We report the presence of CH3NH2 neutral species not only on the surface but also at grain boundaries in the interior of thin polycrystalline films of methylammonium lead iodide perovskite CH3NH3PbI3 (thickness ∼ 50 nm) that were prepared using a standard solution method. Different chemical states for C K-edge were observed at the surfaces and in the interiors of perovskite films. Salient features of σ*(CH3-NH3+: methylammonium cation) and σ*(CH3-NH2: methylamine neutral species) were observed at 290.3 and 292.8 eV in both partial (surface-sensitive) and total (bulk) electron yield modes by near-edge x-ray absorption fine structure measurements. Consistently, two chemical states originated from CH3NH3+ and CH3NH2 in C 1s and N 1s core-level spectra were observed using high-resolution x-ray photoelectron spectroscopy. Using density functional theory calculations, we show that CH3NH2 cannot reside stably in the MAPbI3 perovskite crystal structure. Therefore, we propose that these CH3NH2 neutral species are mainly located on the surface or at grain boundaries in the interior of CH3NH3PbI3 films.

Solar cells based on organometal halide perovskite have been evolving rapidly. The newly achieved efficiencies exceeding 20%1 make perovskite solar cells a promising candidate to compete with even the highest efficiency single crystal Si cells.2–5 Organometal halide perovskite shows several attractive properties for photovoltaic applications, such as wide absorption photon range (∼800 nm),2 weak exciton-binding energy (<50 meV),6 and high carrier mobilities (electron ∼ 8 and hole ∼ 12 cm2/V s).7,8 To further improve device performance, it is crucial to understand the exact chemical structure and the precise role of each interface in devices.9 Especially, several research groups studied the role of the organic cation CH3NH3+ for a better understanding about the material stability and degradation mechanism.10,11 Although a few theoretical studies have been reported,12–14 experimental studies are somewhat lagging behind. If we understand the exact chemical structure and the precise role of each interface in perovskite solar cells, we can establish an optimized fabrication method, taking into consideration necessary interface treatment on electron and hole transport layers, minimizing degradation factors, optimized material design, etc.

In this study, we used near-edge x-ray absorption of fine structure (NEXAFS) and high-resolution x-ray photoelectron spectroscopy (HRXPS) with synchrotron radiation to understand the exact chemical structure of organometal halide perovskite at the surface as well as in the interior of the bulk film. We found the presence of CH3NH2 neutral species at surface and in the interiors of organometal halide perovskite. Based on our experimental results and density functional theory calculations, we propose that the neutral CH3NH2 species mainly reside either on the surface or at grain boundaries in the interior of CH3NH3PbI3 films.

The patterned, fluorine-doped tin oxide glass (FTO Pilkington, 7 Ω/◻) was prepared with HCl and Zn powder and cleaned. On this substrate, we deposited a 100-nm layer of compact layer (cl) of TiO2 using spray pyrolysis with a precursor solution of acetylacetone, Ti (IV) isopropoxyde, and anhydrous ethanol (3:3:2) and then performed post-annealing at 480 °C on a hotplate. Basically, perovskite thin film fabrication followed a modified version of the two-step method of Burschka et al.15 First, a solution of PbI2 in dimethylformamide (460 mg ml−1) was prepared and stirred at 70 °C for ≥2 h. We performed spin-coating at 6000 rpm for 30 s on the cl-TiO2 substrates, previously heated at 70 °C. After the spin coating, the PbI2 layer was dried at 70 °C for 20 min. Then, a 20 mg/ml methylammonium iodide (MAI) solution in 2-propanol (IPA) was prepared and kept at 70 °C. PbI2 films were dipped in the MAI solution for 30 s with gentle shaking of the substrates. After dipping, substrates were rinsed in copious IPA and dried immediately by spinning them on a spin-coater, after which they were annealed for 20 min on the hot plate at 70 °C. To confirm the film thickness and surface morphology, we used an AFM (MFP-3D Asylum Research) and scanning electron microscopy (Quanta 250 FEG, FEI). X-ray diffraction (D8 Discover, Bruker Corporation) and UV-Vis spectrophotometry (Evolution 600 UV-Vis spectrophotometer, Thermo Scientific) were used to determine the crystalline structure and optical bandgap, respectively. All NEXAFS measurements were performed at room temperature and at the beamline 4D of Pohang Light Source II (PLS II), Korea. We used the total electron yield (TEY) and partial electron yield (PEY) detection mode for NEXAFS spectra by recording the sample current normalized to a signal current, which was measured simultaneously using a gold mesh in ultrahigh vacuum (<10−9 Torr). In this case, a p-polarized (≈85%) synchrotron photon beam, with an energy in the range of 279–320 eV and a spectral energy resolution of ΔE = 150 meV, had probing depths of 6 nm and 25 nm, for PEY and TEY modes, respectively. For HRXPS measurements, we performed two experiments using synchrotron radiation (52–860 eV) and a monochromated Al (1486.6 eV) source. HRXPS spectra were obtained at the 10D HRXPS beamline of PLSII. Photon energy was varied from 50 eV to 860 eV to obtain high-quality XPS spectra. Photoelectron signals were recorded with a PHOIBOS 150 electron energy analyzer equipped with a two-dimensional charge-coupled device (2D CCD) detector (Specs GmbH), collecting photoelectrons from the surface normal. The base pressure of the main chamber was maintained below 9.8 × 10−11 Torr. In the case of lab source, we used the monochromated Al (1486.6 eV) and a multi-channel plate analyzer with 180° hemispherical and 165 mm mean radii in the AXIS Ultra DLD (Kratos Surface Analysis). The energy resolution was <0.5 eV in both techniques. Chemical states of all samples were obtained measuring the valence region and the Pb 4f, I 3d, C 1s, and N 1s core-levels spectra. O 1s core-level was also measured to monitor any contamination during the sample transfer. O 1s core-level intensity was below the detection limits of the instruments. Binding energies were calibrated with reference to the Au 4f7/2 level (84.0 eV).16 We fitted both C and N 1s core-level peaks using Doniach-Sŭnjić curves convoluted with Gaussian distribution of 0.5 eV full-width at half-maximum.17 Background due to inelastic scattering was subtracted by the Shirley (or integral) method.18 All calculations were performed using density functional theory (DFT) as implemented in the code of the CASTEP.19 We used ultrasoft Vanderbilt-type pseudopotentials20 and the exchange-correlation functional of the Perdew-Burke-Ernzerhof expression revised for solids (PBEsol) in the generalized gradient approximation (GGA).21,22 Electronic wave functions were expanded by plane waves with an energy cut-off of 280 eV. The k-space integration was performed using a 3 × 3 × 1 mesh in the Brillouin zone (BZ) of a three-dimensionally periodic 3 × 3 × 1 supercell. The geometry optimization for the tetragonal phase was carried out until the Hellmann-Feynman force acting on the atoms was smaller than 0.03 eV/Å. To obtain the most stable atomic structure for (CH3NH3,CH3NH2)PbI3, various initial atomic configuration were considered and all the atoms in the supercell were fully relaxed without any symmetry constraint. Also, we used the Mulliken charge analysis method to analyze charge populations. The binding energy Esafe had the following form:

Ebindingm,n=Etotalm,nEPbI3m×ECH3NH3n×ECH3NH2,
EbindingCH3NH3=(ECH3NH3+PbI3totalEPbI316×ECH3NH3)/16,
EbindingCH3NH2=(ECH3NH2+PbI3totalEPbI316×ECH3NH2)/16,
Ebindingm,n=m×EbindingCH3NH3+n×EbindingCH3NH2,
Esafe=Ebindingm,nEbindingm,n,

where m is the number of CH3NH3 molecules, n is the number of CH3NH2 molecules, Ebindingm,n is the binding energy of CH3NH3 molecule m and CH3NH2 molecule n in the PbI3 3 × 3 × 1 unit cell, Etotalm,n is the total energy of CH3NH3 molecule m and CH3NH2 molecule n in the PbI3 3 × 3 × 1 unit cell, EPbI3 is the total energy of the PbI3 3 × 3 × 1 unit cell, ECH3NH3 is the energy of CH3NH3 single molecule, ECH3NH2 is the energy of CH3NH2 single molecule, EbindingCH3NH3 is the binding energy of CH3NH3 single molecule, EbindingCH3NH2 is the binding energy of CH3NH2 single molecule, ECH3NH3+PbI3total is the total energy of 16 CH3NH3 molecules in the PbI3 3 × 3 × 1 unit cell, and ECH3NH2+PbI3total is the total energy of 16 CH3NH2 molecules in the PbI3 3 × 3 × 1 unit cell

In these forms, we considered various ratios of CH3NH3 and CH3NH2 molecules in the tetragonal-phase PbI3 structure with 16 kinds of mixing conditions. In the equation, negative and positive values of Esafe represent stable and unstable structures of (CH3NH3,CH3NH2)PbI3, respectively.

The solution-prepared CH3NH3PbI3 (MAPbI3) film was measured by NEXAFS with synchrotron radiation. We measured both partial (PEY, probing-depth of ∼6 nm) and total electron yield (TEY, ∼25 nm) modes to see whether electronic structure differed between surface and bulk organometal halide perovskite.23 Different chemical states are observed only in PEY mode (Figure 1(a)) at the binding energy of ∼285 eV. The rest of the spectra are almost the same between the PEY and TEY mode measurements in solution-prepared MAPbI3. The most significant absorption features are the two shoulders at the binding energy of ∼290 eV (Figure 1(a)). To understand these surface and bulk features, we performed curve-fittings for both of them (Figure 1(b)). We found σ*(CH3-NH3+), σ*(CH3-NH2), and σ*(C*=C) at 290.3, 292.8, and 301.4 eV in both measurements.23 The carbon chemical states C*-M (Metal) and π* (C*=C) were found at 282.9 and 285.5 eV in the PEY mode.23 In the TEY mode, however, we could not observe any of these chemical states in the energy range from 280.0 to 285.5 eV.

FIG. 1.

Observation of two organic elements in the bulk and surface of perovskite and their differences in chemical states. (a) C K-edge NEXAFS in PEY (surface sensitivity) and TEY (bulk sensitivity) modes. The photon energy peak at 285 eV and two significant peaks around 290 eV are apparent. (b) Curve-fitting data.

FIG. 1.

Observation of two organic elements in the bulk and surface of perovskite and their differences in chemical states. (a) C K-edge NEXAFS in PEY (surface sensitivity) and TEY (bulk sensitivity) modes. The photon energy peak at 285 eV and two significant peaks around 290 eV are apparent. (b) Curve-fitting data.

Close modal

The photoelectron escaping depth in a thin film depends on the energy of incident photons.24 To investigate chemical structure differences between surface and bulk, we performed high-resolution x-ray photoelectron spectroscopy with synchrotron radiation and a monochromated Al Kα source. The Pb 4f7/2 and I 4d5/2 core-levels spectra at binding energies of 138.7 and 49.7 eV show no significant change in chemical states or shapes for all photon energies measured (Figure 2(a)). There is only a single chemical state for each of these peaks. In the case of C 1s core-level spectra, however, we observed different shapes when different incident photon energies (370, 620, and 1486.7 eV) were used (Figure 2(b)). Using photo energy of 370 eV, the C 1s core-level spectrum shows some tail states around 282.5 eV that do not appear in the case of 620 or 1486.7 eV photon energy. The escaping depth of photoelectrons (inelastic mean free path) with a photon energy of 370 eV is ∼1 nm, while the escaping depth for 620 or 1486.7 eV are ∼2 nm and ∼5 nm, respectively.24 The presence of these tail states only for low photon energy of 370 eV suggests that these tail states are associated with the surface of the solution-prepared MAPbI3 thin film. In the case of N 1s core-level spectra, we could not observe any significant surface states (Figure 2(b)). To understand chemical states of C and N 1s core-levels, we performed curve fittings and obtained several chemical states of C1 (287 eV), C2 (285.4 eV), S1 (283.2 eV), S2 (281.2 eV), N1 (403.3 eV), and N2 (402.7 eV) (Figure 2(b)). We propose that the chemical states of C1, N1 and C2, N2 are from CH3-NH3+ and CH3-NH2.25,26 S1 and S2 are originated from the surface states of C*-M (Metal) and C*=C bonding.27,28 These HRXPS results are consistent with the NEXAFS results.

FIG. 2.

HRXPS measurements with synchrotron radiation. (a) Pb 4f and I 4d core-level spectra measured at various photon energies. No changes of Pb 4f and I 4d chemical states are observed at different photon energies. (b) C and N 1s core-levels spectra with various photon energies and after the optimized curve-fitting analysis. C1, N1 and C2, N2 were assigned to CH3-NH3+ and CH3-NH2 chemical states, respectively. S1 and S2 are species corresponding to carbon metallic chemical states.

FIG. 2.

HRXPS measurements with synchrotron radiation. (a) Pb 4f and I 4d core-level spectra measured at various photon energies. No changes of Pb 4f and I 4d chemical states are observed at different photon energies. (b) C and N 1s core-levels spectra with various photon energies and after the optimized curve-fitting analysis. C1, N1 and C2, N2 were assigned to CH3-NH3+ and CH3-NH2 chemical states, respectively. S1 and S2 are species corresponding to carbon metallic chemical states.

Close modal

Finally, we found CH3NH2 in the organometal halide perovskite with different chemical states between bulk and surface. On the other hand, Pb and I exhibited the same single chemical states in both locations. If CH3NH2 resides inside the perovskite crystal structure, it is expected to introduce more than one chemical states of Pb or I. To determine the location of CH3NH2 in the perovskite, we performed the air exposure experiment (Figure S2 in the supplementary material).29 It has been reported that water (H2O) in air degrades perovskite.30,31 Specifically, CH3NH3+ becomes CH3NH2.30,31 If the C1 and C2 chemical states are originated from CH3NH3+ and CH3NH2, respectively, it is expected that the ratio between C1 and C2 intensities decreases after air exposure. We performed the air exposure experiment at 24 °C and 30% humidity and then measured the sample with HRXPS after 1 and 3 days (Figure S2 in the supplementary material).29 Indeed, the ratio between C1 and C2 intensities decreased after air exposure (Figure S2 in the supplementary material),29 which is consistent with the proposal that the C2 chemical state is originated from CH3NH2 neutral species.

Based on these experimental observations, we propose that CH3NH2 exists in solution-prepared organometal halide perovskites. In addition, one important question arises regarding the location of CH3NH2 in CH3NH3PbI3 films. Therefore, we further investigated the charge populations, density of states (DOS), and system stability of CH3NH3 and CH3NH2 molecules in PbI3 using DFT-based simulations. In these calculations, we considered the tetragonal-phase PbI3 structure (Figure 3(a)).32 First, we investigated the atomic charge states and the C-N bond length of the isolated CH3NH3 and CH3NH2 molecules (Figure 3(b)). Basically, negative or positive values of charge populations are represented by electron accumulation or depletion relative to what it had. Charge populations of CH3NH3 and CH3NH2 molecules are changed in the tetragonal-phase PbI3 in comparison with those of isolated CH3NH3 and CH3NH2 (Figures 3(c) and 3(d)). In particular, the change of charge population of I atoms is very sensitive to the presence of CH3NH3 or CH3NH2 molecules in PbI3. We used the binding energies (Esafe) for the stability of possible structures when CH3NH3 and CH3NH2 coexist in PbI3 (Table S1 in the supplementary material).29 The calculated results of Esafe show positive values for all configurations of (CH3NH3, CH3NH2)PbI3. This indicates that CH3NH2 is unstable in the interior region of PbI3, and instead, it should be located at grain boundaries of thin polycrystalline films of CH3NH3PbI3. In this case, CH3NH2 cannot directly interact with the basic building blocks of PbI3. This conclusion is consistent with the experimental observation of a single chemical state for Pb and I in HRXPS measurements. In addition, DFT calculations of the DOS show different band gaps for CH3NH3PbI3 and CH3NH2PbI3; the band gap of CH3NH3PbI3 was ∼1.70 eV, while CH3NH2PbI3 exhibits no band gap (Figure S3 in the supplementary material).29 The DFT result of CH3NH2PbI3 with no band gap is inconsistent with the experimental result of optical band gap of ∼1.55 eV (Figure S1 in the supplementary material).29 This also suggests that CH3NH2 cannot reside stably inside the CH3NH3PbI3 crystal structure, but rather resides mainly on the surface or at grain boundaries in the interior of thin polycrystalline films of CH3NH3PbI3. A direct correlation between the concentration of CH3NH2 in perovskite cells and their solar cell performance would be particularly helpful. But, unfortunately, it is challenging to accurately determine the total concentration of CH3NH2 for a bulk film of perovskite with a typical thickness of ∼300 nm as those used in perovskite solar cells. Nevertheless, such a correlation can be partially inferred by the study that investigated the effects of annealing conditions on perovskite solar cell performance.33 In that study, the solar-cell efficiency of 10.7% was obtained for perovskite solar cells using perovskite films with post-annealing at 115 °C in N2 environment, while the efficiency was substantially lower (6.7%) with post-annealing at 105 °C.33 Note that the two types of perovskite solar cells were otherwise prepared in an identical way. Presumably, post-annealing at 105 °C was not sufficient to remove the majority of CH3NH2 from the perovskite film and, therefore, could be a likely cause responsible for the lower efficiency.33 Regarding the origin of CH3NH2, it is possible that after perovskite films are prepared, a certain amount of unreacted CH3NH3I still remains in perovskite films even after annealing. CH3NH3I can further decompose to CH3NH2, leading to the presence of CH3NH2 in perovskite films. In addition, when CH3NH3PbI3 perovskite films degrade, CH3NH2 can also be generated as a product, which can be another origin of CH3NH2 in perovskite films.

FIG. 3.

Geometry-optimized atomic structures for PbI3 (a), isolated CH3NH3 and CH3NH2 molecules (b), CH3NH3PbI3 (c), and CH3NH2PbI3 (d). In (b)–(d), charge states of C, N, Pb, and I are also shown using a Mulliken population analysis. It is noted that distinct chemical states of I are observed when CH3NH2 is bonded to the PbI3 structure.

FIG. 3.

Geometry-optimized atomic structures for PbI3 (a), isolated CH3NH3 and CH3NH2 molecules (b), CH3NH3PbI3 (c), and CH3NH2PbI3 (d). In (b)–(d), charge states of C, N, Pb, and I are also shown using a Mulliken population analysis. It is noted that distinct chemical states of I are observed when CH3NH2 is bonded to the PbI3 structure.

Close modal

In conclusion, we found CH3NH2 neutral species on the surface and in the interior of CH3NH3PbI3 films prepared using a typical solution method. Based on the experimental results and density functional theory calculations, we conclude that the CH3NH2 neutral species is mainly located on the surface or at grain boundaries in the interior of CH3NH3PbI3 films, which is expected to affect device performance and degradation. The observation of different chemical states on the surface of CH3NH3PbI3 films such as C-metal and C*=C bonding is also discussed.

This work was supported by funding from the Energy Materials and Surface Sciences Unit of the Okinawa Institute of Science and Technology Graduate University and JSPS KAKENHI Grant No. 15K17925. Y.M.L and H.-K.L are grateful for financial support from the Basic Science Research Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2015R1C1A2A01054543 and NRF-2011-0006744). The experiment at the PLSII was supported, in part, by MEST and POSTECH. Calculations were supported by the Nano Material Technology Development Program (2012M3A7B4049888) through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and Future Planning (MSIP), and the Priority Research Center Program (2010-0020207) through NRF funded by the Ministry of Education (MOE). The authors thank Steven D. Aird, the Technical Editor at Okinawa Institute of Science and Technology Graduate University (OIST) for valuable suggestions in revising the manuscript.

1.
W. S.
Yang
,
J. H.
Noh
,
N. J.
Jeon
,
Y. C.
Kim
,
S.
Ryu
,
J.
Seo
, and
S. I.
Seok
,
Science
348
,
1234
(
2015
).
2.
M. A.
Green
,
A.
Ho-Baillie
, and
H. J.
Snaith
,
Nat. Photonics
8
,
506
(
2014
).
3.
M.
Grätzel
,
Nat. Mater.
13
,
838
(
2014
).
4.
W.
Nie
,
H.
Tsai
,
R.
Asadpour
,
J.-C.
Blancon
,
A. J.
Neukirch
,
G.
Gupta
,
J. J.
Crochet
,
M.
Chhowalla
,
S.
Tretiak
,
M. A.
Alam
,
H.-L.
Wang
, and
A. D.
Mohite
,
Science
347
,
522
(
2015
).
5.
S. D.
Stranks
and
H. J.
Snaith
,
Nat. Nanotechnol.
10
,
391
(
2015
).
6.
V.
D'Innocenzo
,
G.
Grancini
,
M. J. P.
Alcocer
,
A. R. S.
Kandada
,
S. D.
Stranks
,
M. M.
Lee
,
G.
Lanzani
,
H. J.
Snaith
, and
A.
Petrozza
,
Nat. Commun.
5
,
3586
(
2014
).
7.
C.
Wehrenfennig
,
G. E.
Eperon
,
M. B.
Johnston
,
H. J.
Snaith
, and
L. M.
Herz
,
Adv. Mater.
26
,
1584
(
2013
).
8.
C.
Motta
,
F.
El-Mellouhi
, and
S.
Sanvito
,
Sci. Rep.
5
,
12746
(
2015
).
9.
M. A.
Loi
and
J. C.
Hummelen
,
Nat. Mater.
12
,
1087
(
2013
).
10.
W. J.
Yin
,
T.
Shi
, and
Y.
Yan
,
Adv. Mater.
26
,
4653
(
2014
).
11.
J. M.
Azpiroz
,
E.
Mosconi
,
J.
Bisquert
, and
F.
De Angelis
,
Energy Environ. Sci.
8
,
2118
(
2015
).
12.
Y.
Wang
,
T.
Gould
,
J. F.
Dobson
,
H.
Zhang
,
H.
Yang
,
X.
Yao
, and
H.
Zhao
,
Phys. Chem. Chem. Phys.
16
,
1424
(
2013
).
13.
A.
Buin
,
R.
Comin
,
J.
Xu
,
A. H.
Ip
, and
E. H.
Sargent
,
Chem. Mater.
27
,
4405
(
2015
).
14.
I. E.
Castelli
,
J. M.
García-Lastra
,
K. S.
Thygesen
, and
K. W.
Jacobsen
,
APL Mater.
2
,
081514
(
2014
).
15.
J.
Burschka
,
N.
Pellet
,
S.-J.
Moon
,
R.
Humphry-Baker
,
P.
Gao
,
M. K.
Nazeeruddin
, and
M.
Grätzel
,
Nature
499
,
316
(
2013
).
16.
C. D.
Wagner
,
W. M.
Riggs
,
L. E.
Davis
, and
J. F.
Moulder
,
Handbook of X-Ray Photoelectron Spectroscopy, a Reference Book of Standard Data for Use in X-Ray Photoelectron Spectroscopy
(
Perkin–Elmer Corp.
,
1979
).
17.
S.
Doniach
and
M.
Sunjic
,
J. Phys. C: Solid State Phys.
3
,
285
(
1970
).
18.
D. A.
Shirley
,
Phys. Rev. B
5
,
4709
(
1972
).
19.
S. J.
Clark
,
M. D.
Segall
,
C. J.
Pickard
,
P. J.
Hasnip
,
M. I. J.
Probert
,
K.
Refson
, and
M. C.
Payne
,
Z. Kristallogr
220
,
567
570
(
2005
).
20.
D.
Vanderbilt
,
Phys. Rev. B
41
,
7892
(
1990
).
21.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
,
Phys. Rev. Lett.
100
,
136406
(
2008
).
22.
F.
Brivio
,
A. B.
Walker
, and
A.
Walsh
,
APL Mater.
1
,
042111
(
2013
).
23.
J.
Stöhr
,
NEXAFS Spectroscopy
, Springer Series in Surface Sciences Vol. 25 (
Springer
,
1992
).
24.
D.
Briggs
and
M. P.
Seah
,
Practical Surface Analysis, Auger and X-ray Photoelectron Spectroscopy
(
Wiley
,
1990
).
25.
A. P.
Dementjev
,
A.
de Graaf
,
M. C. M.
van de Sanden
,
K. I.
Maslakov
,
A. V.
Naumkin
, and
A. A.
Serov
,
Diamond Relat. Mater.
9
,
1904
(
2000
).
26.
S.
Bhattacharyya
,
C.
Cardinaud
, and
G.
Turban
,
J. Appl. Phys.
83
,
4491
(
1998
).
27.
G.
Deniau
,
L.
Azoulay
,
P.
Jégou
,
G.
Le Chevallier
, and
S.
Palacin
,
Surf. Sci.
600
,
675
(
2006
).
28.
U.
Gelius
,
P. F.
Hedén
,
J.
Hedman
,
B. J.
Lindberg
,
R.
Manne
,
R.
Nordberg
,
C.
Nordling
, and
K.
Siegbahn
,
Phys. Scr.
2
,
70
(
1970
).
29.
See supplementary material at http://dx.doi.org/10.1063/1.4941994 for the sample fabrication, characterization, air-exposure result, and bandgap calculation.
30.
J. M.
Frost
,
K. T.
Butler
,
F.
Brivio
,
C. H.
Hendon
,
M.
van Schilfgaarde
, and
A.
Walsh
,
Nano Lett.
14
,
2584
(
2014
).
31.
Y.
Han
,
S.
Meyer
,
Y.
Dkhissi
,
K.
Weber
,
J. M.
Pringle
,
U.
Bach
,
L.
Spiccia
, and
Y.-B.
Cheng
,
J. Mater. Chem. A
3
,
8139
(
2015
).
32.
T.
Baikie
,
Y.
Fang
,
J. M.
Kadro
,
M.
Schreyer
,
F.
Wei
,
S. G.
Mhaisalkar
,
M.
Graetzel
, and
T. J.
White
,
J. Mater. Chem. A
1
,
5628
(
2013
).
33.
S. R.
Raga
,
M.-C.
Jung
,
M. V.
Lee
,
M. R.
Leyden
,
Y.
Kato
, and
Y. B.
Qi
,
Chem. Mater.
27
,
1597
(
2015
).

Supplementary Material