You are on page 1of 15

Subscriber access provided by READING UNIV

Letter
Modeling the Interaction of Molecular Iodine with MAPbI3:
A Probe of Lead-Halide Perovskites Defect Chemistry
Daniele Meggiolaro, Edoardo Mosconi, and Filippo De Angelis
ACS Energy Lett., Just Accepted Manuscript • DOI: 10.1021/acsenergylett.7b01244 • Publication Date (Web): 18 Jan 2018
Downloaded from http://pubs.acs.org on January 18, 2018

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 Energy Letters 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 14 ACS Energy Letters

1
2
3 Modeling the Interaction of Molecular Iodine with MAPbI3: A
4
5
6
7
Probe of Lead-Halide Perovskites Defect Chemistry
8
9
10 Daniele Meggiolaro,a,b Edoardo Mosconi,a,c * Filippo De Angelis a,b,c *
11
12
13
14
15
a
16 Computational Laboratory for Hybrid/Organic Photovoltaics (CLHYO), CNR-ISTM, Via Elce di
17
18 Sotto 8, 06123, Perugia, Italy.
19
20 b
D3-CompuNet, Istituto Italiano di Tecnologia, Via Morego 30, 16163 Genova, Italy.
21
22
23 c
24 Consortium for Computational Molecular and Materials Sciences (CMS)2, Via Elce di Sotto, 8,
25
26 06123 Perugia, Italy.
27
28
29
30
31
32
33
34 TOC graphics
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 1
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 2 of 14

1
2
3 Abstract
4
5
6 Understanding the defect chemistry of lead-halide perovskites is of paramount importance to further
7
8 progress towards exploitation of these materials. Here we combine recent experimental observations
9
10 on the behavior of MAPbI3 upon exposure to I2 vapor with first-principles calculations to extract a
11
12 global picture of defect chemistry in lead-halide perovskites. By matching the reported
13
14 experimental observables we disclose the origin of the p-doping observed upon exposing MAPbI3
15
16
to I2 and highlight its consequences on the charge/ion transport and trapping activity. Electron/hole
17
18
19 traps related to positive/negative interstitial iodine dominate the defect chemistry in intrinsic
20
21 conditions, while in p-doped MAPbI3 electrons are mainly trapped by positive interstitial iodine and
22
23 neutral lead vacancies. I2 spontaneously dissociates on iodine vacancies, leading to vacancy
24
25 passivation and to the formation of positive interstitial iodine. I2 spontaneously dissociates on non-
26
27 defective MAPbI3 (001) surfaces to form pairs of negative/positive interstitial iodine. Upon trapping
28
29 a hole/electron pair at negative/positive interstitial iodine, I2 release becomes thermodynamically
30
31
32
favored possibly representing a photo-induced trap curing mechanism.
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 2
59
60 ACS Paragon Plus Environment
Page 3 of 14 ACS Energy Letters

1
2
3 Understanding the defect chemistry of lead-halide perovskites is of paramount importance to further
4
5 progress with the practical solar cell exploitation of this outstanding class of materials. The long
6
7 lifetimes and diffusion lengths of photogenerated charge carriers in MAPbI3 and related perovskites
8
9 are suggestive of an apparently low density of charge traps. Clearly, a distinction should be made
10
11
between defects and traps, since most of typical defects in conventional perovskites turn out to be
12
13
14 only shallow traps in MAPbI3.1-4 This, coupled to large polarons protecting photogenerated charge
15
16 carriers,5, 6 has generated the concept of the what is generically referred to as “perovskite defect
17
18 tolerance”. One must recognize that defect tolerance, in the general meaning, is a fundamental
19
20 ingredient of the astonishing properties of this materials class, which would otherwise probably
21
22 only have a mere academic interest.
23
24
Most experimental investigations provide indirect hints on defects, mainly determining the
25
26
27 density of trap states through electrical conductivity measurements. Very low trap densities of
28
29 ∼1010-1011 cm-3 have been reported for MAPbI3 single crystals,7, 8 increasing up to ∼1016 cm-3 for
30
31 polycrystalline thin films.8-10 Disclosing the properties of defects in a semiconductor with atomistic
32
33 resolution can be accessed through first-principles calculations, which are however limited by the
34
35 scale of employed models and the intrinsic accuracy the underlying electronic structure theory. Here
36
37
38
we sought to combine a survey of recent experimental observations on the behavior of MAPbI3
39
40 upon exposure to molecular iodine vapor with high level first-principles calculations to extract a
41
42 global picture of defect chemistry in lead-halide perovskites. We thus use the effect of I2 exposure
43
44 to MAPbI3 as a probe of the underlying material defect chemistry. By matching the reported
45
46 experimental observables to our model defect chemistry, we disclose the origin of the observed p-
47
48 doping upon I2 exposure and highlight its consequences on the charge/ion transport and trapping
49
50
activity.
51
52
53 To our knowledge, four studies have reported on the interaction between I2 vapors and
54
55 MAPbI3.11-14 Here we briefly summarize the main findings of these studies which are relevant to the
56
57 present work:
58 3
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 4 of 14

1
2
3 1) Formation of metallic lead, as revealed by XPS, related to degradation of the perovskite.11
4
5 2) 150 mV shift of the Fermi energy towards the valence band, indicative of p-doping.13
6
7 3) Irreversibility of the p-doping.11, 13
8
9 4) Quenching of the photoluminescence quantum yield.12
10
11
5) Increase of electrical conductivity due to holes. 14
12
13
14 6) Decrease of the ionic conductivity.14
15
16
17
18 We begin by re-investigating native defects in MAPbI3 by state of the art first-principles
19
20 calculations in 2x2x1 tetragonal supercells using hybrid DFT, including dispersion corrections and
21
22 spin-orbit coupling (SOC). We use the modified HSE06 exchange-correlation functional including
23
24
43% exact exchange proposed in Ref. 15, which provides band edge energetics comparable to high-
25
26
27 level GW-SOC calculations.16 Hybrid functionals are important to characterize defects with
28
29 unpaired electrons,17, 18 such as the neutral Ii0 and VI0 defects. The central quantity to our discussion
30
31 is the Defect Formation Energy (DFE), see Supplementary Information. Knowledge of the DFE of
32
33 relevant defects in different charge states allows one to calculate: i) the thermodynamic ionization
34
35 levels, correspond to oxidation and reduction potentials of the defective system, whose energy with
36
37 respect to the band edges determine the defect trapping activity; ii) the defect density; and iii) the
38
39
40
native Fermi level of the system.19
41
42 The DFE diagram of MAPbI3 grown in iodine medium (corresponding to 1:1 ratio of the
43
44 PbI2 and MAI precursors) and excess I2 vapors are reported in Figure 1a and 1b, respectively; the
45
46 associated thermodynamic ionization levels are reported in Figure 1c.
47
48
49
50
51
52
53
54
55
56
57
58 4
59
60 ACS Paragon Plus Environment
Page 5 of 14 ACS Energy Letters

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 1. DFEs diagrams calculated by SOC-HSE06 in iodine medium conditions (a) and under a
24
25 simulated I2 pressure of 10-4 atm (b); c) Thermodynamic transition levels, which are independent of
26
27 the growth conditions.
28
29
30 In iodine-medium conditions the most stable defects are MA interstitials (MAi+), negative
31
32 interstitial iodine (Ii-) and lead vacancies (VPb2-), which pin the Fermi level at mid-gap (0.71 eV
33
34 above VBM over a 1.58 eV calculated band gap), Figure 1a. Positive interstitial iodine (Ii+) and
35
36
37
iodine vacancy (VI+) have a ∼0.2 eV higher DFE, while interstitial lead (Pbi2+) lies at much higher
38
39 DFE. The close to mid-gap native Fermi level is consistent with the intrinsic (or mildly p-doped)
40
41 nature of MAPbI3,13, 20 while thin films show n-doping related to a Pb-rich surface environment.21
42
43 To properly account for defect activity, the defect activation energy to migration should be taken
44
45 into account, since the presence of defects with high activation energy could be kinetically limited.
46
47
Defects related to I and MA ions are the major migrating species,22-24 with I faster than MA, faster
48
49
50 than Pb migration.25 Recent X-ray and neutron diffraction measurements on MAPbI3 revealed
51
52 significant structural disorder associated to interstitial iodine in MAPbI3 films.26
53
54
55
56
57
58 5
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 6 of 14

1
2
3 Table 1. DFEs and defect densities calculated in iodine-medium and at I2 partial pressure of 10-4
4
5 atm.
6
7
8 iodine-medium P(I2) = 10-4 atm
9
10
11 Defect DFE (eV) Density (cm-3) DFE (eV) Density (cm-3)
12
13
14 VPb2- 0.66 ∼ 1010 0.44 ∼ 1014
15
16
17 VPb0 1.10 ∼ 103 0.71 ∼ 1010
18
19
20 MAi+ 0.57 ∼ 1012 0.69 ∼ 1010
21
22
23 Ii+ 0.72 ∼ 1010 0.44 ∼ 1014
24
25
26
Ii- 0.60 ∼ 1012 0.49 ∼ 1014
27
28
29
30 VI + 0.87 ∼ 107 0.98 ∼ 106
31
32
33 Pbi2+ 1.16 ∼ 102 1.38 ∼0
34
35
36
37
38
39
40 Interstitial iodine is an amphoteric defect that can trap both electrons and holes with +/0 and
41
42 0/- ionization levels placed at 0.57 and 0.30 eV below and above the CB and VB edges, Figure 1c.
43
44 Notice that the neutral state of interstitial iodine is unstable, 15 similar to the typical iodine chemistry
45
46 whereby I2- (or a neutral I atom) is unstable with respect to I- and I3-.27 The metastable nature of Ii0
47
48
implies that Ii- should trap two holes to convert to Ii+, with the (-/+) transition placed at mid-gap; the
49
50
51 two-electron nature of the process, however, imparts it a vanishingly small cross section since the
52
53 probability of two consecutive trapping events at the same defect site is vanishingly small. VI+ has a
54
55 transition level resonant with the CB, thus although showing a low activation energy to migration,28
56
57 it can at most represent a shallow trapping site in bulk crystals. VPb is stable in the 2- charge state in
58 6
59
60 ACS Paragon Plus Environment
Page 7 of 14 ACS Energy Letters

1
2
3 a wide range of Fermi energies, thus this defect can trap holes by the (-/2-) transition level whose
4
5 energy falls 0.15 eV above the VBM, suggesting a shallow trapping activity. The (0/2-) transition
6
7 implies the trapping of two holes on the same defect site, and as such it has a vanishingly small
8
9 capture cross section. The deep (0/-) transition falling at mid-gap is not active in iodine-medium
10
11
conditions due to the low density of VPb0, see Table 1. Notably, VPb0 decomposes into VPb2- + Ii+ +
12
13
14 VI+ on the same site, which is more stable than the simple neutral Pb vacancy by 0.40 eV. The (0/-)
15
16 ionization level of VPb is thus the analogous of the (+/0) transition of Ii, falling approximately in
17
18 the same energy range. As previously mentioned, lead vacancies have a large (∼1 eV) migration
19
20 barrier,28 thus their formation could be kinetically limited. In iodine-medium conditions the main
21
22
deep electron and hole traps among native defects are thus related to interstitial iodine. The match
23
24
25
between the calculated Ii+/- density (~1010-1012 cm-3, see Table 1) and the measured trap density
26
27 points at Ii+ and Ii- as the experimentally observed electron and hole traps in MAPbI3 single
28
29 crystals.7
30
31
32 By increasing the iodine partial pressure, mimicking experiments in which the perovskite is
33
34 exposed to I2 vapors, we are shifting the equilibrium towards iodine-rich conditions. To quantify the
35
36 chemical potential shift due to exposure of MAPbI3 to I2 pressures of the order of ∼10-4 atm
37
38
employed in experiments,13, 14
we calculated the variation of I2 chemical potential induced by
39
40
41 pressure referred to iodine rich conditions as:
42
43
44 ∆µ= -RT ln(P/P0) -T∆St (2)
45
46
47 where P is the actual pressure referred to the standard P0=1 atm, T is the absolute temperature (300
48
49 K here), and ∆St is the variation of translation entropy of I2 gas with pressure estimated assuming an
50
51 ideal gas behavior, see Table S1, Supporting Information. We verified that other terms in the Gibbs
52
53 free energy (e.g. thermal contributions to enthalpy due to vibrations, and rotational-vibrational
54
55
contributions to entropy) remain roughly constant in the explored pressure variation range. Notice
56
57
58 7
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 8 of 14

1
2
3 that variation of the chemical potential of a gaseous species in a sizable range (0.2 eV or more) is
4
5 easily achieved by varying its partial pressure, while tuning the chemical potential of a species in
6
7 solution by a factor ∼2.3 RT (∼0.06 eV) implies variation of its activity (concentration) by 1 order
8
9
of magnitude, a situation which can be hardly realized in standard perovskite synthetic conditions.
10
11
12 Upon exposing MAPbI3 to I2 the most notable effect is the stabilization of positive
13
14
15
interstitial iodine and lead vacancies, Figure 1b, with a consequent p-doping of the material.
16
17 Accordingly, the calculated Fermi level shifts by 80 meV, from 0.71 to 0.63 eV above the VBM,
18
19 consistent with the 150 mV work function increase measured by Zohar et al.13. The increase in
20
21 electric conductivity associated to hole transport observed by Senocrate et al.14 is also readily
22
23 explained since holes are the dominant free carriers under p-doped conditions and their
24
25 concentration is predicted to increase by ~3 orders of magnitudes a consequence of the observed p-
26
27
28
doping. Our results are also consistent with the formation of metallic lead observed by Wang et al.
29
30 upon exposure of MAPbI3 to I2,11 related to the stabilization of lead vacancies, i.e. the expulsion of
31
32 lead from its bulk crystal site. Under p-doped conditions the main charge trap is now Ii+ (and to a
33
34 lesser extent VPb0) whose density increases by several orders of magnitude compared to intrinsic
35
36 conditions, see Table 1. Both defects can effectively trap electrons, thereby explaining the
37
38 photoluminescence quenching observed by Li et al. after MAPbI3 interaction with I2.12
39
40
41 Our model has so far allowed us to rationalize most of the experimental observations related
42
43
to I2 exposure, with the exception of the decreased ionic conductivity measured by Senocrate et al.14
44
45
46 These authors ascribed such behavior to the saturation of iodine vacancies, implying a mechanism
47
48 of iodine incorporation mediated by vacancies. This seems a perfectly reasonable explanation, since
49
29
50 iodine vacancy sites would “attract” mobile iodine atoms from I2 to restore a perfect crystal
51
52 lattice. Ion conductivity is a function of both the density of ionic migrating species and their
53
54 migration energy barrier. A decrease in ionic conductivity can be either caused by a reduction of the
55
56
density of migrating defects (e.g. iodine vacancies) or by an increase of the migration energy
57
58 8
59
60 ACS Paragon Plus Environment
Page 9 of 14 ACS Energy Letters

1
2
3 barrier, or both. In this respect, one should notice that iodine vacancies are stable in their positive
4
5 charge state in all conditions, meaning that to preserve electroneutrality I2 dissociation in a vacancy
6
7 should also imply formation of a positive interstitial iodine, as:
8
9
10 VI+ + I2→perf + Ii+ (3)
11
12
13 where perf means the perfect crystal. We first calculated the thermodynamics for reaction (3) in a
14
15 tetragonal 2x2x2 bulk MAPbI3 supercell, finding it favorable by 0.39 (0.72) eV with no corrections
16
17
(including dispersion and entropic corrections, see Computational Details). Thus vacancy filling by
18
19
20 I2 is in all cases a thermodynamically favored reaction which produces Ii+. The question is now
21
22 whether positive interstitial iodine has a higher migration energy barrier than a positive iodine
23
24 vacancy. To check this point we searched for the minimum energy pathway to Ii+ migration, by
25
26 scanning the potential energy surface along a displacement coordinate, followed by short ab initio
27
28 molecular dynamics simulations to properly allow for local atomic rearrangement.22 This procedure
29
30
is expected to deliver lower activation energies than standard calculations by virtue of the effective
31
32
33 relaxation around the migration pathway. In Table 2 we report migration energy barriers for
34
35 relevant defects calculated at the SR-PBE level used for geometry optimizations and refined by
36
37 SOC-HSE06.
38
39
40 Table 2. Migration energy barriers (eV) calculated by SR-PBE and SOC-HSE06 levels of theory on
41
42 the SR-PBE optimized geometries.
43
44
45 SR-PBE SOC-HSE06
46
47
48 VI + 0.09 0.11
49
50
51 Ii+ 0.19 0.24
52
53
54 Ii- 0.08 0.06
55
56
57
58 9
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 10 of 14

1
2
3 As it can be noticed, SR-PBE and SOC-HSE06 levels of theory provide similar activation energies.
4
5 Notably, while VI+ and Ii- have roughly the same energy barrier, we predict a factor 2 increase in the
6
7 migration energy barrier for Ii+ compared to VI+, from ∼0.1 to ∼0.2 eV, consistent with the decrease
8
9 14
of ionic conductivity experimentally measured in Ref. . This points at saturation of iodine
10
11
12 vacancies by I2 leaving behind an Ii+ defect, further contributing to p-doping and
13
14 photoluminescence quenching.
15
16
17 Besides the vacancy saturation pathway, I2 could possibly interact with both the intact
18
19 perovskite bulk and surfaces. A recent computational study reported the favorable interaction of I2
20
21 with MAPbI3 surfaces, leading to split iodine interstitials.30 To investigate the energetics and
22
23 structural features of this processes in the bulk and on surfaces, we explicitly added one I2 molecule
24
25 to a bulk 2x2x2 MAPbI3 tetragonal supercell and to MAI-terminated 2x2x3 slabs exposing (001)
26
27
28
and (110) facets, see Figure 2 and Supporting Information. In these cases the MA cations were
29
30 initially aligned antiparallel to quench the molecular dipole. Most notably, I2 is not stable in its
31
32 molecular form and spontaneously dissociates into a positive and a negative interstitial iodine (Ii+
33
34 and Ii-), similarly to what found by Zhang et al. 30, being stabilized by the bond with I and Pb atoms,
35
36 respectively, see Figure 2.
37
38
39
40
41
42
43
44
45
46
47
48
49
50 Figure 2. a) Starting bulk MAPbI3 structure with added I2 (yellow atoms) and b) final relaxed
51
52
53
structures. Notice in b) the formation of the typical Ii+ trimer and Ii- dimer. c) Optimized structure
54
55 of I2 added at the MAI-terminated MAPbI3 surface (001) facet, delivering the same Ii+/ Ii- pair. I-I
56
57 distances (Å) are also reported.
58 10
59
60 ACS Paragon Plus Environment
Page 11 of 14 ACS Energy Letters

1
2
3 I2 incorporation into bulk MAPbI3 is unfavorable (favorable) by 0.40 (0.02) eV with no corrections
4
5 (including dispersion and entropic corrections), so we expect the vacancy-mediated I2 incorporation
6
7 mechanism to dominate the bulk chemistry. The Ii+ / Ii- intimate defect couple is more stable than
8
9 the non-interacting Ii+ and Ii- defects by ~0.15 eV. This means that formation of interacting Ii+ and
10
11
Ii- defects, which is possible due to their low barriers to migration, could increase their density in
12
13
14 bulk MAPbI3.
15
16
17
I2 dissociation becomes thermodynamically favorable by 0.73 (0.16) eV on MAPbI3 (001)
18
19 facets with no corrections (with corrections) forming the same Ii+ / Ii- interacting defect couple, see
20
21 Figure 2c.30 Notably, I2 dissociation is unfavorable by 0.10 eV on the (110) facet, see Supporting
22
23 Information, suggesting a different reactivity of the various MAPbI3 surfaces. I2 dissociation at
24
25 surfaces could thus still represent a pathway of VI+ saturation. The calculated thermodynamics is
26
27 consistent with the irreversibility of the p-doping observed in Ref. 12
. This is at variance with
28
29
30
MAPbI3 interaction with O2, which is characterized by a much smaller interaction energy,
31
32 accounting for the reversibility of the reaction.31
33
34
35 Finally, it is interesting to notice that the energetics of I2 dissociation and incorporation at
36
37 surfaces can be reverted when introducing an electron/hole pair into the system. By simulating the
38
39 lowest triplet state of the MAPbI3 (001) surface with the Ii+ / Ii- interacting defect couple, we find
40
41 that destabilization of neutral interstitial iodine by electron/hole trapping at Ii+ / Ii-, respectively,
42
43 leads to spontaneous I2 release from the surface, which is now favorable by 0.19 (0.74) eV with no
44
45
corrections (with corrections). This I2 formation pathway following electron/hole capture may
46
47
48 constitute a source of light-induced trap annihilation, analogously to the previously proposed
49
50 Frenkel (vacancy/interstitial) mediated pathway.28 The main difference between the two
51
52 mechanisms is that since I2 would remain close to the perovskite surface, unless it is evacuated as a
53
54 gas, it may reversibly re-dissociate to reform the electron/hole Ii+ / Ii- trapping defects. In a typical
55
56 experiment in which MAPbI3 is exposed to inert atmosphere (by e.g. N2) while irradiating the
57
58 11
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 12 of 14

1
2
3 sample, we thus predict that increasing (decreasing) the pressure of N2 will lead to a decrease
4
5 (increase) of the effectiveness of light-induced trap annihilation, by disfavoring (favoring) the exit
6
7 of gaseous I2. At the same time, removing the formed I2 by e.g. fluxing N2 could lead to more
8
9 effective and partly irreversible decrease in the material trap density.
10
11
12 In summary, we have constructed a global picture of defect chemistry in lead-iodide
13
14
perovskites. Under intrinsic conditions, i.e. when the Fermi level is close to mid-gap, positive and
15
16
17
negative interstitial iodine constitute the main source of electron and hole traps, respectively. Upon
18
19 exposing the material to I2 a Fermi level down shift is calculated, which closely matches the
20
21 experimental work function variation. I2 dissociates on iodine vacancies restoring the pristine
22
23 crystal but leaving behind positive interstitial iodine. The higher migration energy barrier of
24
25 positive interstitial iodine compared to iodine vacancy accounts for the decreased ionic transport,
26
27 while the associated p-doping increases the hole electrical conductivity. Notably, I2 dissociates both
28
29
30
in the bulk and on surfaces to produce a pair of positive/negative interstitial iodine defects, with the
31
32 reaction being strongly favored on (001) surfaces. Upon light irradiation, a hole and an electron can
33
34 be trapped at the positive/negative interstitial iodine defect pair, leading to destabilization of the
35
36 defect pair and I2 exit from the perovskite. This may represent an additional pathway of light-
37
38 induced defect annihilation.
39
40
41 Supporting Information Available: Computational Details, (110) structures, I2 chemical
42
43 potentials.
44
45
46 References:
47
48 (1) Yin, W.-J.; Shi, T.; Yan, Y. Unusual defect physics in CH3NH3PbI3 perovskite solar cell
49 absorber. Appl. Phys. Lett. 2014, 104, 063903.
50
(2) Buin, A.; Comin, R.; Xu, J.; Ip, A. H.; Sargent, E. H. Halide-Dependent Electronic Structure of
51
Organolead Perovskite Materials. Chem. Mater. 2015, 27, 4405-4412.
52
53 (3) Agiorgousis, M. L.; Sun, Y.-Y.; Zeng, H.; Zhang, S. Strong Covalency-Induced Recombination
54 Centers in Perovskite Solar Cell Material CH3NH3PbI3. J. Am. Chem. Soc. 2014, 136, 14570-14575.
55 (4) Walsh, A.; Scanlon, D. O.; Chen, S.; Gong, X. G.; Wei, S.-H. Self-Regulation Mechanism for
56 Charged Point Defects in Hybrid Halide Perovskites. Angew. Chem. Int. Ed. 2014, 53, 1-5.
57
58 12
59
60 ACS Paragon Plus Environment
Page 13 of 14 ACS Energy Letters

1
2
3 (5) Miyata, K.; Meggiolaro, D.; Trinh, M. T.; Joshi, P. P.; Mosconi, E.; Jones, S. C.; De Angelis, F.;
4 Zhu, X. Y. Large polarons in lead halide perovskites. Sci. Adv. 2017, 3.
5 (6) Zhu, X. Y.; Podzorov, V. Charge Carriers in Hybrid Organic–Inorganic Lead Halide Perovskites
6 Might Be Protected as Large Polarons. J. Phys. Chem. Lett. 2015, 6, 4758-4761.
7 (7) Adinolfi, V.; Yuan, M.; Comin, R.; Thibau, E. S.; Shi, D.; Saidaminov, M. I.; Kanjanaboos, P.;
8 Kopilovic, D.; Hoogland, S.; Lu, Z.-H.; Bakr, O. M.; Sargent, E. H. The In-Gap Electronic State
9 Spectrum of Methylammonium Lead Iodide Single-Crystal Perovskites. Adv. Mater. 2016, 28,
10 3406-3410.
11
(8) Chen, Y.; Yi, H. T.; Wu, X.; Haroldson, R.; Gartstein, Y. N.; Rodionov, Y. I.; Tikhonov, K. S.;
12
13
Zakhidov, A.; Zhu, X. Y.; Podzorov, V. Extended carrier lifetimes and diffusion in hybrid
14 perovskites revealed by Hall effect and photoconductivity measurements. Nat. Commun. 2016, 7,
15 12253.
16 (9) Leijtens, T.; Eperon, G. E.; Barker, A. J.; Grancini, G.; Zhang, W.; Ball, J. M.; Kandada, A. R.
17 S.; Snaith, H. J.; Petrozza, A. Carrier trapping and recombination: the role of defect physics in
18 enhancing the open circuit voltage of metal halide perovskite solar cells. Energy Environ. Sci. 2016,
19 9, 3472-3481.
20 (10) deQuilettes, D. W.; Vorpahl, S. M.; Stranks, S. D.; Nagaoka, H.; Eperon, G. E.; Ziffer, M. E.;
21 Snaith, H. J.; Ginger, D. S. Impact of microstructure on local carrier lifetime in perovskite solar
22 cells. Science 2015, 348, 683-686.
23 (11) Wang, S.; Jiang, Y.; Juarez-Perez, Emilio J.; Ono, Luis K.; Qi, Y. Accelerated degradation of
24
methylammonium lead iodide perovskites induced by exposure to iodine vapour. Nat. Energy 2016,
25
2, 16195.
26
27 (12) Li, C.; Guerrero, A.; Zhong, Y.; Gräser, A.; Luna, C. A. M.; Köhler, J.; Bisquert, J.; Hildner,
28 R.; Huettner, S. Real-Time Observation of Iodide Ion Migration in Methylammonium Lead Halide
29 Perovskites. Small 2017, 13, 1701711-n/a.
30 (13) Zohar, A.; Levine, I.; Gupta, S.; Davidson, O.; Azulay, D.; Millo, O.; Balberg, I.; Hodes, G.;
31 Cahen, D. What Is the Mechanism of MAPbI3 p-Doping by I2? Insights from Optoelectronic
32 Properties. ACS Energy Lett. 2017, 2, 2408-2414.
33 (14) Senocrate, A.; Moudrakovski, I.; Kim, G. Y.; Yang, T.-Y.; Gregori, G.; Grätzel, M.; Maier, J.
34 The Nature of Ion Conduction in Methylammonium Lead Iodide: A Multimethod Approach.
35 Angew. Chem. Int. Ed. 2017, 56, 7755-7759.
36 (15) Du, M.-H. Density Functional Calculations of Native Defects in CH3NH3PbI3: Effects of Spin–
37 Orbit Coupling and Self-Interaction Error. J. Phys. Chem. Lett. 2015, 6, 1461-1466.
38
(16) Umari, P.; Mosconi, E.; De Angelis, F. Relativistic GW calculations on CH3NH3PbI3 and
39
40
CH3NH3SnI3 Perovskites for Solar Cell Applications. Sci. Rep. 2014, 4, 4467.
41 (17) Alkauskas, A.; Pasquarello, A. Band-edge problem in the theoretical determination of defect
42 energy levels: The O vacancy in ZnO as a benchmark case. Phys. Rev. B 2011, 84, 125206.
43 (18) Iwata, J.-I. S., Keisuke; Oshiyama, Atsushi A precaution for the hybrid density functional
44 calculation of open-shell systems. 2016, arXiv:1605.00338.
45 (19) Van de Walle, C. G.; Neugebauer, J. First-principles calculations for defects and impurities:
46 Applications to III-nitrides. J. Appl. Phys. 2004, 95, 3851-3879.
47 (20) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M.; Alarousu, E.; Buin, A.; Chen, Y.; Hoogland, S.;
48 Rothenberger, A.; Katsiev, K.; Losovyj, Y.; Zhang, X.; Dowben, P. A.; Mohammed, O. F.; Sargent,
49 E. H.; Bakr, O. M. Low trap-state density and long carrier diffusion in organolead trihalide
50 perovskite single crystals. Science 2015, 347, 519-522.
51
(21) Sadoughi, G.; Starr, D. E.; Handick, E.; Stranks, S. D.; Gorgoi, M.; Wilks, R. G.; Bär, M.;
52
53
Snaith, H. J. Observation and Mediation of the Presence of Metallic Lead in Organic–Inorganic
54 Perovskite Films. ACS Appl. Mater. Interfaces 2015, 7, 13440-13444.
55 (22) Azpiroz, J. M.; Mosconi, E.; Bisquert, J.; De Angelis, F. Defect migration in methylammonium
56 lead iodide and its role in perovskite solar cell operation. Energy Environ. Sci. 2015, 8, 2118-2127.
57
58 13
59
60 ACS Paragon Plus Environment
ACS Energy Letters Page 14 of 14

1
2
3 (23) Ming, W.; Chen, S.; Du, M.-H. Chemical instability leads to unusual chemical-potential-
4 independent defect formation and diffusion in perovskite solar cell material CH3NH3PbI3. J. Mater.
5 Chem. A 2016, 4, 16975-16981.
6 (24) Yang, J.-H.; Yin, W.-J.; Park, J.-S.; Wei, S.-H. Fast self-diffusion of ions in CH3NH3PbI3: the
7 interstiticaly mechanism versus vacancy-assisted mechanism. J. Mater. Chem. A 2016, 4, 13105-
8 13112.
9 (25) Eames, C.; Frost, J. M.; Barnes, P. R. F.; O'Regan, B. C.; Walsh, A.; Islam, M. S. Ionic
10 transport in hybrid lead iodide perovskite solar cells. Nat. Commun. 2015, 6, 7497.
11
(26) Minns, J. L.; Zajdel, P.; Chernyshov, D.; van Beek, W.; Green, M. A. Structure and interstitial
12
13
iodide migration in hybrid perovskite methylammonium lead iodide. Nat. Commun. 2017, 8, 15152.
14 (27) Boschloo, G.; Hagfeldt, A. Characteristics of the Iodide/Triiodide Redox Mediator in Dye-
15 Sensitized Solar Cells. Acc. Chem. Res. 2009, 42, 1819-1826.
16 (28) Mosconi, E.; Meggiolaro, D.; Snaith, H. J.; Stranks, S. D.; De Angelis, F. Light-induced
17 annihilation of Frenkel defects in organo-lead halide perovskites. Energy Environ. Sci. 2016, 9,
18 3180-3187.
19 (29) Mosconi, E.; De Angelis, F. Mobile Ions in Organohalide Perovskites: Interplay of Electronic
20 Structure and Dynamics. ACS Energy Lett. 2016, 1, 182-188.
21 (30) Zhang, L.; Sit, P. H. L. Ab initio study of the role of iodine in the degradation of CH3NH3PbI3.
22 J. Mater. Chem. A 2017, 5, 23976-23986.
23 (31) deQuilettes, D. W.; Zhang, W.; Burlakov, V. M.; Graham, D. J.; Leijtens, T.; Osherov, A.;
24
Bulovic, V.; Snaith, H. J.; Ginger, D. S.; Stranks, S. D. Photo-induced halide redistribution in
25
organic-inorganic perovskite films. Nat. Commun. 2016, 7, 11683.
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 14
59
60 ACS Paragon Plus Environment

You might also like