Transition from the Tetragonal to Cubic Phase of ... - ACS Publications

5 downloads 87258 Views 2MB Size Report
Sep 29, 2015 - around 20% in 2015.12 In order to further improve the performance of ... PSCs,26 and good understanding including theoretical calcu- lations on the ..... the Australian National Fabrication Facility, a company established ... 2012, 2, 591. (12) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.;.
Letter pubs.acs.org/JPCL

Transition from the Tetragonal to Cubic Phase of Organohalide Perovskite: The Role of Chlorine in Crystal Formation of CH3NH3PbI3 on TiO2 Substrates Qiong Wang, Miaoqiang Lyu, Meng Zhang, Jung-Ho Yun, Hongjun Chen, and Lianzhou Wang* Nanomaterials Centre, School of Chemical Engineering and Australian Institute for Bioengineering and Nanotechnology, The University of Queensland, Brisbane, Queensland 4072, Australia S Supporting Information *

ABSTRACT: The role of chlorine in the superior electronic property and photovoltaic performance of CH3NH3PbI(3−x)Clx perovskite has attracted recent research attention. Here, we study the impact of chlorine in the perspective of the crystal structure of the perovskite layer, which can provide important understanding of its excellent charge mobility and extended lifetimes. In particular, we find that in the presence of chlorine (PbCl2 or CH3NH3Cl), when CH3NH3PbI3 films are deposited on a TiO2 mesoporous layer instead of a planar TiO2 substrate, a stable cubic phase rather than the commonly observed tetragonal phase is formed in CH3NH3PbI3 perovskite at room temperature. The relative peak intensity of two major facets of cubic CH3NH3PbI3 crystals, (100)C and (200)C facets, can also be easily tuned, depending on the film thickness. Furthermore, compared with pristine CH3NH3PbI3 perovskite films, in the presence of chlorine, CH3NH3PbI3 crystals grown on planar substrates exhibit strong preferred orientations on (110)T and (220)T facets.

O

can be obtained,4 which implies that an electron-selective layer is not necessitated for CH3NH3PbI3−xClx perovskite.23 Although single-crystal CH3NH3PbI3 perovskite of large grain sizes prepared by advanced techniques can also exhibit very long charge diffusion lengths ranging from a few micrometers to 100 μm,24,25 it is still of high importance to study the role of chlorine on the crystal formation of a perovskite film prepared from the simple one-step spin-coating method.26,27 In particular, it was recently reported that perovskite films prepared from precursor solutions with and without chlorine share the same crystal structure and are actually the same material as chlorine would be evaporated as CH3NH3Cl during the annealing process.28 Hence, it is suggested that the expression of CH3NH3PbI3−xClx is indeed not accurate because techniques such as X-ray photoelectron spectroscopy (XPS) or energy dispersive spectroscopy (EDS) have been used to demonstrate zero or a trace amount of chlorine on the surface of the final perovskite layer.29,30 However, Starr et al.,31 demonstrated an inhomogeneous chlorine distribution in CH3NH3PbI3−xClx layers as they observed a higher average concentration of chlorine throughout the perovskite layer than on the surface. In their work, they also discussed the possible electronic consequences of the physical location of chlorine in the perovskite layer, which contributed to superior charge mobility and lifetimes compared with the chlorine-free CH3NH3PbI3 perovskite layer. Colella et al.27 conducted

rganohalide lead perovskite solar cells (PSCs) have seen significant progress in boosting the power conversion efficiency (PCE) in the past 5 years,1,2 which is primarily attributed to the outstanding photophysical properties of organic−inorganic lead perovskite, including high absorption coefficient,3 extraordinary long charge diffusion lengths/lifetimes,4,5 high carrier mobilities,6 low exciton binding energy,7,8 and strong photoluminescence efficiency.9,10 In addition, the solution processability makes PSCs one of the most promising new-generation solar cells that could compete with siliconbased solar cells, with the PCE surging from 9% in 201211 to around 20% in 2015.12 In order to further improve the performance of PSCs, extensive work has been done to investigate a number of challenging issues including the working mechanism and stability of PSCs,13,14 the cause of hysteresis in PSCs,15,16 and the influence of preparation techniques on film morphology and crystal formation of the perovskite layer.17,18 In particular, the influence of chlorine in the precursor solution of perovskites has drawn recent attention due to its significant impact on the morphology and electronic property of the perovskite layer and thus on the photovoltaic performance of PSCs.19−21 For polycrystalline perovskite, CH3NH3PbI3, prepared from the spin-coating or dip-coating method, the electron diffusion length is on the order of or slightly shorter than the absorption depth (∼100 nm),5 which indicates that an efficient mesoporous electron-transport layer is generally required.22 However, for polycrystalline perovskite, CH3NH3PbI3−xClx deposited from chlorine-containing solutions using the one-step spin-coating method, surprisingly longer rang electron and hole diffusion lengths of around 1 μm © 2015 American Chemical Society

Received: August 4, 2015 Accepted: September 29, 2015 Published: September 29, 2015 4379

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384

Letter

The Journal of Physical Chemistry Letters

Figure 1. XRD patterns of CH3NH3PbI3 films deposited from precursors S1, S2, and S3 on planar TiO2 substrates (TiO2−PS) (a) and on TiO2 mesoporous substrates (TiO2−MS, ∼500 nm) (b). (c) CH3NH3PbI3 films deposited from the S2 precursor on planar and mesoporous TiO2 substrates with different film thicknesses: TiO2−MS-1 for 300 nm; TiO2−MS-2 for 500 nm. (Thicker TiO2 mesoporous substrates were prepared by depositing twice the TiO2 solution via the spin-coating method; see the Supporting Information. XRD measurements of the anatase TiO2 compact layer and S1 and S3 on TiO2 substrates of different film thicknesses are given in S-Figures 3 and 4, respectively.)

CH3NH3Cl in S1 under the same molar ratio of CH3NH3I and PbI2. Sample S3 was prepared by dissolving CH3NH3I and PbCl2 at a molar ratio of 3:1 with a total mass weight of 40 wt % in DMF solvent. Detailed experimental conditions are given in the Supporting Information. The transition between tetragonal and cubic crystal phases has been observed in bromide-substituted lead triiodide perovskite.37 It is reported that as the concentration of bromide in CH3NH3PbI3−xBrx increases, (004) and (220) facets of the tetragonal I4/mcm phase located in the 2θ range of 28−29° are gradually merged into one single peak corresponding to the (200) facet in the cubic Pm3m ̅ phase, indicating increased symmetry. As a result, X-ray diffraction (XRD) patterns of CH3NH3PbI3 perovskite films prepared from S1, S2, and S3 precursors deposited on a planar TiO2 compact layer were measured, and the results are given in Figure 1a. It is clear that all samples have the (004) facet that sits close to the (220) facet, indicating the tetragonal phase of the perovskite crystal.38 Meanwhile, it is also noted that the relative peak intensity between the (004)T facet and (220)T facet is quite different in the three samples. Compared with pristine CH3NH3PbI3 perovskite prepared from the S1 precursor, CH3NH3PbI3 perovskite prepared from the CH3NH3Cl-containing precursor (S2) presents the (004)T facet as the preferential facet. However, the one prepared from the PbCl2-containing precursor (S3) exhibits the preferred (220)T facet. Interestingly, when the perovskite film was deposited on the TiO 2 mesoporous layer with a thickness of around 500 nm, the (004)T facet disappeared and merged into the (200)C facet for the CH3NH3PbI3 perovskite deposited from S2 and S3 precursors, Figure 1b. This implies that CH3NH3PbI3 crystals have transited from the tetragonal phase to the cubic phase.39,40 Calculation of the XRD pattern of S3 in Figure 1b gives the lattice parameter of around 6.265 Å, which matches well with the computer modulation data of the cubic CH3NH3PbI3.41 The influence of the TiO2 mesoporous layer thickness on the crystal structure of CH3NH3PbI3 perovskite is presented in Figure 1c. It shows that as the film thickness of the TiO2 mesoporous layer increases from 0 (planar substrate) to around 300 nm, the peak intensity of the (004)T facet is greatly decreased while the (220)T facet is kept intact. Further increase in the film thickness of photoanodes leads to the disappearance of the (004)T facet and prominent enhancement in the (200)C facet. Eventually, as the film thickness of the TiO2 mesoporous layer increases, the perovskite layer deposited from the

combined angle-resolved X-ray photoelectron spectroscopy (AR-XPS) and first-principles DFT modeling in investigation of the CH3NH3PbI3−xClx/TiO2 interface and revealed that chloride was preferentially located in close proximity to the perovskite/TiO2 interface due to an increased chloride−TiO2 surface affinity. In light of the arguments on the role of chlorine in determining some important properties including charge mobility and lifetimes of halide perovskite films,32,33 it is of importance to further investigate the influence of chlorine on the crystal structure of perovskite materials, which is expected to provide better understanding of the structure−property relation of the material system. In this study, we focus on a TiO2 substrate because TiO2 has been widely used as an electron-selective layer in conventional PSCs,26 and good understanding including theoretical calculations on the interaction between halide and TiO2 have been reported.34,35 For instance, Mosconi et al.36 used first-principles electronic structure calculations, suggesting that the presence of chlorine improved the preferred orientation for a perovskite film grown on TiO2 due to enhanced binding energy between the substrate and perovskite, thereby leading to enhanced stability. Here, we adopt two chloride precursors, PbCl2 and CH3NH3Cl, and three substrates, a planar TiO2 compact layer and TiO 2 mesoporous layers with two different film thicknesses, to study the role of chlorine in the crystal structure of CH3NH3PbI3 perovskite. By comparing the crystal structure of CH3NH3PbI3 perovskite deposited on a planar substrate and TiO2 mesoporous layer, we find that the resulting perovskite experiences transition from the tetragonal to cubic phase as the film thickness of the TiO2 mesoporous layer increases. Under the same mesoporous layer, the relative peak intensity of two major facets of cubic CH3NH3PbI3 crystals, (100)C and (200)C facets, can be tuned. Compared with the pristine CH3NH3PbI3 perovskite film, the chlorine-involved CH3NH3PbI3 perovskite film exhibits significantly enhanced peak intensity of the (110)T and (220)T facets for TiO2 planar substrates. The finding here provides new fundamental understanding of the crystal structural control of the organic−inorganic hybrid perovskite that could be associated with desirable optoelectronic properties. Pristine CH3NH3PbI3 perovskite films were prepared by spin-coating a solution containing an equal molar ratio of CH3NH3I and PbI2 at a concentration of 40 wt % in N,Ndimethylformamide (DMF) solvent onto certain substrates. Samples prepared under this condition are referred as S1. Then, sample S2 was prepared by dissolving an additional amount of 4380

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384

Letter

The Journal of Physical Chemistry Letters

perovskite formation process and thus promotes the growth of crystal domains during annealing due to an initial introduction of a CH3NH3+ rich environment,28 while the latter changes the nucleation dynamic and alters the growth kinetic and crystallite sizes upon chloride inclusion.21 Although the existence of chlorine in the final perovskite film is still under debate (our XPS measurement of perovskite prepared from chlorine-containing precursors given in the Supporting Information shows no existence of chlorine in the final product, S-Figure 2), 31,43 our study shows that CH3NH3PbI3 perovskite prepared from chlorine-containing precursors experiences transition from the tetragonal to cubic phase when it is deposited on a thick TiO2 mesoporous layer. It has been reported that the cubic phase of CH3NH3PbI3 perovskite transits to the tetragonal phase at around 54 °C, and CH3NH3PbI3 perovskite should present a tetragonal structure at room temperature.44,45 Therefore, it is very interesting to observe the formation of cubic CH3NH3PbI3 perovskite as the final product in our study. More explanation on this finding will be given in the following section. Figure 3 presents the XRD patterns of CH3NH3PbI3 films deposited from precursors S1, S2, and S3 on ∼300 nm

CH3NH3Cl-containing precursor (S2) gradually transforms from the tetragonal phase to the cubic phase. It is known that when a planar substrate or thin TiO2 mesoporous layer is used as a photoanode, CH3NH3PbI3 perovskite tends to form a capping layer, and when the photoanode is composed of a thick TiO2 mesoporous layer, it forms a CH3NH3PbI3 perovskite sensitized structure, as illustrated in Scheme 142 (see the Supporting Information for Scheme 1. Illustration of CH3NH3PbI3 Perovskite Deposited from the S2 Precursor on Different Substratesa

a

Substrates: (a) TiO2 planar and mesoporous layers of around (b) 300 and (c) 500 nm thickness.

cross-sectional images, S-Figure 1). On planar substrates, it is noted that two chlorine resources, CH3NH3Cl and PbCl2, can affect the (004)T and (220)T facets of CH3NH3PbI3 perovskite differently, as judged from XRD data in Figure 1. SEM images of CH3NH3PbI3 perovskite films prepared from S1, S2, and S3 precursors on planar substrates are given in Figure 2. It can be

Figure 3. XRD patterns of CH3NH3PbI3 films deposited from precursors S1, S2, and S3 on ∼300 nm mesoporous TiO2 substrates. (Note: the y scales of the peak intensity of three samples are different.)

mesoporous TiO2 substrates. (Absorbance spectra of these three samples are given in the Supporting Information, S-Figure 5). It shows that the sample deposited from S1 exhibits weak peaks, as indexed according to ref 41. In this sample, the peak intensities of the (110)T and (220)T facets are as weak as that of other facets of CH3NH3PbI3 crystals. However, these two facets are greatly enhanced in samples prepared from S2 source. The clear split of the two facets (004) and (220) at 2θ ranging from 28 to 29° signifies the tetragonal phase of CH3NH3PbI3 perovskite. The sample spin-coated from S3 precursor exhibits one strong peak at 2θ of around 28.5°, corresponding to the (200) facet of the cubic CH3NH3PbI3 perovskite. Therefore, compared with pristine CH3NH3PbI3 perovskite, the crystallinity and/or the quantity of the two facets at 2θ in range of 14−15 and 28−29° that are exposed at the surface are significantly improved for CH3NH3PbI3 perovskite deposited from chlorine-containing precursors, while other facets are barely influenced. It has been theoretically calculated that both CH3NH3PbI3 and CH3NH3PbI(3−x)Clx would exhibit preferred orientation on the (110) facet due to the binding of perovskite halide atoms to undercoordinated Ti(IV) atoms of the anatase TiO2.36

Figure 2. Scanning electron microscopy (SEM) images of the surface morphology of CH3NH3PbI3 perovskite films deposited from precursorsS1 (a1, a2), S2 (b1, b2), and S3 (c1, c2) on planar substrates.

seen that jigsaw-like crystals with grain edges are formed for samples prepared from S2. However, bulk crystals with clearly ordered plates on the surface are characterized for samples prepared from S3. Compared with pristine CH3NH3PbI3 perovskite, it implies that in the presence of CH3NH3Cl and PbCl2, the nucleation and growth of CH3NH3PbI3 crystals are modified. Recent reports suggest that these two chlorine resources have different impacts on the nucleation process and crystal growth process, where the former slows down the 4381

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384

Letter

The Journal of Physical Chemistry Letters However, in our experiment, we find that chlorine-free CH3NH3PbI3 perovskite exhibits no any clear preference of the (110) facet. As shown in Figure 3, (202) and (114) facets exhibit comparable peak intensity to (110) and (220) facets. On the other hand, on the basis of their work, the binding energy of Cl−Ti (−27.3 eV) is slightly increased compared with the of the I−Ti bond (−24.2 eV), which cannot be the dominant reason for the significantly enhanced preferential (110) and (220) facets in chlorine-involved perovskites. Meanwhile, they calculated that the (110) facet of the tetragonal perovskite showed a higher (∼7 eV) binding energy to anatase TiO2 surface than that of the (001) facet of the pseudocubic phase.36 We also used anatase TiO2 as the substrate (S-Figure 3), but as observed in Figure 1, perovskite films prepared from chlorine precursors (S2 and S3) exhibit the preferred cubic phase to the tetragonal phase for mesoporous substrates. As a result, it is difficult to explain these differences using their theory, or this is because the interfacial chlorine atoms at the interface between the perovskite and TiO2 are not substantial in our situation.36,46 Recently, several works have been published, trying to reveal the fundamental understanding of the crystal growth of perovskite from the perspective of the path that a system traverses. Up to now, it has been widely accepted that in an unoptimized one-step solution deposition, the formation of intermediates seems to kinetically dominate the initial nucleation, causing poor crystallinity and short-rang order.21 Jeon et al.47 found that a mixed solvent (γ-butyrolactone and dimethyl sulfoxide, DMSO) could be used to tune the growth kinetics of perovskite by creating a new intermediate phase, CH3NH3I−PbI2−DMSO, which ultimately produced a much more crystalline CH3NH3PbI3 film. Alternatively, chloride addition opens up another route to circumvent perovskite nucleation from an amorphous phase through creating the kinetically accessible and structurally coherent PbCl2 and CH3NH3PbCl3 intermediates.48 Tidhar et al.49 found that the presence of PbCl2 nanocrystals could act as heterogeneous nucleation sites for the formation of perovskite crystals in solution. Colella et al.19 suggested that the formation of the CH3NH3PbCl3 intermediate phase largely guided perovskite nucleation and growth in chloride-containing systems. More encouragingly, Pistor et al. monitored the evolution of different crystalline phases during thin film growth by co-evaporating CH3NH3I and PbCl2 using in situ XRD and found that all XRD peaks of films deposited under low PbCl2 fluxes and measured at room temperature could be indexed to cubic CH3NH3PbI3.50 They speculated that the incorporation of chlorine into CH3NH3PbI3−xClx perovskite could effectively lower the cubic−tetragonal transition temperature and stabilize the cubic modification at room temperature. However, more detailed data on the CH3NH3PbI3−CH3NH3PbCl3 system and its phases are desirable to elucidate this point.51 It is worth noting that the relative peak intensity between (100)C and (200)C facets can be tuned. Precursor S4 containing CH3NH3I and PbCl2 at a molar ratio of 3:1 and a mass weight concentration of 60 wt % is used to deposit CH3NH3PbI3 perovskite films on photoanodes composed of a TiO 2 mesoporous layer of around 300 nm. XRD patterns of CH3NH3PbI3 perovskite films spin-coated from S3 and S4 precursors on conductive glass and mesoporous TiO 2 substrates are given in Figure 4. The peak intensity ratios between two major facets of CH3NH3PbI3 perovskite prepared under the above conditions are summarized in Table 1. It can

Figure 4. XRD patterns of CH3NH3PbI3 films deposited from S3 and S4 precursors on top of the nonconductive side of FTO and ∼300 nm mesoporous TiO2 substrates.

Table 1. Relative Peak Intensity of (100)C and (200)C Facets of CH3NH3PbI3 Crystals Given in Figure 4 samples peak intensity ratio of the (100)C to (200)C facet

S3− glass

S3−TiO2 MS

S4−TiO2 MS

1.05

0.95

0.77

be seen that under the same concentration of CH3NH3PbI3 precursor (S3), the peak intensity ratio of the (100)C facet to the (200)C facet is decreased from 1.05 to 0.95 when the substrates are changed from glass to the mesoporous TiO2 layer. With the same film thickness of TiO2 mesoporous layers of around 300 nm, a higher concentration of CH3NH3I/PbCl2 in DMF solvent (S4) leads the a peak intensity ratio of the (100)C facet to the (200)C facet dropping down to around 0.77. Therefore, for CH3NH3PbI3 perovskite prepared from PbCl2 involved precursor, the choice of substrates can subtly affect crystal orientations of CH3NH3PbI3 crystals.52 Comparing the crystal structure of CH3NH3PbI3 perovskite deposited from S3 with the S4 precursor, CH3NH3PbI3 crystals exhibit the preferred orientation of the (200)C facet over the (100)C facet for the more concentrated precursor. As a result, in the presence of chlorine, CH3NH3PbI3 crystals have more exposure of (110)T and (220)T facets (planar substrate) or (100)C and (200)C facets (mesoporous substrate). For many inorganic semiconductors, such as TiO2, the crystal facet is highly related to the surface electronic structure and thus determines the charge mobility.53,54 Particularly, it has been demonstrated in photoelectrochemical studies that nanowires or nanotubes of inorganic semiconductors generally show superiority in electron transportation along the oriented facet.55 Therefore, in the future, the effect of crystal orientation of CH3NH3PbI3 crystals on the charge lifetime and mobility of the CH3NH3PbI3 perovskite film needs to be studied to provide guidance for optimizing the photovoltaic performance of PSCs.30 In summary, we studied the influence of chlorine on the crystal structure of CH3NH3PbI3 perovskite using two different chlorine sources, PbCl2 and CH3NH3Cl, deposited on three different substrates. It is found that for the photoanodes composed of a thick TiO2 mesoporous layer (∼500 nm), under the involvement of chlorine, CH3NH3PbI3 crystals tend to form a cubic structure rather than a tetragonal structure, which is very different from what is generally observed for CH3NH3PbI3 perovskite at room temperature. It is speculated that the existence of chlorine can significantly reduce the cubic to 4382

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384

Letter

The Journal of Physical Chemistry Letters

(6) Oga, H.; Saeki, A.; Ogomi, Y.; Hayase, S.; Seki, S. Improved Understanding of The Electronic and Energetic Landscapes of Perovskite Solar Cells: High Local Charge Carrier Mobility, Reduced Recombination, and Extremely Shallow Traps. J. Am. Chem. Soc. 2014, 136, 13818−13825. (7) Miyata, A.; Mitioglu, A.; Plochocka, P.; Portugall, O.; Wang, J. T.W.; Stranks, S. D.; Snaith, H. J.; Nicholas, R. J. Direct Measurement of The Exciton Binding Energy and Effective Masses for Charge Carriers in Organic-Inorganic Tri-Halide Perovskites. Nat. Phys. 2015, 11, 582−587. (8) D’Innocenzo, V.; Grancini, G.; Alcocer, M. J. P.; Kandada, A. R. S.; Stranks, S. D.; Lee, M. M.; Lanzani, G.; Snaith, H. J.; Petrozza, A. Excitons versus Free Charges in Organo-Lead Tri-Halide Perovskites. Nat. Commun. 2014, 5, 3586. (9) Deschler, F.; Price, M.; Pathak, S.; Klintberg, L. E.; Jarausch, D. D.; Higler, R.; Huttner, S.; Leijtens, T.; Stranks, S. D.; Snaith, H. J.; et al. High Photoluminescence Efficiency and Optically Pumped Lasing in Solution-Processed Mixed Halide Perovskite Semiconductors. J. Phys. Chem. Lett. 2014, 5, 1421−1426. (10) Zhu, H.; Fu, Y.; Meng, F.; Wu, X.; Gong, Z.; Ding, Q.; Gustafsson, M. V.; Trinh, M. T.; Jin, S.; Zhu, X. Y. Lead Halide Perovskite Nanowire Lasers with Low Lasing Thresholds and High Quality Factors. Nat. Mater. 2015, 14, 636−642. (11) Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. (12) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. High-Performance Photovoltaic Perovskite Layers Fabricated through Intramolecular Exchange. Science 2015, 348, 1234−1237. (13) Edri, E.; Kirmayer, S.; Mukhopadhyay, S.; Gartsman, K.; Hodes, G.; Cahen, D. Elucidating The Charge Carrier Separation and Working Mechanism of CH3NH3PbI3−XClX Perovskite Solar Cells. Nat. Commun. 2014, 5, 3461. (14) Li, X.; Ibrahim Dar, M.; Yi, C.; Luo, J.; Tschumi, M.; Zakeeruddin, S. M.; Nazeeruddin, M. K.; Han, H.; Grätzel, M. Improved Performance and Stability of Perovskite Solar Cells by Crystal Crosslinking with Alkylphosphonic Acid ω-Ammonium Chlorides. Nat. Chem. 2015, 7, 703−711. (15) Unger, E. L.; Hoke, E. T.; Bailie, C. D.; Nguyen, W. H.; Bowring, A. R.; Heumuller, T.; Christoforo, M. G.; Mcgehee, M. D. Hysteresis and Transient Behavior in Current-Voltage Measurements of Hybrid-Perovskite Absorber Solar Cells. Energy Environ. Sci. 2014, 7, 3690−3698. (16) Wu, B.; Fu, K.; Yantara, N.; Xing, G.; Sun, S.; Sum, T. C.; Mathews, N. Charge Accumulation and Hysteresis in Perovskite-Based Solar Cells: An Electro-Optical Analysis. Adv. Energy Mater. 2015, DOI: 10.1002/aenm.201500829. (17) Wang, Q.; Chen, H. J.; Liu, G.; Wang, L. Z. Control of OrganicInorganic Halide Perovskites in Solid-State Solar Cells: A Perspective. Sci. Bull. 2015, 60, 405−418. (18) Zhang, W.; Saliba, M.; Moore, D. T.; Pathak, S. K.; Hörantner, M. T.; Stergiopoulos, T.; Stranks, S. D.; Eperon, G. E.; AlexanderWebber, J. A.; Abate, A. Ultrasmooth Organic−Inorganic Perovskite Thin-Film Formation and Crystallization for Efficient Planar Heterojunction Solar Cells. Nat. Commun. 2015, 6, 6142. (19) Colella, S.; Mosconi, E.; Fedeli, P.; Listorti, A.; Gazza, F.; Orlandi, F.; Ferro, P.; Besagni, T.; Rizzo, A.; Calestani, G.; et al. MAPbI3−XClX Mixed Halide Perovskite for Hybrid Solar Cells: The Role of Chloride as Dopant on The Transport and Structural Properties. Chem. Mater. 2013, 25, 4613−4618. (20) Unger, E. L.; Bowring, A. R.; Tassone, C. J.; Pool, V. L.; GoldParker, A.; Cheacharoen, R.; Stone, K. H.; Hoke, E. T.; Toney, M. F.; Mcgehee, M. D. Chloride in Lead Chloride-Derived Organo-Metal Halides for Perovskite-Absorber Solar Cells. Chem. Mater. 2014, 26, 7158−7165.

tetragonal transition temperature, thereby leading to the stable presence of cubic CH3NH3PbI3 perovskite at room temperature. Moreover, it is demonstrated that with the presence of chlorine, the preferred orientation of CH3NH3PbI3 crystals on TiO2 substrates can be shifted from the (110)T to (220)T facet (planar substrate) or the (100)C to (200)C facet (mesoporous substrate), confirming the influence of substrates on preferential orientation of perovskite. The findings reported herein provide new understanding of crystal formation of the perovskite layer and the interaction at the perovskite/TiO2 interface under the influence of chlorine, which could provide further guidance on the design of better PSCs. Future work on the effect of crystal orientation of CH3NH3PbI3 perovskite on its electronic properties, such as charge separation/transport, charge carrier mobilities, and trap states/densities needs to be conducted under the assistance of techniques, such as transient absorption spectroscopy, time-resolved photoluminescence, or newly applied confocal fluorescence microscopy.33



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpclett.5b01682. Experimental details, SEM, XPS, XRD, and UV−vis data (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the Australian Research Council (ARC) through its Discovery and Future Fellowship schemes. This work was performed in part at the QLD node of the Australian National Fabrication Facility, a company established under the National Collaborative Research Infrastructure Strategy to provide nano- and microfabrication facilities for Australia’s researchers. Q.W. acknowledges the support from the Chinese Scholarship Council (CSC).



REFERENCES

(1) Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Gratzel, M. Sequential Deposition as A Route to High-Performance Perovskite-Sensitized Solar Cells. Nature 2013, 499, 316−319. (2) Liu, M.; Johnston, M. B.; Snaith, H. J. Efficient Planar Heterojunction Perovskite Solar Cells by Vapour Deposition. Nature 2013, 501, 395−398. (3) De Wolf, S.; Holovsky, J.; Moon, S.-J.; Löper, P.; Niesen, B.; Ledinsky, M.; Haug, F.-J.; Yum, J.-H.; Ballif, C. Organometallic Halide Perovskites: Sharp Optical Absorption Edge and its Relation to Photovoltaic Performance. J. Phys. Chem. Lett. 2014, 5, 1035−1039. (4) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J. P.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341− 344. (5) Xing, G.; Mathews, N.; Sun, S.; Lim, S. S.; Lam, Y. M.; Grätzel, M.; Mhaisalkar, S.; Sum, T. C. Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342, 344−347. 4383

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384

Letter

The Journal of Physical Chemistry Letters (21) Williams, S. T.; Zuo, F.; Chueh, C.-C.; Liao, C.-Y.; Liang, P.-W.; Jen, A. K. Y. Role of Chloride in The Morphological Evolution of Organo-Lead Halide Perovskite Thin Films. ACS Nano 2014, 8, 10640−10654. (22) Edri, E.; Kirmayer, S.; Henning, A.; Mukhopadhyay, S.; Gartsman, K.; Rosenwaks, Y.; Hodes, G.; Cahen, D. Why Lead Methylammonium Tri-iodide Perovskite-Based Solar Cells Require A Mesoporous Electron Transporting Scaffold (but not Necessarily A Hole Conductor). Nano Lett. 2014, 14, 1000−1004. (23) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643−647. (24) Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-Hole Diffusion Lengths > 175 μm in SolutionGrown CH3NH3PbI3 Single Crystals. Science 2015, 347, 967−970. (25) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M.; Alarousu, E.; Buin, A.; Chen, Y.; Hoogland, S.; Rothenberger, A.; Katsiev, K.; et al. Low Trap-State Density and Long Carrier Diffusion in Organolead Trihalide Perovskite Single Crystals. Science 2015, 347, 519−522. (26) Pellegrino, G.; Colella, S.; Deretzis, I.; Condorelli, G. G.; Smecca, E.; Gigli, G.; La Magna, A.; Alberti, A. Texture of MAPbI3 Layers Assisted by Chloride on Flat TiO2 Substrates. J. Phys. Chem. C 2015, 119, 19808−19816. (27) Colella, S.; Mosconi, E.; Pellegrino, G.; Alberti, A.; Guerra, V. L. P.; Masi, S.; Listorti, A.; Rizzo, A.; Condorelli, G. G.; De Angelis, F.; Gigli, G. Elusive Presence of Chloride in Mixed Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 3532−3538. (28) Yu, H.; Wang, F.; Xie, F.; Li, W.; Chen, J.; Zhao, N. The Role of Chlorine in the Formation Process of “CH3NH3PbI3‑XClX” Perovskite. Adv. Funct. Mater. 2014, 24, 7102−7108. (29) Edri, E.; Kirmayer, S.; Kulbak, M.; Hodes, G.; Cahen, D. Chloride Inclusion and Hole Transport Material Doping to Improve Methyl Ammonium Lead Bromide Perovskite-Based High OpenCircuit Voltage Solar Cells. J. Phys. Chem. Lett. 2014, 5, 429−433. (30) Chen, Q.; Zhou, H.; Fang, Y.; Stieg, A. Z.; Song, T.-B.; Wang, H.-H.; Xu, X.; Liu, Y.; Lu, S.; You, J. The Optoelectronic Role of Chlorine in CH3NH3PbI3(Cl)-based Perovskite Solar Cells. Nat. Commun. 2015, 6, 7269. (31) Starr, D. E.; Sadoughi, G.; Handick, E.; Wilks, R. G.; Alsmeier, J. H.; Kohler, L.; Gorgoi, M.; Snaith, H. J.; Bar, M. Direct Observation of An Inhomogeneous Chlorine Distribution in CH3NH3PbI3−XClX Layers: Surface Depletion and Interface Enrichment. Energy Environ. Sci. 2015, 8, 1609−1615. (32) Wu, X.; Trinh, M. T.; Niesner, D.; Zhu, H.; Norman, Z.; Owen, J. S.; Yaffe, O.; Kudisch, B. J.; Zhu, X. Y. Trap States in Lead Iodide Perovskites. J. Am. Chem. Soc. 2015, 137, 2089−2096. (33) De Quilettes, D. W.; Vorpahl, S. M.; Stranks, S. D.; Nagaoka, H.; Eperon, G. E.; Ziffer, M. E.; Snaith, H. J.; Ginger, D. S. Impact of Microstructure on Local Carrier Lifetime in Perovskite Solar Cells. Science 2015, 348, 683−686. (34) Hebenstreit, E. L. D.; Hebenstreit, W.; Geisler, H.; Ventrice, C. A., Jr; Hite, D. A.; Sprunger, P. T.; Diebold, U. The Adsorption of Chlorine on TiO2 (1 1 0) Studied with Scanning Tunneling Microscopy and Photoemission Spectroscopy. Surf. Sci. 2002, 505, 336−348. (35) Vogtenhuber, D.; Podloucky, R.; Redinger, J. Ab Initio Study of Atomic Cl Adsorption on Stoichiometric and Reduced Rutile TiO2 (110) Surfaces. Surf. Sci. 2000, 454−456, 369−373. (36) Mosconi, E.; Ronca, E.; De Angelis, F. First-Principles Investigation of The TiO2/Organohalide Perovskites Interface: The Role of Interfacial Chlorine. J. Phys. Chem. Lett. 2014, 5, 2619−2625. (37) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. I. Chemical Management for Colorful, Efficient, and Stable Inorganic− Organic Hybrid Nanostructured Solar Cells. Nano Lett. 2013, 13, 1764−1769. (38) Zhang, D. D.; Eaton, S. W.; Yu, Y.; Dou, L. T.; Yang, P. D. Solution-Phase Synthesis of Cesium Lead Halide Perovskite Nanowires. J. Am. Chem. Soc. 2015, 137, 9230−9233.

(39) Poglitsch, A.; Weber, D. Dynamic Disorder in Methylammoniumtrihalogenoplumbates (II) Observed by Millimeter-Wave Spectroscopy. J. Chem. Phys. 1987, 87, 6373−6378. (40) Kawamura, Y.; Mashiyama, H.; Hasebe, K. Structural Study on Cubic−Tetragonal Transition of CH3NH3PbI3. J. Phys. Soc. Jpn. 2002, 71, 1694−1697. (41) Baikie, T.; Fang, Y.; Kadro, J. M.; Schreyer, M.; Wei, F.; Mhaisalkar, S. G.; Graetzel, M.; White, T. J. Synthesis and Crystal Chemistry of The Hybrid Perovskite (CH3NH3)PbI3 for Solid-State Sensitised Solar Cell Applications. J. Mater. Chem. A 2013, 1, 5628− 5641. (42) Im, J.-H.; Jang, I.-H.; Pellet, N.; Grätzel, M.; Park, N.-G. Growth Of CH3NH3PbI3 Cuboids with Controlled Size for High-Efficiency Perovskite Solar Cells. Nat. Nanotechnol. 2014, 9, 927−932. (43) Dar, M. I.; Arora, N.; Gao, P.; Ahmad, S.; Grätzel, M.; Nazeeruddin, M. K. Investigation Regarding The Role of Chloride in Organic−Inorganic Halide Perovskites Obtained from Chloride Containing Precursors. Nano Lett. 2014, 14, 6991−6996. (44) Onoda-Yamamuro, N.; Matsuo, T.; Suga, H. Calorimetric and IR Spectroscopic Studies of Phase Transitions in Methylammonium Trihalogenoplumbates (II). J. Phys. Chem. Solids 1990, 51, 1383−1395. (45) Wasylishen, R. E.; Knop, O.; Macdonald, J. B. Cation Rotation in Methylammonium Lead Halides. Solid State Commun. 1985, 56, 581−582. (46) Roiati, V.; Mosconi, E.; Listorti, A.; Colella, S.; Gigli, G.; De Angelis, F. Stark Effect in Perovskite/TiO2 Solar Cells: Evidence of Local Interfacial Order. Nano Lett. 2014, 14, 2168−2174. (47) Jeon, N. J.; Noh, J. H.; Kim, Y. C.; Yang, W. S.; Ryu, S.; Seok, S. I. Solvent Engineering for High-Performance Inorganic−Organic Hybrid Perovskite Solar Cells. Nat. Mater. 2014, 13, 897−903. (48) Williams, S. T.; Chueh, C.-C.; Jen, A. K. Y. Navigating OrganoLead Halide Perovskite Phase Space via Nucleation Kinetics Toward a Deeper Understanding of Perovskite Phase Transformations and Structure−Property Relationships. Small 2015, 11, 3088−3096. (49) Tidhar, Y.; Edri, E.; Weissman, H.; Zohar, D.; Hodes, G.; Cahen, D.; Rybtchinski, B.; Kirmayer, S. Crystallization of Methyl Ammonium Lead Halide Perovskites: Implications for Photovoltaic Applications. J. Am. Chem. Soc. 2014, 136, 13249−13256. (50) Pistor, P.; Borchert, J.; Fränzel, W.; Csuk, R.; Scheer, R. Monitoring The Phase Formation of Coevaporated Lead Halide Perovskite Thin Films by In Situ X-Ray Diffraction. J. Phys. Chem. Lett. 2014, 5, 3308−3312. (51) Moore, D. T.; Sai, H.; Tan, K. W.; Smilgies, D.-M.; Zhang, W.; Snaith, H. J.; Wiesner, U.; Estroff, L. A. Crystallization Kinetics of Organic−Inorganic Trihalide Perovskites and The Role of The Lead Anion in Crystal Growth. J. Am. Chem. Soc. 2015, 137, 2350−2358. (52) Listorti, A.; Juarez-Perez, E. J.; Frontera, C.; Roiati, V.; GarciaAndrade, L.; Colella, S.; Rizzo, A.; Ortiz, P.; Mora-Sero, I. Effect of Mesostructured Layer upon Crystalline Properties and Device Performance on Perovskite Solar Cells. J. Phys. Chem. Lett. 2015, 6, 1628−1637. (53) Li, C. H.; Koenigsmann, C.; Ding, W. D.; Rudshteyn, B.; Yang, K. R.; Regan, K. P.; Konezny, S. J.; Batista, V. S.; Brudvig, G. W.; Schmuttenmaer, C. A.; et al. Facet-Dependent Photoelectrochemical Performance of TiO2 Nanostructures: An Experimental and Computational Study. J. Am. Chem. Soc. 2015, 137, 1520−1529. (54) Pan, J.; Liu, G.; Lu, G. M.; Cheng, H. M. On The True Photoreactivity Order Of {001}, {010}, and {101} Facets of Anatase TiO2 Crystals. Angew. Chem., Int. Ed. 2011, 50, 2133−2137. (55) Zhu, K.; Neale, N. R.; Miedaner, A.; Frank, A. J. Enhanced Charge-Collection Efficiencies and Light Scattering in Dye-Sensitized Solar Cells Using Oriented TiO2 Nanotubes Arrays. Nano Lett. 2007, 7, 69−74.

4384

DOI: 10.1021/acs.jpclett.5b01682 J. Phys. Chem. Lett. 2015, 6, 4379−4384