Compositional Engineering To Improve the ... - ACS Publications


Compositional Engineering To Improve the...

2 downloads 105 Views 890KB Size

Subscriber access provided by READING UNIV

Article

Compositional Engineering to Improve the Stability of Lead Halide Perovskites: A Comparative Study of Cationic and Anionic Dopants Kianoosh Poorkazem, and Timothy L Kelly ACS Appl. Energy Mater., Just Accepted Manuscript • DOI: 10.1021/acsaem.7b00065 • Publication Date (Web): 05 Dec 2017 Downloaded from http://pubs.acs.org on December 5, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Energy Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Compositional Engineering to Improve the Stability of Lead Halide Perovskites: A Comparative Study of Cationic and Anionic Dopants Kianoosh Poorkazem and Timothy L. Kelly* Department of Chemistry, University of Saskatchewan, 110 Science Place, Saskatoon, SK, S7N 5C9, Canada Keywords: photovoltaics, humidity, oxygen, degradation, longevity

ABSTRACT

The instability of perovskite solar cells is the single greatest barrier to their commercialization. While a number of studies have now looked at the effect of perovskite composition on device stability, many of these have examined only a single compositional variable. With many of these studies having been carried out under different environmental conditions, and still others lacking environmental controls entirely, it is often difficult to compare the relative effect of various cationic or anionic dopants. To address this knowledge gap, we fabricated CH3NH3PbI3-based solar cells where either the methylammonium or iodide ions were replaced with 20 mol% of a dopant ion (ethylammonium, formamidinium, bromide, or chloride). We then assessed their

ACS Paragon Plus Environment

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

stability in either a controlled 85% relative humidity environment or under one sun illumination in air; both conditions have been previously shown to rapidly decompose CH3NH3PbI3. Of the dopants studied, the formamidinium cations imparted the best moisture resistance and the resulting perovskite displayed the lowest photochemical reactivity. We attribute the improved stability of the formamidinium-doped perovskite to the more delocalized positive charge of the formamidinium cation.

1. Introduction Lead halide perovskites are an exciting class of hybrid organic-inorganic semiconductor, and have found application in a wide variety of optoelectronic devices, such as solar cells,1 lightemitting diodes,2 photodetectors,3 and lasers4. In particular, enormous interest has been paid to perovskite solar cells (PSCs), since the perovskites are derived from relatively Earth-abundant materials, and the devices can be manufactured using highly scalable, low-temperature fabrication processes. However, in order to be commercially viable, PSCs need to be robust enough to withstand a wide variety of environmental conditions. Unfortunately, this is not yet the case, and lead halide perovskites have been shown to be unstable under a variety of real-world conditions, such as high temperatures,5 humidity,6,7 and the synergetic effects of oxygen and light.8 Prolonged exposure to harsh conditions leads to perovskite decomposition, and ultimately, to device failure. One of the most deleterious environmental stimuli for perovskites is high relative humidity (RH). In high RH environments, the methylammonium (MA) lead iodide perovskite absorbs water, which can lead to the reversible formation of a crystalline monohydrate phase.9

ACS Paragon Plus Environment

2

Page 3 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Upon prolonged exposure, further hydration occurs, producing both a dihydrate phase and PbI2.6,7,9 Ultimately, the hydrate phases have reduced dimensionality, leading to a loss of both light absorption and carrier transport, rendering them useless in device applications. In addition to water vapor, Bryant et al. showed that the synergetic effect of molecular oxygen and light was highly detrimental to CH3NH3PbI3 PSCs.8 The authors demonstrated that the photoexcitation of CH3NH3PbI3 in the presence of oxygen led to the formation of superoxide anions via the reduction of oxygen by conduction band electrons. The superoxide then readily deprotonated the methylammonium cation to decompose the perovskite into PbI2 and CH3NH2.10 As a result of these environmental instabilities, unencapsulated PSCs can fail in a matter of hours or days, depending on the perovskite composition and the device architecture. Numerous strategies have been proposed to improve the stability of PSCs, including the modification of interfacial layers,11-16 altering the perovskite deposition method,17-23 or changing the perovskite composition.24-33 Although the improvement of interfacial layers and electrodes is obviously extremely important, tuning the perovskite composition itself is a highly attractive strategy, in that a more stable perovskite layer could be used in a variety of different device architectures, and would also be useful in other perovskite-based optoelectronic devices. In an example of this approach, the hydrophobicity of 1,1,1-trifluoroethylammonium iodide was used to impart a higher moisture resistance to CH3NH3PbI3-based cells; however, the improvement was limited to a small amount of this additive, since the addition of larger amounts would shift the perovskite structure from 3D to 2D, sacrificing carrier mobility in the process.33 Therefore, changes to the perovskite composition should ideally be done without reducing the dimensionality.

ACS Paragon Plus Environment

3

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

In general, a lead halide perovskite of the general formula APbX3 is prepared by the reaction of an AX salt with a lead halide (PbX2), where the A-site cation is of the appropriate size to occupy the void spaces of the PbX3 framework. Although the CH3NH3PbI3 perovskite was the first material used in PSCs, other monovalent cations, such as formamidinium (FA), can be readily substituted in the A-site.5,34 Ethylammonium (EA) has been used in a few studies;35-37 however, the larger size of EA in pure EAPbI3 causes the formation of 1D chain-like structures. As for the halides, iodide can be partially replaced by bromide or chloride in order to tune the optoelectronic properties of the resulting material. Although chloride is too small to directly substitute iodide in the APbX3 lattice, its presence in the perovskite precursor solution has been shown to improve the crystallization of the perovskite layer, and affects the photovoltaic performance of fabricated devices.38,39 While individual studies have looked at the effect of some of these ion substitutions on perovskite stability, they are often carried out using different testing protocols or environmental conditions, and no single study has directly compared their efficacy. Here, we have carried out a comparative study on how various ion substitutions affect the stability of lead halide perovskites. Thin films of MAPbI3 and four different mixed-ion perovskites − MAPbI3−xBrx, MAPbI3(Cl), (MA)1−x(EA)xPbI3, and (MA)1−x(FA)xPbI3 − were prepared. Their stability in an 85% RH environment was evaluated, as was their photochemical stability in the presence of molecular oxygen. Overall, the FA-doped samples appeared to have an increased resistance to the effects of moisture, and a lower photochemical reactivity, while the various other ion substitutions had little to no impact on the stability of the films.

2. Experimental Section

ACS Paragon Plus Environment

4

Page 5 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

2.1. Perovskite film fabrication. Glass slides (2.5 cm × 2.5 cm) were cleaned by sequentially sonicating for 20 min in a 2% (v/v) aqueous Extran 300 solution, deionized water, and isopropanol. They were dried for at least 2 h in an oven and UV/ozone treated for 15 min immediately prior to use. A combination of interlayer diffusion40 and solvent annealing63 (with minor modifications) was used to prepare the perovskite thin films. In a nitrogen-filled glovebox, a 1.00 M solution of PbI2 in anhydrous N,N-dimethylformamide was prepared and pre-heated to 100 °C, as was the glass substrate. A PbI2 layer was deposited by spin coating at 3000 r.p.m. for 30 s, followed by 3 min of drying at room temperature, and 5 min of annealing at 100 °C. The film was removed from the glovebox, and 700 µL of a 0.050 M solution of MAI (20 mol% of the MAI was replaced by MABr, MACl, EAI or FAI for the mixed-ion perovskites) in anhydrous 2propanol was dispensed evenly across the film. After 45 s, the substrate was spun at 4000 r.p.m. for 20 s. One of two annealing treatments was then used. In method A, the samples were placed on a hot plate at either 140 °C (for (MA)1-x(FA)xPbI3) or 100 °C (all other samples) for a period of 5 min. In method B, the same temperatures were used, but the annealing was carried out in the glovebox, in the presence of DMF vapor (as shown in Figure S7). Method A was used for the samples in Figures 1-3, 5, S2-S3, S5, and S10-S11, while method B was used to prepare samples for all other data sets. For P3HT coated samples, 20 mg P3HT, 3.4 µL 4-tert-butylpyridine, and 6.8 µL of a Li-TFSI solution (28 mg/mL in acetonitrile) were dissolved in 1 mL of anhydrous chlorobenzene, and a thin film deposited by spin coating at 1000 r.p.m. for 30 s and at 3000 r.p.m. for 10 s.

2.2. Characterization. pXRD patterns were measured by a PANalytical Empyrean diffractometer operating with a Co source. Absorption spectra were acquired on a Cary 6000i

ACS Paragon Plus Environment

5

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 28

UV-vis-NIR spectrophotometer. SEM images were taken by a Hitachi SU8010 microscope. Fluorescence spectra were acquired on a PTI fluorometer using an excitation wavelength of 460 nm. Atomic force microscopy was carried out using a Dimension Hybrid Nanoscope system (Bruker/Veeco Metrology). Solar cell efficiency measurements were made using a 450 W Class AAA solar simulator, equipped with an AM1.5G filter (Sol3A, Oriel Instruments). The illumination intensity was set to 1 sun using a standard silicon reference cell (91150V, Oriel Instruments). The active area of the cell was defined to be 0.101 cm2 using a non-reflective metal aperture mask. All measurements were made inside the glovebox using a Keithley 2420 sourcemeasure unit. Density functional theory calculations were carried out using the Spartan software package; optimized equilibrium geometries and electrostatic potential maps were calculated at the M06/6-31+G(d) level of theory.

3. Results and discussion 3.1. Fabrication and characterization of perovskite thin films Perovskite films were first fabricated and characterized based on a two-step spin-coating process (Figure 1a). In all cases, a thin film of PbI2 was first deposited by spin-coating from a N,Ndimethylformamide solution. In the second step, a low concentration (0.050 M) solution of MAI in 2-propanol was spin-coated onto the PbI2 layer; the solution was left quiescent for 45 s prior to spinning in order to allow time for the MAI to intercalate into the PbI2 thin film and for conversion to the perovskite to occur.40 To produce the various mixed-ion perovskites, 20 mol% of the MAI in the second step was replaced by the appropriate halide salt (either MABr, MACl, EAI, or FAI, Figure 1a). Importantly, FA-containing perovskites often require a thermal annealing step (with temperatures as high as 170 ºC) to go from a yellow, hexagonal δ-phase to a black, trigonal α-phase.41,42 In this work, we observed that thermal annealing at 140 °C was

ACS Paragon Plus Environment

6

Page 7 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

sufficient to produce dark films of (MA)1−x(FA)xPbI3, with no δ-phase observable in the pXRD pattern (Figure 1b). The relatively lower temperatures required here are likely due to the substantial mole fraction of MAI used in this case, as opposed to the higher temperatures required to produce pure FAPbI3. All other films were annealed at 100 °C, with details provided in the Experimental Section.

ACS Paragon Plus Environment

7

ACS Applied Energy Materials

b

c (α hν )2 (µm-2eV2)

§

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

16.5

16.0

(MA)1–x(FA)xPbI3 (MA)1–x(EA)xPbI3 MAPbI3(Cl)

30

MAPbI3 MAPbI3-xBrx MAPbI3(Cl) MA(1-x)EAxPbI3 MA(1-x)FAxPbI3

20

10

MAPbI3–xBrx MAPbI3

10

15

20

25

30

35

40

45

50

0 1.54 1.56 1.58 1.60 1.62 1.64 1.66 1.68 1.70

2θ (degree)

Energy (eV)

Figure 1. (a) Schematic of the two-step spin-coating procedure used for the deposition of mixedion perovskites; (b) pXRD patterns and (c) Tauc plots of perovskite thin films. In (b), the peak marked with § is due to residual PbI2, and the inset shows an expanded view of the perovskite peak at ca. 16°; in (c), experimental data are shown as solid lines, while least squares fits to the linear region are shown as dashed lines.

To assess whether the various dopants were actually incorporated into the perovskite lattice, the pXRD patterns of each film were measured (Figure 1b). The most prominent perovskite peak is the (110) reflection, observed at ca. 16° 2θ (Figure 1b, inset). With bromide substitution, this peak shifts to higher angle, consistent with a decrease in the lattice parameters, owing to the substitution of the smaller bromide ion. For both the (MA)1−x(EA)xPbI3 and (MA)1−x(FA)xPbI3 samples, the peak shifts to smaller diffraction angles, commensurate with the larger size of the EA and FA cations. No shift is observed for the MAPbI3(Cl) sample; consistent

ACS Paragon Plus Environment

8

Page 9 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

with other literature reports, the chloride ion is too small to effectively replace iodide in the lattice, and so no ion substitution occurs in this sample. Similar shifts (or lack thereof) are observed in the other peaks of the diffraction pattern, and no additional peaks are observed. These shifts were consistent with literature reports of mixed-ion perovskites with these particular dopants;35,43,44 however, we nonetheless synthesized powdered samples of these same materials for comparison, and the same shifts were observed in the pXRD patterns of the powders (Figure S1). The inclusion of the different dopants in the perovskite lattice was further evaluated by absorption spectroscopy (Figure S2). Tauc plots (Figure 1c) were used to estimate the optical bandgap of each sample. The data revealed a blue-shift in the bandgap of MAPbI3−xBrx (1.60 eV) relative to MAPbI3 (1.58 eV), and a red-shift in the bandgap of (MA)1−x(FA)xPbI3 (1.56 eV); given the bandgaps of pure MAPbBr3 (2.24 eV)45 and α-FAPbI3 (1.41 eV)46 single crystals, this is consistent with partial ion substitution in the thin films prepared here. In addition, we also observed a small red shift for the (MA)1−x(EA)xPbI3 and MAPbI3(Cl) samples (both 1.56 eV). This may be due to the variations in crystallite size observed (Figure 2, top row). As shown, the crystallite size of MAPbI3−xBrx and (MA)1−x(FA)xPbI3 were very similar to MAPbI3; however, the MAPbI3(Cl) and (MA)1−x(EA)xPbI3 crystallites were significantly larger. Depth histograms using atomic force microscopy (Figure S3) further confirmed this size difference. The larger crystallite sizes observed in these two perovskite samples results in a higher surface roughness and stronger Mie scattering. This can in turn produce an artificial red-shift in the optical bandgap, as noted by Tian and Scheblykin.47

ACS Paragon Plus Environment

9

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

Figure 2. SEM images of perovskite thin films before (top row) and after (bottom row) exposure to an 85% RH environment for 4 h. Scale bars are all 1 µm.

3.2. Moisture stability The effect of humidity on the various perovskite thin films was then evaluated. Samples were stored at room temperature in a gas-tight desiccator fitted with inlet and outlet ports (Figure S4); a stream of nitrogen gas at 85% RH (prepared using a system of bubblers and flow controllers)6 was introduced into the chamber, and the humidity was validated downstream of the sample chamber by way of an external sensor. SEM was used to capture the morphological changes occurring in the perovskite films in the early stages of moisture-induced decomposition (Figure 2, Figure S5). In most cases, after 4 h of exposure, two distinct regions could be observed in the SEM images – a bright region where individual perovskite crystallites were still intact, and a darker region with a less well defined, more amorphous appearance. However, even in the brighter, less disturbed regions of the images, there is evidence of rounding at the edges of the crystallites, and adjacent particles have begun to fuse together. This is similar to what Gangishetty et al. observed in MAPbI3 films grown in a 40-60% RH environment.48 It suggests

ACS Paragon Plus Environment

10

Page 11 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

that in the early stages of the hydration process, water is absorbed by the perovskite crystallites; this in turn facilitates ionic diffusion in the solid, disrupting the crystalline perovskite lattice and leading to increased amorphization of the film. In order to evaluate these changes in crystallinity more directly, pXRD was used to monitor the diffraction patterns of the films as a function of exposure time to water vapor (Figure 3). However, one of the problems with using conventional benchtop diffraction to follow moisture-induced changes in the perovskite structure is that the various hydrate phases are unstable at the low relative humidities typically found in ambient conditions.7 Therefore, we first coated the perovskite films with a thin layer of the hole-transport material poly(3hexylthiophene) (P3HT),49 which has been shown to decelerate the moisture-induced decomposition of MAPbI3;6 it therefore also serves to decelerate the reverse dehydration process, enabling us to detect the formation of any intermediate hydrate phases. The pXRD patterns of MAPbI3 follow the general trend expected from previous in situ absorbance spectroscopy and in situ grazing incidence wide angle X-ray scattering (GIWAXS) experiments; after an initial induction period of 12 – 24 h, in which the water vapor permeates the P3HT capping layer to reach the perovskite below, there is the appearance of a crystalline monohydrate phase (identified by diagnostic diffraction peaks at ca. 9.8 and 12.1°), while the perovskite diffraction peaks decrease substantially in intensity. Upon further exposure to water vapor, the hydrated intermediate also disappears, and is replaced by PbI2, as indicated by the growth of the (001) PbI2 reflection at 14.6°. Very similar behavior is observed for the MAPbI3(Cl) and (MA)1−x(EA)xPbI3 samples, both of which show evidence of intermediate monohydrate phases. However, in contrast, no hydrate phases were observed in the MAPbI3−xBrx and (MA)1−x(FA)xPbI3 samples, regardless of exposure time. Additionally, if the normalized intensity

ACS Paragon Plus Environment

11

ACS Applied Energy Materials

of the perovskite reflection at 16° (the most intense perovskite reflection, and therefore the most sensitive to changes in perovskite crystallinity) is plotted as a function of time (Figure 3b), these two samples display the slowest perovskite decomposition kinetics. The relatively better crystal stability of MAPbI3−xBrx compared to MAPbI3 is in line with early experiments on bromidedoped PSCs, although the conditions employed here (85% RH) were substantially harsher than what was studied previously.50 An analysis of the peak width via the Scherrer equation shows relatively little change in crystallite size over the course of the experiment (Figure S6), suggesting that it is the number of perovskite crystallites, rather than their size, that decreases over the course of the experiment.

a

MAPbI3

MAPbI3–xBrx

MAPbI3(Cl)

b

(MA)1–x(EA)xPbI3 (MA)1–x(FA)xPbI3

Normalized Intensity

190 h 140 h 90 h 45 h 24 h 12 h 4h 0h

MAPbI3

(MA)1–x(EA)xPbI3

MAPbI3–xBrx

(MA)1–x(FA)xPbI3

MAPbI3(Cl)

1.2

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

1.0 0.8 0.6 0.4 0.2 0.0

10

15

10

15

10

15

10

15

10

15

2θ θ (degree)

0

24

48

72

96

120

144

168

192

Time (h)

Figure 3. (a) pXRD patterns of P3HT-coated perovskite films after exposure to an 85% RH environment for various times; (b) normalized intensity of the perovskite peak at ca. 16° as a function of exposure time.

As part of this work, we also observed that the annealing conditions used to prepare the perovskite films had an effect on their stability. Samples that were annealed for 5 min under ambient laboratory conditions showed slightly less resistance to the effects of moisture than

ACS Paragon Plus Environment

12

Page 13 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

those that were solvent annealed in the presence of DMF vapor (Figure S7-S8). However, the films otherwise appeared to be quite similar (Figure S9); with this in mind, solvent-annealing was used to prepare the films for further tests of their humidity and photochemical stability. Since the perovskite undergoes a large decrease in extinction coefficient during decomposition, optical absorption spectroscopy provides a very sensitive means of monitoring the decomposition process (Figure 4). Even a simple visual inspection of perovskite thin films exposed to 85% RH for a period of 20 h (Figure 4a) is revealing; except for (MA)1−x(FA)xPbI3, all of the other samples showed evidence of almost complete decomposition. On continued exposure, the (MA)1−x(FA)xPbI3 films only showed signs of significant decomposition after 100 h, again suggesting that the FA was more effective at stabilizing the perovskite structure than other types of ion-substitutions. In order to provide more quantitative information, the optical absorption spectra of P3HT-coated perovskites were monitored (Figure S10), and the normalized absorbance at 520 nm is shown in Figure 4b. Consistent with the more qualitative study, the (MA)1−x(FA)xPbI3 sample showed the slowest decomposition kinetics, showing less than a 10% loss of absorbance after 100 h; in contrast, the other four samples decomposed 4-6 times faster. The behavior of the (MA)1−x(FA)xPbI3 perovskite was also different in one other aspect. A Tauc plot (Figure 4c) reveals that the bandgap of the perovskite shifts over time, from 1.56 eV originally to 1.45 eV after > 300 h of exposure to the humid environment. This is much closer to the absorption edge of α-phase FAPbI3 (1.41 eV).46 No such shift was observed in the spectra of the other perovskites. This suggests that the MA cation preferentially interacts with water and is gradually depleted from the perovskite structure, leaving behind a material enriched in the FA component. It underscores the impact that the FA cation can have over the stability of the resultant perovskite.

ACS Paragon Plus Environment

13

ACS Applied Energy Materials

c MAPbI3 MAPbI3–xBrx MAPbI3(Cl) (MA)1–x(EA)xPbI3 (MA)1–x(FA)xPbI3

1.0

0.8

0.6

0.4

0h 23 h 32 h 49 h 69 h 90 h 112 h 157 h 207 h 255 h 326 h

50

40

(αhν )2 (µm-2eV-2)

b Normalized Absorbance

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

30

20

10

0.2 0

100

200

300

400

500

0 1.40

1.45

Time (h)

1.50

1.55

1.60

1.65

1.70

Energy (eV)

Figure 4. (a) Photographs of perovskite thin films, (b) normalized absorbance at 520 nm for P3HT-coated perovskite thin films, and (c) Tauc plot for P3HT-coated (MA)1−x(FA)xPbI3 as a function of exposure time to an 85% RH environment. The thickness of the perovskite layer was assumed to be constant throughout the damp test.

3.3. Thermal stability and device lifetime The previous studies suggest that out of the various cationic and anionic dopants studied, FA imparts the best protection against the effects of moisture. In order to see whether the same trend would hold true for other environmental stimuli, we subjected the various perovskite films to thermal treatment at 85 °C in the glovebox (0% RH), and measured pXRD patterns and UV/vis spectra as a function of heating time (Figure 5, Figure S11). All samples displayed the same

ACS Paragon Plus Environment

14

Page 15 of 28

trend, which was a slow, monotonic decline in the intensity of both the perovskite diffraction peaks and optical absorption bands over a period of ca. 600 h. Analysis of the films by SEM at the early stages of decomposition showed relatively little change in crystallite morphology, and relatively little difference between the various perovskite compositions (Figure S12). This is supported by the line widths of the pXRD peaks (Figue S13), which display little in the way of a trend over the course of the experiment. The data suggests that at least at the doping levels studied here (x = 0.20), none of the ions studied impart any additional thermal stability.

b MAPbI3 MAPbI3–xBrx MAPbI3(Cl) (MA)1–x(EA)xPbI3 (MA)1–x(FA)xPbI3

1.2 1.0 0.8 0.6 0.4 0.2

Normalized Absorbance

a Normalized Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1.0

MAPbI3 MAPbI3–xBrx MAPbI3(Cl) (MA)1–x(EA)xPbI3 (MA)1–x(FA)xPbI3

0.8

0.6

0.4 0.0 0

200

400

600

800

1000

0

Time (h)

200

400

600

800

1000

Time (h)

Figure 5. (a) Normalized intensity of the perovskite peak at ca. 16°, and (b) normalized absorbance at 520 nm as a function of annealing time at 85 °C.

In order to elucidate the impact of these two environmental factors – elevated humidity and temperature – on device performance, we prepared a series of perovskite solar cells using either a FTO/TiO2/perovskite/P3HT/Au or FTO/TiO2/perovskite/P3HT/Ag device architecture. They were exposed to the same environmental conditions as described previously, and their PCE was measured as a function of time (Figure S14). The devices stored in a humid environment showed a steady decline in performance over a period of ca. one month, while the devices stored

ACS Paragon Plus Environment

15

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

at elevated temperature degraded slowly over the first 10 days, followed by rapid cell failure. In both studies, there was little impact of perovskite composition on the device lifetime; however, given the better moisture resistance of the FA-containing perovskites, this suggests that the perovskite itself is not the limiting factor for device lifetime. Instead, we ascribe the cell degradation observed under humid conditions to an increased mobility of iodide ions in the presence of water; this in turns leads to the corrosion of the silver top contact and the formation of AgI, as has been noted elsewhere.51-54 The data suggest that the stability of electrodes and interfaces may be more significant than the decomposition of the perovskite absorber layer.

3.4. Photochemical stability Based on the results of the moisture and thermal stability tests, we identified the FA cation as the most promising candidate for producing more stable perovskite layers. We therefore prepared two additional perovskite films with higher FA content (x = 0.4 and 0.6), and evaluated their photochemical stability (at 1 sun intensity) in the presence of molecular oxygen (i.e., in dry air). Again, the effects of perovskite decomposition are apparent from even a qualitative visual inspection (Figure 6). A color change from brown to yellow began to occur for both the 20% and 40% FA samples after only 16 h of irradiation time, highlighting the deleterious effects of the photochemically-generated superoxide anion.8 Promisingly, the sample with 60% FA showed a greater resistance to photochemical bleaching, with the appearance of yellow decomposition products only apparent after > 60 h of continuous irradiation.

ACS Paragon Plus Environment

16

Page 17 of 28

Normalized Absorbance

b 1.0

20% FA 40% FA 60% FA

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0

12

24

36

48

60

Time (h)

d

820

Emission Intensity (a.u.)

Band Edge (nm)

c

810

800

1.2x106

20% FA 40% FA 60% FA

1.0x106 8.0x105 6.0x105 4.0x105 2.0x105

790

0.0 0

12

24

36

0

12

24

Time (h)

36

48

60

Time (h)

e

f

1.2

0.8

Controls (N2 / dark)

20% FA 40% FA 60% FA

0.6 0.4 0.2

Normalized Intensity

1.0

Normalized Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1.0 20% FA 40% FA 60% FA

0.8 0.6 0.4 0.2

0.0 0

12

24

36

48

60

0

12

24

36

48

60

Time (h)

Time (h)

Figure 6. (a) Photographs, (b) normalized absorbance at 520 nm, (c) band edge position (as determined from Tauc plots, (d) photoluminescence intensity, (e) normalized intensity (relative to the intensity at t = 0 h) of the perovskite diffraction peak at ca. 16°, and (f) normalized intensity (relative to the intensity at t = 65 h) of the PbI2 (001) diffraction peak as a function of irradiation time in air. All data points are the average of two separate measurements.

ACS Paragon Plus Environment

17

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 28

A more quantitative measure of the photochemical stability of the FA-containing samples can be obtained by monitoring their absorption spectra as a function of irradiation time. There is very little change in the spectra of any of the samples when stored in the dark (Figure S15), highlighting the photochemical nature of this particular decomposition mechanism. For the samples irradiated in air, the absorbance at 520 nm is normalized and plotted as a function of time (Figure 6b), showing that the 60% FA sample decomposes more slowly (and to a lesser overall extent) than the samples with lower FA content. This is accompanied by a shift in the position of the band edge (Figure 6c), as determined by Tauc plots of the UV/vis data (Figure S16). The band edge red-shifts for all three samples, although the red-shift is less pronounced for samples with higher FA content. This is consistent with the superoxide anion preferentially deprotonating the MA cation, leading to an enrichment in FA content over the course of the decomposition process. Changes in the photoluminescence spectra of the films were also observed over the course of the photochemical study (Figure S17, Figure 6d). There is a marked increase in the photoluminescence intensity at early stages of the decomposition process, followed by a pronounced drop at later times. The low emission intensity observed initially may be due to the presence of numerous non-radiative recombination centers and trap states within the perovskite film. As noted by others, the nature of these trap states may be under-coordinated lead ions in the lead halide framework.55 These trap states can therefore be passivated by using Lewis bases such as pyridine or thiophene. This is supported by another recent study, which reported an order of magnitude higher photoluminescence when the PbI2:FAI molar ratio was changed from 1:1 to 0.9:1, producing a slight excess of iodide ions.56 In the present case, it may be possible that as the

ACS Paragon Plus Environment

18

Page 19 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

photogenerated superoxide deprotonates the methylammonium cation, it leaves behind additional iodide ions that can help passivate trap states; alternatively, the superoxide anion may coordinate to the lead ion itself. However, in either case, as the decomposition process unfolds further, the photoluminescence intensity inevitably drops to zero. The photochemical decomposition process was also followed by pXRD (Figure S18, Figure 6e,f). The intensity of the perovskite peak at ca. 16° 2θ is plotted in Figure 6e, while the intensity of the PbI2 (001) reflection is plotted in Figure 6f; a Scherrer analysis of the line widths is shown in Figure S19. Again, control samples that were stored under nitrogen in the dark showed no evidence of decomposition. When irradiated in air, the intensities of the perovskite reflections drop rapidly, with the rate decreasing as the amount of FA is increased; this is accompanied by a rapid increase in the amount of PbI2 present. The perovskite crystallite size calculated via the Scherrer equation also shrinks during the irradiation process, highlighting the degradation of individual perovskite grains. Notably, in all three samples a small amount of δphase FAPbI3 (2θ = 13.6°) was observed at intermediate times (16 h), that quickly disappeared as the decomposition progressed. This may be a result of the preferential decomposition of the MA cation, which leaves behind FA-rich domains that crystallize into the δ-phase. At later times, even the FA-rich perovskite domains decompose, leaving behind PbI2 as the only crystalline byproduct. All of these results are consistent with previous reports suggesting that FAPbI357 and MA0.5FA0.5PbI358 have better photochemical stability than MAPbI3; however, they more clearly separate out photochemical and moisture-induced effects, both of which we have shown to have an impact on the stability of perovskite thin films. The moisture-induced and photochemical decomposition processes are strongly influenced by the ammonium group of the MA cation: hydrogen bonding between the oxygen of

ACS Paragon Plus Environment

19

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 28

a water molecule the hydrogen of the MA cation drives the formation of crystalline hydrate phases,6,59 while the relative acidity of the MA cation leaves it vulnerable to photogenerated superoxide.8,10,60 In contrast, the positive charge on FA is more declocalized, which can be easily visualized in an electrostatic potential map (Figure 7). This leads to a weaker interaction and a longer hydrogen bond between the FA cation and water, and is likely responsible for the improved moisture resistance of the FA-containing perovskites. Photochemically, the reduced acidity of the FA cation makes it less vulnerable to deprotonation by photogenerated superoxide. Other studies have also shown that the MA cation is vulnerable to deprotonation by basic surface groups,61 while the same reaction does not occur with FAPbI3.62

Figure 7. Electrostatic potential maps of the MA and FA cations, and the optimized geometry of a 1:1 complex with water. All calculations were carried out at the M06/6-31+G(d) level of theory.

4. Conclusion

ACS Paragon Plus Environment

20

Page 21 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

By fabricating a series of perovskite samples with various ion substitutions, we have systematically evaluated how dopants can affect the stability of lead halide perovskites. The FA cation was the most effective at stabilizing perovskite films, and provided increased moisture resistance and photochemical stability. We attribute this to the more delocalized positive charge on the FA cation, which reduces its ability to hydrogen bond with water and lowers its acidity. In addition, by carefully controlling the relevant environmental variables (humidity, temperature, light, and atmosphere), we have been able to separate out their individual contributions to the overall decomposition of lead halide perovskites. Our results suggest that FA-based perovskites are among the most promising candidates for efficient perovskite solar cells with long-term stability.

Supporting Information. Materials, synthetic and device fabrication procedures, pXRD patterns of films and powders, absorbance spectra, AFM images, SEM images, Tauc plots, device moisture resistance and thermal stability data, photoluminescence spectra.

Corresponding Author * E-mail: [email protected]

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

ACS Paragon Plus Environment

21

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 28

Acknowledgements The Natural Sciences and Engineering Research Council of Canada (NSERC) and the University of Saskatchewan are acknowledged for financial support. T.L.K. is a Canada Research Chair in Photovoltaics. The research was undertaken, in part, thanks to funding from the Canada Research Chair program. SEM images were acquired at the Western College of Veterinary Medicine Imaging Centre. Dr. Eiko Kawamura is gratefully acknowledged for assistance with SEM imaging.

References (1) (2) (3) (4)

(5)

(6)

(7)

Ye, M.; Hong, X.; Zhang, F.; Liu, X., Recent Advancements in Perovskite Solar Cells: Flexibility, Stability and Large Scale. J. Mater. Chem. A 2016, 4, 6755-6771. Stranks, S. D.; Snaith, H. J., Metal-Halide Perovskites for Photovoltaic and LightEmitting Devices. Nat. Nanotechnol. 2015, 10, 391-402. Chen, J.; Zhou, S.; Jin, S.; Li, H.; Zhai, T., Crystal Organometal Halide Perovskites with Promising Optoelectronic Applications. J. Mater. Chem. C 2016, 4, 11-27. Wei, H.; Fang, Y.; Mulligan, P.; Chuirazzi, W.; Fang, H.-H.; Wang, C.; Ecker, B. R.; Gao, Y.; Loi, M. A.; Cao, L.; Huang, J., Sensitive X-Ray Detectors Made of Methylammonium Lead Tribromide Perovskite Single Crystals. Nat. Photonics 2016, 10, 333-339. Eperon, G. E.; Stranks, S. D.; Menelaou, C.; Johnston, M. B.; Herz, L. M.; Snaith, H. J., Formamidinium Lead Trihalide: A Broadly Tunable Perovskite for Efficient Planar Heterojunction Solar Cells. Energy Environ. Sci. 2014, 7, 982-988. Yang, J.; Siempelkamp, B. D.; Liu, D.; Kelly, T. L., Investigation of Ch3nh3pbi3 Degradation Rates and Mechanisms in Controlled Humidity Environments Using in Situ Techniques. ACS Nano 2015, 9, 1955-1963. Leguy, A. M. A.; Hu, Y.; Campoy-Quiles, M.; Alonso, M. I.; Weber, O. J.; Azarhoosh, P.; van Schilfgaarde, M.; Weller, M. T.; Bein, T.; Nelson, J.; Docampo, P.; Barnes, P. R. F., Reversible Hydration of Ch3nh3pbi3 in Films, Single Crystals, and Solar Cells. Chem. Mater. 2015, 27, 3397-3407.

ACS Paragon Plus Environment

22

Page 23 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

(8)

(9)

(10)

(11)

(12)

(13)

(14)

(15)

(16)

(17) (18)

(19)

(20)

Bryant, D.; Aristidou, N.; Pont, S.; Sanchez-Molina, I.; Chotchunangatchaval, T.; Wheeler, S.; Durrant, J. R.; Haque, S. A., Light and Oxygen Induced Degradation Limits the Operational Stability of Methylammonium Lead Triiodide Perovskite Solar Cells. Energy Environ. Sci. 2016, 9, 1655-1660. Song, Z.; Abate, A.; Watthage, S. C.; Liyanage, G. K.; Phillips, A. B.; Steiner, U.; Graetzel, M.; Heben, M. J., Perovskite Solar Cell Stability in Humid Air: Partially Reversible Phase Transitions in the Pbi2-Ch3nh3i-H2o System. Adv. Energy Mater. 2016, 6, 1600846. Aristidou, N.; Eames, C.; Sanchez-Molina, I.; Bu, X.; Kosco, J.; Islam, M. S.; Haque, S. A., Fast Oxygen Diffusion and Iodide Defects Mediate Oxygen-Induced Degradation of Perovskite Solar Cells. Nat. Commun. 2017, 8, 15218. Yin, J.; Cao, J.; He, X.; Yuan, S.; Sun, S.; Li, J.; Zheng, N.; Lin, L., Improved Stability of Perovskite Solar Cells in Ambient Air by Controlling the Mesoporous Layer. J. Mater. Chem. A 2015, 3, 16860-16866. Fakharuddin, A.; Di Giacomo, F.; Palma, A. L.; Matteocci, F.; Ahmed, I.; Razza, S.; D’Epifanio, A.; Licoccia, S.; Ismail, J.; Di Carlo, A.; Brown, T. M.; Jose, R., Vertical Tio2 Nanorods as a Medium for Stable and High-Efficiency Perovskite Solar Modules. ACS Nano 2015, 9, 8420-8429. Yang, J.; Fransishyn, K. M.; Kelly, T. L., Comparing the Effect of Mesoporous and Planar Metal Oxides on the Stability of Methylammonium Lead Iodide Thin Films. Chem. Mater. 2016, 28, 7344-7352. Chen, H.; Hou, Y.; Halbig, C. E.; Chen, S.; Zhang, H.; Li, N.; Guo, F.; Tang, X.; Gasparini, N.; Levchuk, I.; Kahmann, S.; Ramirez Quiroz, C. O.; Osvet, A.; Eigler, S.; Brabec, C. J., Extending the Environmental Lifetime of Unpackaged Perovskite Solar Cells through Interfacial Design. J. Mater. Chem. A 2016, 4, 11604-11610. Tan, H.; Jain, A.; Voznyy, O.; Lan, X.; García de Arquer, F. P.; Fan, J. Z.; QuinteroBermudez, R.; Yuan, M.; Zhang, B.; Zhao, Y.; Fan, F.; Li, P.; Quan, L. N.; Zhao, Y.; Lu, Z.-H.; Yang, Z.; Hoogland, S.; Sargent, E. H., Efficient and Stable Solution-Processed Planar Perovskite Solar Cells Via Contact Passivation. Science 2017, 355, 722-726. Lee, J.; Kim, J.; Lee, C.-L.; Kim, G.; Kim, T. K.; Back, H.; Jung, S.; Yu, K.; Hong, S.; Lee, S.; Kim, S.; Jeong, S.; Kang, H.; Lee, K., A Printable Organic Electron Transport Layer for Low-Temperature-Processed, Hysteresis-Free, and Stable Planar Perovskite Solar Cells. Adv. Energy Mater. 2017, 1700226. Zhao, Y.; Wei, J.; Li, H.; Yan, Y.; Zhou, W.; Yu, D.; Zhao, Q., A Polymer Scaffold for Self-Healing Perovskite Solar Cells. Nat. Commun. 2016, 7, 10228. Masi, S.; Rizzo, A.; Munir, R.; Listorti, A.; Giuri, A.; Esposito Corcione, C.; Treat, N. D.; Gigli, G.; Amassian, A.; Stingelin, N.; Colella, S., Organic Gelators as Growth Control Agents for Stable and Reproducible Hybrid Perovskite-Based Solar Cells. Adv. Energy Mater. 2017, 1602600. Lu, J.; Jiang, L.; Li, W.; Li, F.; Pai, N. K.; Scully, A. D.; Tsai, C.-M.; Bach, U.; Simonov, A. N.; Cheng, Y.-B.; Spiccia, L., Diammonium and Monoammonium Mixed-OrganicCation Perovskites for High Performance Solar Cells with Improved Stability. Adv. Energy Mater. 2017, 1700444. Long, M.; Zhang, T.; Chai, Y.; Ng, C.-F.; Mak, T. C. W.; Xu, J.; Yan, K., Nonstoichiometric Acid–Base Reaction as Reliable Synthetic Route to Highly Stable Ch3nh3pbi3 Perovskite Film. Nat. Commun. 2016, 7, 13503.

ACS Paragon Plus Environment

23

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(21)

(22)

(23)

(24)

(25)

(26)

(27)

(28)

(29)

(30)

(31)

(32)

Page 24 of 28

Bi, D.; Yi, C.; Luo, J.; Décoppet, J.-D.; Zhang, F.; Zakeeruddin, Shaik M.; Li, X.; Hagfeldt, A.; Grätzel, M., Polymer-Templated Nucleation and Crystal Growth of Perovskite Films for Solar Cells with Efficiency Greater Than 21%. Nat. Energy 2016, 1, 16142. Qin, C.; Matsushima, T.; Fujihara, T.; Adachi, C., Multifunctional Benzoquinone Additive for Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2017, 29, 1083808. Sun, Y.; Wu, Y.; Fang, X.; Xu, L.; Ma, Z.; Lu, Y.; Zhang, W.-H.; Yu, Q.; Yuan, N.; Ding, J., Long-Term Stability of Organic-Inorganic Hybrid Perovskite Solar Cells with High Efficiency under High Humidity Conditions. J. Mater. Chem. A 2017, 5, 13741379. Tsai, H.; Nie, W.; Blancon, J.-C.; Stoumpos, C. C.; Asadpour, R.; Harutyunyan, B.; Neukirch, A. J.; Verduzco, R.; Crochet, J. J.; Tretiak, S.; Pedesseau, L.; Even, J.; Alam, M. A.; Gupta, G.; Lou, J.; Ajayan, P. M.; Bedzyk, M. J.; Kanatzidis, M. G.; Mohite, A. D., High-Efficiency Two-Dimensional Ruddlesden–Popper Perovskite Solar Cells. Nature 2016, 536, 312-316. Yang, M.; Zhang, T.; Schulz, P.; Li, Z.; Li, G.; Kim, D. H.; Guo, N.; Berry, J. J.; Zhu, K.; Zhao, Y., Facile Fabrication of Large-Grain Ch3nh3pbi3−Xbrx Films for HighEfficiency Solar Cells Via Ch3nh3br-Selective Ostwald Ripening. Nat. Commun. 2016, 7, 12305. Saliba, M.; Matsui, T.; Domanski, K.; Seo, J.-Y.; Ummadisingu, A.; Zakeeruddin, S. M.; Correa-Baena, J.-P.; Tress, W. R.; Abate, A.; Hagfeldt, A.; Grätzel, M., Incorporation of Rubidium Cations into Perovskite Solar Cells Improves Photovoltaic Performance. Science 2016, 354, 206-209. Eperon, G. E.; Leijtens, T.; Bush, K. A.; Prasanna, R.; Green, T.; Wang, J. T.-W.; McMeekin, D. P.; Volonakis, G.; Milot, R. L.; May, R.; Palmstrom, A.; Slotcavage, D. J.; Belisle, R. A.; Patel, J. B.; Parrott, E. S.; Sutton, R. J.; Ma, W.; Moghadam, F.; Conings, B.; Babayigit, A.; Boyen, H.-G.; Bent, S.; Giustino, F.; Herz, L. M.; Johnston, M. B.; McGehee, M. D.; Snaith, H. J., Perovskite-Perovskite Tandem Photovoltaics with Optimized Bandgaps. Science 2016, 354, 861-865. Yang, Z.; Rajagopal, A.; Chueh, C.-C.; Jo, S. B.; Liu, B.; Zhao, T.; Jen, A. K. Y., Stable Low-Bandgap Pb–Sn Binary Perovskites for Tandem Solar Cells. Adv. Mater. 2016, 28, 8990-8997. Wang, F.; Geng, W.; Zhou, Y.; Fang, H.-H.; Tong, C.-J.; Loi, M. A.; Liu, L.-M.; Zhao, N., Phenylalkylamine Passivation of Organolead Halide Perovskites Enabling HighEfficiency and Air-Stable Photovoltaic Cells. Adv. Mater. 2016, 28, 9986-9992. Li, N.; Zhu, Z.; Chueh, C.-C.; Liu, H.; Peng, B.; Petrone, A.; Li, X.; Wang, L.; Jen, A. K. Y., Mixed Cation Faxpea1–Xpbi3 with Enhanced Phase and Ambient Stability toward High-Performance Perovskite Solar Cells. Adv. Energy Mater. 2017, 7, 1601307. Liang, J.; Wang, C.; Wang, Y.; Xu, Z.; Lu, Z.; Ma, Y.; Zhu, H.; Hu, Y.; Xiao, C.; Yi, X.; Zhu, G.; Lv, H.; Ma, L.; Chen, T.; Tie, Z.; Jin, Z.; Liu, J., All-Inorganic Perovskite Solar Cells. J. Am. Chem. Soc. 2016, 138, 15829-15832. Saliba, M.; Matsui, T.; Seo, J.-Y.; Domanski, K.; Correa-Baena, J.-P.; Nazeeruddin, M. K.; Zakeeruddin, S. M.; Tress, W.; Abate, A.; Hagfeldt, A.; Gratzel, M., CesiumContaining Triple Cation Perovskite Solar Cells: Improved Stability, Reproducibility and High Efficiency. Energy Environ. Sci. 2016, 9, 1989-1997.

ACS Paragon Plus Environment

24

Page 25 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

(33)

(34)

(35)

(36)

(37)

(38)

(39)

(40)

(41)

(42)

(43)

(44)

(45)

(46)

Bi, D.; Gao, P.; Scopelliti, R.; Oveisi, E.; Luo, J.; Grätzel, M.; Hagfeldt, A.; Nazeeruddin, M. K., High-Performance Perovskite Solar Cells with Enhanced Environmental Stability Based on Amphiphile-Modified Ch3nh3pbi3. Adv. Mater. 2016, 28, 2910-2915. Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G., Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and nearInfrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019-9038. Im, J.-H.; Chung, J.; Kim, S.-J.; Park, N.-G., Synthesis, Structure, and Photovoltaic Property of a Nanocrystalline 2h Perovskite-Type Novel Sensitizer (Ch3ch2nh3)Pbi3. Nanoscale Res. Lett. 2012, 7, 1-7. Gholipour, S.; Ali, A. M.; Correa-Baena, J.-P.; Turren-Cruz, S.-H.; Tajabadi, F.; Tress, W.; Taghavinia, N.; Grätzel, M.; Abate, A.; De Angelis, F.; Gaggioli, C. A.; Mosconi, E.; Hagfeldt, A.; Saliba, M., Globularity-Selected Large Molecules for a New Generation of Multication Perovskites. Adv. Mater. 2017, 10.1002/adma.201702005, 1702005. Hsu, H.-L.; Chang, C.-C.; Chen, C.-P.; Jiang, B.-H.; Jeng, R.-J.; Cheng, C.-H., HighPerformance and High-Durability Perovskite Photovoltaic Devices Prepared Using Ethylammonium Iodide as an Additive. J. Mater. Chem. A 2015, 3, 9271-9277. 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. Chen, Q.; Zhou, H.; Fang, Y.; Stieg, A. Z.; Song, T.-B.; Wang, H.-H.; Xu, X.; Liu, Y.; Lu, S.; You, J.; Sun, P.; McKay, J.; Goorsky, M. S.; Yang, Y., The Optoelectronic Role of Chlorine in Ch3nh3pbi3(Cl)-Based Perovskite Solar Cells. Nat. Commun. 2015, 6, 7269. 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. Jeon, N. J.; Noh, J. H.; Yang, W. S.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I., Compositional Engineering of Perovskite Materials for High-Performance Solar Cells. Nature 2015, 517, 476-480. Binek, A.; Hanusch, F. C.; Docampo, P.; Bein, T., Stabilization of the Trigonal HighTemperature Phase of Formamidinium Lead Iodide. J. Phys. Chem. Lett. 2015, 6, 12491253. Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I., HighPerformance Photovoltaic Perovskite Layers Fabricated through Intramolecular Exchange. Science 2015, 348, 1234-1237. Jesper Jacobsson, T.; Correa-Baena, J.-P.; Pazoki, M.; Saliba, M.; Schenk, K.; Gratzel, M.; Hagfeldt, A., Exploration of the Compositional Space for Mixed Lead Halogen Perovskites for High Efficiency Solar Cells. Energy Environ. Sci. 2016, 10.1039/c6ee00030d. Liu, Y.; Yang, Z.; Cui, D.; Ren, X.; Sun, J.; Liu, X.; Zhang, J.; Wei, Q.; Fan, H.; Yu, F.; Zhang, X.; Zhao, C.; Liu, S., Two-Inch-Sized Perovskite Ch3nh3pbx3 (X = Cl, Br, I) Crystals: Growth and Characterization. Adv. Mater. 2015, 27, 5176-5183. Zhumekenov, A. A.; Saidaminov, M. I.; Haque, M. A.; Alarousu, E.; Sarmah, S. P.; Murali, B.; Dursun, I.; Miao, X.-H.; Abdelhady, A. L.; Wu, T.; Mohammed, O. F.; Bakr, O. M., Formamidinium Lead Halide Perovskite Crystals with Unprecedented Long Carrier Dynamics and Diffusion Length. ACS Energy Lett. 2016, 1, 32-37.

ACS Paragon Plus Environment

25

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(47)

(48)

(49) (50)

(51)

(52)

(53)

(54)

(55)

(56)

(57)

(58)

(59)

(60)

Page 26 of 28

Tian, Y.; Scheblykin, I. G., Artifacts in Absorption Measurements of Organometal Halide Perovskite Materials: What Are the Real Spectra? J. Phys. Chem. Lett. 2015, 6, 34663470. Gangishetty, M. K.; Scott, R. W. J.; Kelly, T. L., Effect of Relative Humidity on Crystal Growth, Device Performance and Hysteresis in Planar Heterojunction Perovskite Solar Cells. Nanoscale 2016, 8, 6300-6307. Liu, D.; Yang, J.; Kelly, T. L., Compact Layer Free Perovskite Solar Cells with 13.5% Efficiency. J. Am. Chem. Soc. 2014, 136, 17116-17122. 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. Li, J.; Dong, Q.; Li, N.; Wang, L., Direct Evidence of Ion Diffusion for the SilverElectrode-Induced Thermal Degradation of Inverted Perovskite Solar Cells. Adv. Energy Mater. 2017, 10.1002/aenm.201602922, 1602922-n/a. Back, H.; Kim, G.; Kim, J.; Kong, J.; Kim, T. K.; Kang, H.; Kim, H.; Lee, J.; Lee, S.; Lee, K., Achieving Long-Term Stable Perovskite Solar Cells Via Ion Neutralization. Energy Environ. Sci. 2016, 9, 1258-1263. Han, Y.; Meyer, S.; Dkhissi, Y.; Weber, K.; Pringle, J. M.; Bach, U.; Spiccia, L.; Cheng, Y.-B., Degradation Observations of Encapsulated Planar Ch3nh3pbi3 Perovskite Solar Cells at High Temperatures and Humidity. J. Mater. Chem. A 2015, 3, 8139-8147. Deng, Y.; Dong, Q.; Bi, C.; Yuan, Y.; Huang, J., Air-Stable, Efficient Mixed-Cation Perovskite Solar Cells with Cu Electrode by Scalable Fabrication of Active Layer. Adv. Energy Mater. 2016, 6, 1600372. Noel, N. K.; Abate, A.; Stranks, S. D.; Parrott, E. S.; Burlakov, V. M.; Goriely, A.; Snaith, H. J., Enhanced Photoluminescence and Solar Cell Performance Via Lewis Base Passivation of Organic–Inorganic Lead Halide Perovskites. ACS Nano 2014, 8, 98159821. Jacobsson, T. J.; Correa-Baena, J.-P.; Halvani Anaraki, E.; Philippe, B.; Stranks, S. D.; Bouduban, M. E. F.; Tress, W.; Schenk, K.; Teuscher, J.; Moser, J.-E.; Rensmo, H.; Hagfeldt, A., Unreacted Pbi2 as a Double-Edged Sword for Enhancing the Performance of Perovskite Solar Cells. J. Am. Chem. Soc. 2016, 138, 10331-10343. Govindhasamy, M.; Soichiro, T.; Gai, M.; Shusaku, K.; Hitoshi, N.; Tomokazu, U.; Hiroshi, I.; Seigo, I., Light Stability Tests of Methylammonium and Formamidinium PbHalide Perovskites for Solar Cell Applications. Jpn. J. Appl. Phys. 2015, 54, 08KF08. Lee, J.-W.; Kim, D.-H.; Kim, H.-S.; Seo, S.-W.; Cho, S. M.; Park, N.-G., Formamidinium and Cesium Hybridization for Photo- and Moisture-Stable Perovskite Solar Cell. Adv. Energy Mater. 2015, 5, 1501310. Vincent, B. R.; Robertson, K. N.; Cameron, T. S.; Knop, O., Alkylammonium Lead Halides. Part 1. Isolated Pbi64− Ions in (Ch3nh3)4pbi6•2h2o. Can. J. Chem. 1987, 65, 1042-1046. Aristidou, N.; Sanchez-Molina, I.; Chotchuangchutchaval, T.; Brown, M.; Martinez, L.; Rath, T.; Haque, S. A., The Role of Oxygen in the Degradation of Methylammonium Lead Trihalide Perovskite Photoactive Layers. Angew. Chem. Int. Ed. 2015, 54, 82088212.

ACS Paragon Plus Environment

26

Page 27 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

(61)

(62)

(63)

Yang, J.; Siempelkamp, B. D.; Mosconi, E.; De Angelis, F.; Kelly, T. L., Origin of the Thermal Instability in Ch3nh3pbi3 Thin Films Deposited on Zno. Chem. Mater. 2015, 27, 4229-4236. Song, J.; Hu, W.; Wang, X.-F.; Chen, G.; Tian, W.; Miyasaka, T., Hc(Nh2)2pbi3 as a Thermally Stable Absorber for Efficient Zno-Based Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 8435-8443. Xiao, Z.; Dong, Q.; Bi, C.; Shao, Y.; Yuan, Y.; Huang, J., Solvent Annealing of Perovskite-Induced Crystal Growth for Photovoltaic-Device Efficiency Enhancement. Adv. Mater. 2014, 26, 6503-6509.

ACS Paragon Plus Environment

27

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 28

ToC Figure

ACS Paragon Plus Environment

28