[40] With a small binding energy, the energy loss in charge generation and transport could be minimized, which could not be omitted in dye-sensitized and polymer solar cells.[41] As previously mentioned, the high coverage and uniformity of the CH3NH3PbI3 layer was another key for the higher performance than its counterparts. However, the fill factor of PSCF was relatively lower, which decreased the photovoltaic performance. Electrochemical impedance spectra were further studied for the PSCF (Figure S19a, Supporting Information), and the low fill factor was found to be derived from the high series resistance of the CNT sheet electrode (Figure S19b, Supporting Information). Silver is introduced to increase the conductivity of CNT sheet back contact electrode through a thermal deposition. The conductivity of this composite back contact electrode was increased from

≈500 to ≈1000 S cm−1 (Figure S20, Supporting Information). However, with the increasing silver thickness, the transmittance was decreased (Figure S21, Supporting Information). The optimal power conversion efficiency of 7.1% occurred at the silver thickness of 5 nm with much enhanced fill factor of 0.56 while slightly decreased short-circuit current density of 14.5 mA cm−2. The photovoltaic performance may be further largely enhanced by developing highly conductive and transparent electrode materials. Based on the solution process, the fabrication can be easily scaled up for a continuous production (Figure S22, Supporting Information). The PSCF was flexible and can be deformed into various structures (Figure 5a,b). The photovoltaic parameters are further compared in Figure 5c. JSC was slightly decreased while FF had been increased due to a better contact between perovskite layer and CNT sheet, and the PCE could be maintained by above 90% after twisting for 400 cycles. Interestingly, the CNT sheet adhered to the perovskite crystal more intensely after twisting, which protected the perovskite layer from peeling off (Figure S23, Supporting Information) In summary, high-performance perovskite solar cells in a fiber shape have been developed through an electrochemical deposition process. They display VOC of 0.85 V and PCE of 7.1%, both of which, to the best of our knowledge, represent the highest values in the all-solid-state fiber-shaped solar cells. However, for the PSCFs, it is important to further enhance their producibility aiming at practical applications (Figure S24, Supporting Information). Due to a coaxial configuration, these PSCFs are flexible and stable, and can be further woven into smart clothes to power the next-generation portable and wearable electronic devices.

Experimental Section Preparation of TiO2 Nanotube Array: Titanium wire (250 µm) was sequentially washed with acetone, isopropanol, and water in ultrasonic bath, followed by drying in air. The anodization of the titanium

Figure 5. Flexible characterization of the PSCF. a,b) Photographs of a PSCF being tied and twisted, respectively. c) Dependence of photovoltaic parameters on twisting cycle number. small 2016, 12, No. 18, 2419–2424

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.small-journal.com

2423

communications www.MaterialsViews.com

wire was proceeded in an electrolyte containing 0.27 M NH4F in a mixture solvent of water and glycerol (volume ratio, 1/1). It was operated in a two-electrode system with the titanium wire anode and platinum plate cathode at an anodized voltage of 10–30 V. After anodizing and washing by water, the resulting titanium wire was annealed at 500 °C for 60 min. The modified titanium wire was immersed into a TiCl4 aqueous solution (concentration of 30 × 10−3 M) for 30 min at 70 °C and annealed at 450 °C for 30 min to deposit a thin layer of TiO2 nanoparticles on the outer surface. Synthesis of Perovskite Layer: A cathodic deposition of PbO was also made with an oxidic titanium wire as the cathode and a platinum plate as the anode at an electrolyte containing 2 × 10−3 M Pb(NO3)2 and 0.2 M H2O2 at a constant current density of 5 mA cm−2. The resulting titanium wire was annealed at 200 °C for 30 min to complete the transformation to a PbO layer. The PbO was dipped in a hydroiodic acid aqueous solution (concentration of 0.1 M) and dried at 80 °C for 60 min to eliminate the trace water. After cooled down to room temperature, this PbI2-coated wire was dipped into a CH3NH3I/isopropanol solution (concentration of 10 mg mL−1) for 60 s and annealed at 80 °C for 30 min to obtain the light-harvesting perovskite layer. Fabrication of PSCF: After deposition of the TiO2 nanotube array and perovskite layer, the modified Ti wire was fixed to two synchronous motors at the two ends. A transparent CNT sheet was drawn out of a spinnable array and attached to the modified Ti wire. With the rotation of the two synchronous motors, this CNT sheet was wrapped around the perovskite layer to produce a PSCF.[27] For a close contact between CNT sheet and perovskite layer, the completed PSCF was dipped into isopropanol and dried at 80 °C for 30 min. According to the generally accepted method,[10–12] the effective area was calculated by multiplying the diameter of the modified Ti wire and the length of the PSCF, i.e., the projection area of the PSCF.

Supporting Information Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements This work was supported by the National Natural Science Foundation of China grants 21225417, 51573027, and 51403038, Science and Technology Commission of Shanghai Municipality grants 15XD1500400, 12nm0503200, and 15JC1490200, and the Program for Outstanding Young Scholars from Organization Department of the Communist Party of China Central Committee.

[1] M. Kaltenbrunner, G. Adam, E. D. Głowacki, M. Drack, R. Schwödiauer, L. Leonat, D. H. Apaydin, H. Groiss, M. C. Scharber, M. S. White, N. S. Sariciftci, S. Bauer, Nat. Mater. 2015, 14, 1032. [2] Y. Chen, Y. Zhao, Z. Liang, Energy Environ. Sci. 2015, 8, 401. [3] G. Zhu, B. Peng, J. Chen, Q. Jing, Z. L. Wang, Nano Energy 2015, 14, 126. [4] T. Chen, L. Qiu, Z. Yang, H. Peng, Chem. Soc. Rev. 2013, 42, 5031. [5] T. Chen, L. Qiu, H. G. Kia, Z. Yang, H. Peng, Adv. Mater. 2012, 24, 4623.

2424 www.small-journal.com

[6] T. Chen, L. Qiu, Z. Yang, Z. Cai, J. Ren, H. Li, H. Lin, X. Sun, H. Peng, Angew. Chem., Int. Ed. 2012, 51, 11977. [7] Z. Yang, H. Sun, T. Chen, L. Qiu, Y. Luo, H. Peng, Angew. Chem., Int. Ed. 2013, 52, 7545. [8] H. Li, J. Guo, H. Sun, X. Fang, D. Wang, H. Peng, ChemNanoMat 2015, 1, 399. [9] H. Li, Z. Yang, L. Qiu, X. Fang, H. Sun, P. Chen, S. Pan, H. Peng, J. Mater. Chem. A 2014, 2, 3841. [10] D. Wang, S. Hou, H. Wu, C. Zhang, Z. Chu, D. Zou, J. Mater. Chem. 2011, 21, 6383. [11] M. R. Lee, R. D. Eckert, K. Forberich, G. Dennler, C. J. Brabec, R. A. Gaudiana, Science 2009, 324, 232. [12] D. Liu, M. Zhao, Y. Li, Z. Bian, L. Zhang, Y. Shang, X. Xia, S. Zhang, D. Yun, Z. Liu, A. Cao, C. Huang, ACS Nano 2012, 6, 11027. [13] Z. Zhang, Z. Yang, Z. Wu, G. Guan, S. Pan, Y. Zhang, H. Li, J. Deng, B. Sun, H. Peng, Adv. Energy Mater. 2014, 4, 1301750. [14] A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem. Soc. 2009, 131, 6050. [15] W. S. Yang, J. H. Noh, N. J. Jeon, Y. C. Kim, S. Ryu, J. Seo, S. I. Seok, Science 2015, 348, 1234. [16] A. Yella, C. L. Mai, S. M. Zakeeruddin, S. N. Chang, C. H. Hsieh, C. Y. Yeh, M. Gratzel, Angew. Chem., Int. Ed. 2014, 53, 2973. [17] S. Mathew, A. Yella, P. Gao, R. Humphry-Baker, B. F. Curchod, N. Ashari-Astani, I. Tavernelli, U. Rothlisberger, M. K. Nazeeruddin, M. Gratzel, Nat. Chem. 2014, 6, 242. [18] Y. Diao, Y. Zhou, T. Kurosawa, L. Shaw, C. Wang, S. Park, Y. Guo, J. A. Reinspach, K. Gu, X. Gu, B. C. Tee, C. Pang, H. Yan, D. Zhao, M. F. Toney, S. C. Mannsfeld, Z. Bao, Nat. Commun. 2015, 6, 7955. [19] A. Dualeh, N. Tétreault, T. Moehl, P. Gao, M. K. Nazeeruddin, M. Grätzel, Adv. Funct. Mater. 2014, 24, 3250. [20] G. E. Eperon, V. M. Burlakov, P. Docampo, A. Goriely, H. J. Snaith, Adv. Funct. Mater. 2014, 24, 151. [21] M. Liu, M. B. Johnston, H. J. Snaith, Nature 2013, 501, 395. [22] J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin, M. Gratzel, Nature 2013, 499, 316. [23] Y. Y. Zhou, M. J. Yang, A. L. Vasiliev, H. F. Garces, Y. X. Zhao, D. Wang, S. P. Pang, K. Zhu, N. P. Padture, J. Mater. Chem. A 2015, 3, 9249. [24] Q. Chen, H. Zhou, Z. Hong, S. Luo, H. S. Duan, H. H. Wang, Y. Liu, G. Li, Y. Yang, J. Am. Chem. Soc. 2014, 136, 622. [25] R. Li, X. Xiang, X. Tong, J. Zou, Q. Li, Adv. Mater. 2015, 27, 3831. [26] M. Lee, Y. Ko, Y. Jun, J. Mater. Chem. A 2015. 3, 19310. [27] L. Qiu, J. Deng, X. Lu, Z. Yang, H. Peng, Angew. Chem., Int. Ed. 2014, 53, 10425. [28] S. He, L. Qiu, X. Fang, G. Guan, P. Chen, Z. Zhang, H. Peng, J. Mater. Chem. A 2015, 3, 9406. [29] Y. Zhao, K. Zhu, J. Phys. Chem. Lett. 2014, 5, 4175. [30] P. Roy, S. Berger, P. Schmuki, Angew. Chem., Int. Ed. 2011, 50, 2904. [31] Y. Wu, X. Yang, H. Chen, K. Zhang, C. Qin, J. Liu, W. Peng, A. Islam, E. Bi, F. Ye, M. Yin, P. Zhang, L. Han, Appl. Phys. Express 2014, 7, 052301. [32] T. Chen, L. Qiu, H. Li, H. Peng, J. Mater. Chem. 2012, 22, 23655. [33] C. Ying, C. Shi, N. Wu, J. Zhang, M. Wang, Nanoscale 2015, 7, 12092. [34] H.-S. Ko, J.-W. Lee, N.-G. Park, J. Mater. Chem. A 2015, 3, 8808. [35] L. Wang, C. McCleese, A. Kovalsky, Y. Zhao, C. Burda, J. Am. Chem. Soc. 2014, 136, 12205. [36] H. Sun, H. Li, X. You, Z. Yang, J. Deng, L. Qiu, H. Peng, J. Mater. Chem. A 2014, 2, 345. [37] L. Qiu, X. Sun, Z. Yang, W. Guo, H. Peng, Acta Chim. Sin. 2012, 70, 1523. [38] Z. Li, S. A. Kulkarni, P. P. Boix, E. Shi, A. Cao, K. Fu, S. K. Batabyal, J. Zhang, Q. Xiong, L. H. Wong, N. Mathews, S. G. Mhaisalkar, ACS Nano 2014, 8, 6797. [39] S. N. Habisreutinger, T. Leijtens, G. E. Eperon, S. D. Stranks, R. J. Nicholas, H. J. Snaith, Nano Lett. 2014, 14, 5561. [40] H. J. Snaith, J. Phys. Chem. Lett. 2013, 4, 3623. [41] S. Collavini, S. F. Volker, J. L. Delgado, Angew. Chem., Int. Ed. 2015, 54, 9757. Received: January 31, 2016 Revised: February 29, 2016 Published online: March 22, 2016

© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

small 2016, 12, No. 18, 2419–2424