You are on page 1of 10

View Article Online

View Journal

PCCP
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: J. A. Idígoras, A.
Todinova, J. R. Sanchez-Valencia, A. Barranco, A. Borrás and J. A. Anta, Phys. Chem. Chem. Phys., 2016,
DOI: 10.1039/C6CP01265E.

This is an Accepted Manuscript, which has been through the


Royal Society of Chemistry peer review process and has been
accepted for publication.

Accepted Manuscripts are published online shortly after


acceptance, before technical editing, formatting and proof reading.
Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
standard Terms & Conditions and the Ethical guidelines still
apply. In no event shall the Royal Society of Chemistry be held
responsible for any errors or omissions in this Accepted Manuscript
or any consequences arising from the use of any information it
contains.

www.rsc.org/pccp
Page 1 of 9 Please
Physical do not adjust
Chemistry margins
Chemical Physics
View Article Online
DOI: 10.1039/C6CP01265E

Journal Name

Physical Chemistry Chemical Physics Accepted Manuscript


ARTICLE

The Interaction between Hybrid Organic-Inorganic Halide


Perovskite and Selective Contacts in Perovskite Solar Cells: an
Received 00th January 20xx,
Accepted 00th January 20xx
Infrared Spectroscopy Study
J. Idígoras,*,a A. Todinova,a J.R. Sánchez-Valencia,b A. Barranco,b A. Borrásb and J.A. Anta.a
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

DOI: 10.1039/x0xx00000x

www.rsc.org/ The interaction of hybrid organic-inorganic halide perovskite and selective contacts is crucial to get efficient, stable and
hysteresis-free perovskite-based solar cells. In this report, we analyze the vibrational properties of methylammonium lead
halide perovskites deposited on different substrates by infrared absorption (IR) measurements (4000-500 cm-1). The
materials employed as substrates are not only characterized by a different chemical nature (TiO2, ZnO and Al2O3), but also
by different morphologies. For all of them, we have investigated the influence of these substrate properties on the
perovskite formation and its degradation by humidity. The effect of selective-hole contact (Spiro-OmeTad and P3HT) layers
on the degradation rate by moisture has also been studied. Our IR results reveal the existence of a strong interaction
between perovskite and all ZnO materials considered, evidenced by a shift of the peaks related to the N-H vibrational
modes. The interaction even induces a morphological change of ZnO nanoparticles after perovskite deposition, pointing to
an acid-base reaction that takes place through the NH3+ groups of the methylammonium cation. Our IR and X-ray
diffraction results also indicate that this specific interaction favors perovskite decomposition and PbI2 formation for
ZnO/perovskite films submitted to humid conditions. Although no interaction is observed for TiO2, Al2O3, and hole
selective contact, the morphology and chemical nature of both contacts appear to play an important role in the rate of
degradation upon moisture exposure.

layer to provide a porous substrate on which the perovskite is


1. Introduction deposited. Additionally, on top of that, a hole-selective layer is also
deposited. In these photovoltaic devices, the charge extraction is
The origin of hybrid organic-inorganic perovskite solar cells (PSCs) is carried out thanks to these selective contacts: TiO2 (n-contact layer)
based on the field of dye-sensitized solar cells.1,2 When CH3NH3PbI3 and HSM (p-contact layer). Thereby, the electrons are transferred
(MAPbI3) was for the first time used as sensitizer of TiO2, a power to TiO2, a wide bandgap semiconductor metal oxide, whereas holes
conversion efficiency of 3.5% was achieved in combination with a are collected by the HSM. Thus, the TiO2 layer does not only act as
liquid electrolyte.3 The power conversion efficiency was raised to scaffold of the perovskite material, but also extracts the
6.5% after optimization of titania surface, perovskite processing and photogenerated electrons. Nevertheless, in 2012 Snaith et al.6
the concentration of liquid electrolyte.4 Nevertheless, a significant demonstrated that the perovskite material in contact with an
improvement of efficiency and stability took place when MAPbI3 insulating mesoporous Al2O3 scaffold was also capable of producing
was combined with solid organic hole-selective materials (HSMs)5 in efficiencies exceeding 12%. In addition, special attention has
a thin film6,7 or mesoporous8–10 configuration. In particular, certified recently been paid to ZnO as n-contact layer. This oxide draws
power conversion efficiencies higher than 20% has recently been interest due to its capability to fabricate flexible devices and to
reported.11,12 reduce the production costs because of the wide variety of
Most state-of-the-art PSCs employ a compact and mesoporous TiO2 techniques and the lower temperature processing that is required
to deposit this material.13–21 Additionally, organic n-type materials
such as fullerenes can also be used as electron selective
contacts.22,23
The most extended perovskite materials used in photovoltaic
application are hybrid organic-inorganic compounds with a 3D
crystal structure of general chemical formula ABX3 (A = organic
cation, B = Pb and X = halide). Methylammonium (CH3NH3+, MA) or
formamidinium (HC(NH2)2+, FA) are the most common organic
cations, whereas I, Cl and Br are the halides. The chemical nature of
this combination of species not only characterizes the structural

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 1

Please do not adjust margins


Please
Physical do not adjust
Chemistry margins
Chemical Physics Page 2 of 9
View Article Online
DOI: 10.1039/C6CP01265E
ARTICLE Journal Name
24,25
properties, but also the optoelectronic behavior, such as the 2.1 n-contact oxide layers and MAPbI3 film preparation
26–29 25,30
hysteresis effects or the determination of the bandgap. For Silicon wafer was cut and cleaned by sonication in acetone and 2-
instance, a bandgap of 1.57 and 1.48 eV were found for MAPbI3 and propanol followed by drying in a N2 stream. The n-contact oxide
30
FAPbI3, respectively. layers were deposited using different materials and methods (Table
Bearing in mind that the interaction between both organic and

Physical Chemistry Chemical Physics Accepted Manuscript


1). To deposit dense TiO2 layer (cTiO2), 0.65 ml of Titanium (IV)
inorganic components and the chemical nature of the perovskite is isopropoxide (Sigma Aldrich, 97%) and 0.38 ml of Acetylacetone
governed by a variety of chemical bonds, including hydrogen bonds, (Sigma Aldrich, 99%) were mixed with 5 ml of ethanol. This solution
the study of the vibrational modes of these perovskite materials can was spin-coated at 3000 rpm for 60s over the substrates. The
provide valuable clues to understand their optoelectronic mesoporous TiO2 layers were prepared with the 18NR-T (mTiO2)
properties, their photovoltaic behavior and their stability. Several and 18NR-AO (sTiO2) acquired from Dyesol. The commercial TiO2
studies about the vibrational properties of perovskite materials nanoparticle pastes were diluted 2:7 (w:w) in pure ethanol and
have been carried out by Raman and infrared (IR) spectroscopy. spin-coated (5000 rpm, 30 s). The TiO2 layers were sintered then 30
31
Quarti et al. combined computational and experimental data to min at 500°C. The ZnO layers were prepared using a colloidal
32
explain the basis of Raman spectra of MAPbI3. Gottesman et al. suspension of ZnO nanoparticles obtained by a forced hydrolysis
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

observed extremely slow changes in Raman spectra under 44


method diluted 1:1 (w:w) in ethanol (hZnO) or using a solution in
illumination conditions with respect to darkness, which were ethanol of commercial ZnO nanoparticles: Pi-Kem (pZnO) and
33
associated to structural changes. Ledinsky et al. studied the Evonik (eZnO). These solutions were spreaded onto the substrate by
relationship between the structural changes of perovskite and its spin-coating (5000 rpm, 30 s) and then calcined 30 min at 450°C.
degradation. These authors also discussed their Raman spectra Al2O3 thin films were prepared under glancing angle configuration
based on the ratio of iodine-to-bromine atoms in perovskite layers. in an electron bombardment evaporator in a high vacuum system
34 -6
Onoda-Yamamuro et al. showed the variation of the IR spectra of with a base pressure below 5 x 10 mbar. Aluminum pellets were
35 -4
hybrid halide perovskites at different temperatures. Mosconi et al. used as precursor and the evaporation was carried out at 5 x 10
investigated the influence of temperature on IR spectra of MAPbI3 mbar by dosing oxygen to the system. The substrates were placed
to study the dynamics of MA in these hybrid perovskites. Pérez- o o
under oblique angles of 85 and 65 with respect to the evaporator
36
Osorio et al. reported the vibrational properties of MAPbI3 using and deposition was performed at room temperature. To ensure the
Raman and IR data, and combining theoretical and experimental oxidation of the thin films, an additional annealing in air was
37
data. Glaser et al. analyzed the shifts of the peaks associated to performed at 500°C during 12 hours. The thickness of the samples
+
the NH3 stretching vibrations when the halide atoms in the hybrid was monitored through a Quartz Crystal Monitor, which at the
38
halide perovskites are changed. Müller et al. proposed a mentioned oblique geometries result ca. 300 and 530 nm thickness
molecular model that explains the interaction of water molecules o o
for Al2O3(85 ) and Al2O3(65 ) thin films, respectively. These
and MA by IR spectra performed in different ambient conditions. experimental conditions ensure highly transparent and porous thin
Finally, very recently, Bakulin et al resolved on the time-scale the 45,46
films with a microstructure consisting in tilted columns.
39
reorientation of MAI on MAPbI3 lattice by 2-dimensional IR.
Up to the date, all previous works have studied the vibrational
properties of hybrid organic-inorganic halide perovskites focusing Table 1: Summary of the substrates studied.
on the interaction between the organic cations and halide ions
Material Preparation Label
and/or the influence of different conditions. Nevertheless, it is well-
known the importance of the perovskite/substrate interface to get Compact TiO2 Spin-coated from precursors
cTiO2
23,40–43
efficient, stable and hysteresis–free devices. In particular, the
nature and structural properties of the materials that can be used Mesoporous TiO2 Spin-coated from commercial
mTiO2
as scaffold and/or electron selective layer, not only determine the Particle size: 20 nm paste (18NR-T, Dyesol)
perovskite morphology and loading, but also the quality of the
Mesoporous TiO2 Spin-coated from commercial
interaction in the interface. To the best of our knowledge, there is sTiO2
Particle size: 20/100 nm paste (18NR-AO, Dyesol)
no report yet where (1) the interaction between the hybrid halide
perovskite (MAPbI3) and substrates (TiO2, ZnO and Al2O3) and (2) its Mesoporous ZnO Spin-coated from solution of
eZnO
Particle size: 20/50 nm commercial nanoparticles (Evonik)
influence on the degradation upon moisture exposure have been
studied. In this work, it is shown by FTIR spectroscopy analysis how Mesoporous ZnO Spin-coated from paste obtained
hZnO
the perovskite/ZnO interaction is different with respect to TiO2 and Particle size: 40/100nm from forced hydrolysis method
Al2O3. Taking into account the shift of the position peaks related to
Mesoporous ZnO Spin-coated from solution of
N-H (stretching and bending) and its instability under the presence Particle size: 100/200 nm commercial nanoparticles (Pi-Kem)
pZnO
of acidic groups, we infer that the perovskite interacts with the ZnO
surface through the amine group. In addition, different mechanisms Mesoporous Al2O3 Glancing Angle – Physical Vapor o
Al2O3(65 )
Nanocolumns 65o Deposition (GLAD)
of moisture-induced degradation take place depending on the
nature and morphological properties of substrates. Mesoporous Al2O3 Glancing Angle – Physical Vapor o
Al2O3(85 )
Nanocolumns 85o Deposition (GLAD)

2. Experimental

2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins


Page 3 of 9 Please
Physical do not adjust
Chemistry margins
Chemical Physics
View Article Online
DOI: 10.1039/C6CP01265E
Journal Name ARTICLE

For the perovskite (MAPbI3) synthesis, a 40 wt % solution that


contains a mixture of PbCl2:CH3NH3I (1:3 in molar ratio) using DMF
as solvent was deposited by spin-coating (2000 rpm, 60 s) onto the
different substrates. The film was subsequently annealed at 100 °C
for 1h in a vacuum and under a relative humidity lower than 15%.

Physical Chemistry Chemical Physics Accepted Manuscript


The CH3NH3I precursor was acquired from Dyesol.
The hole-selective material (HSM) solutions were prepared by
dissolving 72.3mg of Spiro-OmeTad (2,2′,7,7′-tetrakis(N,N-di-4-
methoxyphenyamine)-9,9-spirobifluorene) or 20mg of P3HT
(Poly(3-hexylthiophene-2,5-diyl) in 1mL of chlorobenzene. Then 50
μL of each HSM solution were dropped on the substrate and spin
coated at 4000 rpm for 30 s. Figure 1: FTIR spectra of perovskite film deposited on mTiO2, hZnO and Al2O3(85 )
o

-1 -1
substrates in the range of (A) 4000-2500 cm and (B) 1800-800 cm .
2.2 FTIR spectroscopy of thin films
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

The infrared spectra were measured in transmittance mode using a


Bruker IFS 66/S spectrometer equipped with a DTGS detector. The Figure 1 and Table 2 show the FTIR spectra and characteristic
spectra were obtained by averaging 350 scans at a spectral vibrational bands of MAPbI3 deposited on mTiO2, hZnO and
-1 o
resolution of 4 cm and 2.2 KHz as scanner velocity in the 4000-500 Al2O3(85 ) substrates, whose main peaks are related to the
-1 10,36
cm range. Spectra were measured under nitrogen atmosphere and frequency of vibrational modes of the organic cation (see
non-polarized light. All spectra were referenced to the spectrum of Supporting Information, Figure S2). In Figure 1A, the weak and
-1
the bare-Si substrate. broad peak at around 3500 cm is related to O-H stretch vibrations
as a consequence of carrying out the measurements in ambient
2.3 Microstructural characterization 38
conditions. However, the strong peaks between 3300-2900 cm
-1
+
SEM analyses of the samples were performed in S5200 and S4800 are related to the N-H stretch vibrations associated with the NH3
10,36–38,47
microscopes from Hitachi, working at between 2 and 5 kV. group of the MA cation. In particular, this doublet has been
Secondary and backscattered electron (BSE) images were recorded attributed to the symmetric and asymmetric N-H stretch vibrations.
at the same time in the S5200 apparatus giving additional These two latter vibrational modes exhibit the position peaks at
-1 -1
composition information. X-ray diffraction patterns in a Bragg– 3175 cm (asym, v1) and 3130 cm (sym, v2) when the perovskite is
o
Brentano configuration were taken in a D8 Discover diffractometer deposited on mTiO2 and Al2O3(85 ). However, a significant shift of
-1
(Bruker AXS). these peaks (≈25 cm ) was found for hZnO. In particular, the
asymmetric and symmetric vibrational modes were assigned to
-1
peaks at 3205 and 3155 cm , respectively. In contrast, no shift was
3. Results and Discussion found for the asymmetric (v3) and symmetric (v4) C-H stretch
-1 10,36,37
The thin films employed in this work were prepared from different vibrations (2975 and 2932 cm , respectively).
methods and materials obtaining substrates with different Figure 1B shows weaker vibrational modes of MAPbI3 mainly
microstructures and morphologies (see Supporting Information, related to the bending and rocking modes in the range 1800-800
-1
Figure S1). A compact layer of TiO2 was used for cTiO2 substrate, cm . Following Refs. [10, 36, 37], we propose the following
whereas layer consisting of 20 and 20-100 nm of TiO2 nanoparticles attribution of these peaks to the vibrational modes of MAPbI3: v7
-1 + -1
was employed in mTiO2 and sTiO2, respectively. eZnO, hZnO and (1582 cm ) is the asymmetric NH3 bending, v9 (1468 cm ) is the
+ -1
pZnO substrates were prepared using ZnO nanoparticles of 20-50, symmetric NH3 bending, v10 (1422 cm ) is the asymmetric CH3
-1
40-100 and 100-200 nm, respectively. In contrast to these bending, v11 and v13 (1250 and 910 cm , respectively) are the CH3-
o + -1
nanoparticle films, a layer of Al2O3 nanocolumns deposited with 85 NH3 rocking and v12 (960 cm ) is the C-N stretching. In Figure 1B,
o o
and 65 with respect to the source were used as Al2O3(85 ) and the v7, v9, v10 and v11 peaks are observed at the same position when
o o
Al2O3(65 ), respectively. MAPbI3 is deposited on mTiO2, hZnO and Al2O3(85 ) substrates. In
contrast, significant changes appear when hZnO is compared with
o
respect to mTiO2 and Al2O3(85 ). Firstly, v12 and v13 peaks undergo a
-1
similar shift (≈25 cm ) to v1 and v2. Secondly, the v7/v9 ratio is
larger, so that the asymmetric N-H bending seems to be stronger.
Thirdly, the v9 peak changes its shape and an additional (v8) peak
+
becomes visible, also attributed to bending vibrations of the NH3 .
Thereby, bearing in mind that the v5 and v6 peaks were attributed to
37
vibrational coupled modes (v7+v11 and v8/9+v11, respectively), only
the v6 peak shifts. Since, similar results were found for TiO2, ZnO
and Al2O3 substrates of different preparation procedures and
structural properties (see Supporting information, Figure S3), we
infer that the interactions found are attributable to the chemical

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 3

Please do not adjust margins


Please
Physical do not adjust
Chemistry margins
Chemical Physics Page 4 of 9
View Article Online
DOI: 10.1039/C6CP01265E

Journal Name

Physical Chemistry Chemical Physics Accepted Manuscript


ARTICLE

nature of the oxide only, and not to the different morphologies of for ZnO. Thereby, the changes in the FTIR spectra seem to be
the substrate and the perovskite film deposited on it. Thus, it seems mainly due to the different substrates used and that the
+
that the specific interface area between perovskite and substrates, interaction appears to occur through the NH3 group, suggesting
which in fact determine the electron injection and recombination an acid-base reaction. This possible reaction could take place as
rates in complete PSCs, does not affect the frequency of vibrational a consequence of the instability of ZnO under the presence of
49
modes of MAPbI3. acidic group in contrast to TiO2 and Al2 O3. This hypothesis
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

seems to be supported by Figure 2, where a morphological


Table 2: Characteristic frequencies in the IR spectrum of perovskite films change of nanoparticles appears for hZnO substrates only after
the deposition of the perovskite. In particular, Figure 2e shows
Vibration cm-1 Attributed to Affected by
the formation of aciculate ZnO particles in contrast to the
ν1 3175 N-H stretch (asym) ZnO (≈+25 cm-1) spherical form of the hZnO nanoparticles (Figure 2b). The size of
the ZnO particles has also considerably increased after the
ν2 3130 N-H stretch (sym) ZnO (≈+25 cm-1) deposition from 100 to 200 nm in diameter to lengths over 500
nm (see Supporting Information, Figure S4). These observations
ν3 2975 C-H stretch (asym) - confirm the existence of a chemical reaction between perovskite
and ZnO that even produces a re-dissolution of the ZnO
ν4 2932 C-H stretch (sym) - substrate and generation of a new morphology.

ν5 2828 v 7+v11 -

-1
ν6 2702 v 8/9+v11 ZnO (≈+25 cm )

+
ν7 1582 NH3 bending (asym) -

+
ν8 1484 NH3 bending (sym) Only appears for ZnO

+
ν9 1468 NH3 bending (sym) -

ν10 1422 CH3 bending (asym) -

ν11 1250 CH3-NH3+ rocking -

ν12 960 C-N stretch ZnO (≈+25 cm-1)

ν13 950 CH3-NH3+ rocking ZnO (≈+25 cm-1)

A similar displacement of peaks was also found previously by


37
Glaser when different halides (X= Cl, Br and I) were used.
These changes were tentatively explained by the Lorentz-Lorentz
shift consequence of the different interactions between the MA
and Pb-X. However, in our case the same perovskite (MAPbI3) Figure 2: SEM of mTiO2 (top), hZnO (middle) and Al2O3(85o) (bottom) substrates
has been used and not all peaks undergo a shift. As it was before (left) and after (right) deposition of perovskite layer.
mentioned above, only (asym and sym) N-H stretch, (sym) N-H
bend and C-N rock vibrational modes of MAPbI3 are displaced It is well-known that moisture exposure shows both detrimental
48
when ZnO was used as substrate. In particular, Bellanato et al. and beneficial effects on perovskite materials. For instance, during
+
showed that the interaction between NH3 group and the the perovskite processing small amounts of water can improve the
support materials can produce not only the displacement of the optoelectronic properties, and hence the efficiency of the
9,38,50,51
position, but also a variation in the intensity of the peaks or even devices. Nevertheless, its prolonged exposure produces the
the appearance of new peaks, which could explain our results formation of the colorless hydrated CH3NH3PbI3·H2O or

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 4

Please do not adjust margins


Page 5 of 9 Please
Physical do not adjust
Chemistry margins
Chemical Physics
View Article Online
DOI: 10.1039/C6CP01265E
Journal Name ARTICLE

(CH3NH3)4PbI6·2H2O with negative consequences for its photovoltaic to the influence of the nature of the substrate itself on the
52–55
performance. In order to study the interaction between water degradation of the perovskite. To investigate the effects of water
molecules and perovskite materials, the optoelectronic properties, molecules on perovskite, the samples were stored under humid
perovskite morphology and crystal structure were analyzed in conditions (85% RH) for different times (see Scheme 1 in the
numerous papers. However, in this paper attention has been paid Supporting Information) at room temperature and in darkness.

Physical Chemistry Chemical Physics Accepted Manuscript


Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

Figure 3: (A) FTIR spectra of perovskite film deposited on mTiO2 films before (red line) and after (blue line) moisture exposure for 1, 3 and 5 hours in the range of 4000-750 cm-1. (B)
FTIR spectra of perovskite film deposited on hZnO and Al2O3(85o) before (red line) and after (blue line) moisture exposure for 3 hours in the range of 4000-600 cm-1. Labels indicate
the fraction of signal intensity that remains after moisture exposure.

Figure 3A shows the FTIR spectra of perovskite deposited on the the perovskite and water molecules. For both Al2O3 and ZnO films
mTiO2 layer before and after the moisture exposure for 1, 3 and 5 substrates, the spectral changes after moisture exposure show the
hours. Müller et al.38 proposed a molecular model in which the same behavior than for TiO2 in the sense that the intensity of the
water molecules modify the hydrogen bonds between MA and main peaks (v1, v2, v9 and v13) decrease. However, several
halide atoms forming new hydrogen bonds. Thereby, a shift and differences can be observed. In the case of Al2O3 substrates, it
+
decrease of the asymmetric NH3 stretch vibration mode (v1) was seems that this oxide promotes the inclusion of water droplets as
found. In contrast to this result, Figure 3A shows a decrease of the suggested by the behavior of the broad peak associated to the O-H
-1
whole FTIR spectrum after moisture exposure where the peaks stretching (≈3500 cm ), which increases after the exposure. This is
appear at the same wavenumber. The decrease of the most intense sort of expected as alumina is a hygroscopic material that tends to
absorption bands (v1, v2, v9 and v13) is stronger for longer exposition capture ambient moisture.
-1
times under moisture. For instance, the signals corresponding to v1 When ZnO was used as substrate, the broad 3500 cm peak gives
-1
and v2 vibrational modes are reduced to 77% and 65% of their initial place to two well-defined peaks at 3500 and 3450 cm , associated
38
values after 1 and 5 hours, respectively. Since these modes to O-H vibrations in separated water molecules. In contrast to
correspond to the N-H stretching vibrations, this observation Al2O3, this observation indicates that isolated water molecules
strongly suggests that water molecules interact again with the percolate within the structure in the case of ZnO. Additionally, v13
+
perovskite via the ammonium group. The behavior of the v9 peak peak, which was assigned to CH3-NH3 rocking vibrational mode,
-1 -1
(N-H bending) corroborates this interpretation. The formation of shifts from 925 to 910 cm and a new peak appears at 720 cm .
hydrates likely modifies the relative concentration of these groups Similar results were found for pZnO and eZnO (see Supporting
in the sample. In addition, bearing in mind that the perovskite Information, Figure S6). Nevertheless, a new additional peak at
-1
degradation is carried out by water surface adsorption and 1045 cm appears after moisture exposure for both pZnO and eZnO
38,56
infiltration processes, it has been found that degradation with respect to hZnO. This difference could be related to the
process also depends on the nanostructure of substrates (see different synthesis methods of ZnO, which determine the number
Supporting Information, Figure S5). The general observation is that of oxygen vacancies and then its reactivity. For instance, a clear
49
the degree of degradation increases in the order sTiO2 > cTiO2 > difference was reported by Dong et al. during the perovskite
mTiO2. However, no direct relation between particle size and formation process when ZnO films obtained by ALD and solution
degradation rate is detected. methods were employed as substrates. In any case, all these
On the other hand, it has been found that the perovskite observations clearly indicate that the ZnO substrate induces a
degradation also depends on the nature of substrates. Figure 3B chemical change of the perovskite in the presence of moisture, and
o
shows FTIR spectra of perovskite deposited on hZnO and Al2O3(85 ) that this change is associated to the chemical nature of the oxide
layers before and after moisture exposure to study the influence of rather to the structural properties such as surface area, particle size
the chemical nature of the substrate on the interaction between or porosity. However, the fact that the perovskite tends to react

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 5

Please do not adjust margins


Please
Physical do not adjust
Chemistry margins
Chemical Physics Page 6 of 9
View Article Online
DOI: 10.1039/C6CP01265E
ARTICLE Journal Name

with the ZnO, and change its morphology, can also contribute to an characteristic peaks associated with PbI2 appear neither before
enhanced percolation of water molecules into the interface. or after moisture exposure. This result brings to light the specific
The peculiar results obtained when ZnO is used as substrate, both interaction that takes place between perovskite material and
with and without moisture, highlight the importance of specific substrates and it seems to point at different perovskite
interactions between perovskite and ZnO. In this regard, Zhang et formation and degradation processes, involving the

Physical Chemistry Chemical Physics Accepted Manuscript


42
al. claimed recently that the following decomposition is catalyzed decomposition of the perovskite in accordance to the reaction
by ZnO indicated above.
CH3NH3PbI3 CH3NH2 + PbI2 + HI

The formation of the amino group favors the acid-base reaction


+
with water, which explains why all peaks related to the NH3 group
become affected. The PbI2 compound cannot be detected in the
studied FTIR range (4000 - 500 cm-1).36 However, in Figure 4 the
yellowish coloration of the film after water exposure suggest that
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

PbI2 has formed in the presence of ZnO.

Figure 4: Images of perovskite film deposited on (A, B) mTiO2 and (C, D) hZnO layers (A,
C) before and (B, D) after moisture exposure for 5 hours

In order to analyze the influence of humidity on the crystal


Figure 5: X-ray diffraction for perovskite deposited on (A) mTiO2, (B) hZnO and (C)
structure of perovskite the XRD patterns were recorded. Figure 5 o
Al2O3(85 ) substrates (red line) before and (blue line) after moisture exposure for 5
shows the XRD patterns of perovskite deposited on mTiO2 , hZnO hours.
o
and Al2O3(85 ) substrates before and after moisture exposure
(85% RH , 5h). Characteristic diffraction peaks of perovskite
appear when the MAPbI3 layer was deposited on all the different Taking into account the effects of prolonged moisture exposure
substrates. For instance, the stronger diffraction peak, related to on photoelectronic properties of PSCs, special attention has
the (110), (220) and (310) crystal planes of tetragonal been paid to ensure the long-term device stability, which is one
o o o of the most critical issues for its commercialization. Different
perovskite, appear at 14.1 , 28.5 and 31.9 , respectively. Similar
42,51–53 strategies have been carried out to guarantee the stability of the
results were previously reported. When the MAPbI3 layer
57–60
was subjected under moisture exposure, the characteristic perovskite films. At this point, it is important to remark that
diffraction peaks remained at the same positions. In spite of the the HSMs does not only collect the photogenerated holes, but
52
decrease of the signal intensity, the persistence of these also protects the underlying perovskite films against humidity.
diffraction peaks guarantees the presence of crystalline MAPbI3. In order to determine how the FTIR technique can help to assess
Nevertheless, there are some observable changes between the how the nature of HSM affects the perovskite degradation
XRD patterns of perovskite before and after moisture exposure process, we have performed FTIR measurements of perovskite
in relation to the nature of substrates. For instance, when films in combination with two of the most employed HSMs.
MAPbI3 layer was deposited on mTiO2, the diffraction peaks only Figure 6 shows the FTIR spectra of MAPbI3 deposited on mTiO2 and
suffer a decrease of signal intensity. However, additional hZnO films before and after moisture exposure (RH 85%, 25h) with
diffraction peaks appear when MAPbI3 layer was deposited on Spiro-OmeTad and P3HT as HSMs and the uncovered perovskite
hZnO not only after moisture exposure, but also before. The film, which was used as control. It is important to state that the
diffraction peaks at 12.7o, 25.95o, 38.7o and 52.48o were presence of the HSM does not modify the FTIR spectra of the
attributed to the presence of crystalline PbI2.42,52,53 In particular, perovskite, whose peaks appear at the same wavenumber. The
similar results were found by Zang et al. as a consequence of the combination of both materials produces the superposition of FTIR
decomposition of MAPbI3 due to a long annealing time.42 In any spectra of perovskite and HSMs. The FTIR spectra of both HSMs are
case, after moisture exposure the signal of diffraction peak shown in Supporting Information (see Figure S7). For both mTiO2
associated to the crystalline MAPbI3 decrease, whereas the and hZnO substrates, the FTIR spectral changes after moisture
intensity of peaks attributed to PbI2 increase. A similar result exposure show the same trends described above independently of
was found when the MAPbI3 layer was deposited on Al2O3(85o) which HSM has been deposited. However, the signal of the peaks
substrate after moisture exposure. However, in this case no associated with the HSMs does not undergo any changes. This

6 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins


Page 7 of 9 Please
Physical do not adjust
Chemistry margins
Chemical Physics
View Article Online
DOI: 10.1039/C6CP01265E

Journal Name

Physical Chemistry Chemical Physics Accepted Manuscript


ARTICLE
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

Figure 6: FTIR spectra of bare perovskite film and in combination with HSMs before (red line) and after (blue line) moisture exposure for 25 hours deposited on (A) mTiO2 layers in
the range of 4000-750 cm-1 and (B) hZnO layers in the range of 4000-650 cm-1.

effect confirms that the decrease of the signal intensity of the peaks In relation to the chemical nature of substrates, comparison of FTIR
related to the vibrational modes of perovskite is a consequence of spectra of perovskite deposited on ZnO and the rest of the
its degradation. In order to determine the influence of HSMs on the materials (TiO2 an Al2O3) shows that the peaks corresponding to N-
-1
perovskite degradation process, we have analyzed the decrease of H vibrational modes become shifted (≈25 cm ), which suggests the
v1 and v2 vibrational modes intensities of MAPbI3. When mTiO2 film existence of a specific interaction between perovskite and ZnO. This
was used as substrate, these peaks are reduced to 82% and 92% of observation is common to the three types of ZnO materials studied
+
the original signal in combination with Spiro-OmeTad and P3HT, and points out that this interaction is driven by the NH3 group of
respectively. However, for the uncovered perovskite the signal is the organic cation. This specific interaction appears to be also
reduced to 47%. Even for the case of hZnO, for which as already behind the stronger perovskite degradation by moisture in the
discussed the degradation is faster, the same trend is maintained. presence of ZnO. SEM images show that the perovskite does even
These results infer that P3HT acts as more efficient barrier than induce a change of the morphology of the ZnO substrate after
Spiro-OmeTad to prevent the perovskite degradation upon deposition. Furthermore, FTIR spectra and XRD measurements
humidity making more difficult the penetration/percolation of strongly suggest that this oxide favors the degradation of the
water molecules. Similar results were previously reported using UV- perovskite and the formation of PbI2. In addition, it is demonstrated
Vis spectra as testing tool.52 that the morphological properties of substrates and the presence of
HSM also play an important role in the degradation rate of
perovskite. In particular, a slower degradation rate was found when
4. Conclusions P3HT was used as HSM.
In this report, we have performed an experimental analysis of the The results reported here that the interaction between perovskite
vibrational properties of hybrid organic-inorganic halide perovskite and selective contacts is crucial in the formation and stability of an
in the presence of various substrates used as selective layers in interface suited for photovoltaic performance. In fact, the particular
perovskite-based solar cells. In particular, we studied the influence chemical nature of perovskite and metal oxide, as well as the
of the chemical nature and morphological properties of the morphology of the latter, can lead to a dramatic impact on the
substrates on the perovskite formation and the influence of these stability of the active film. On the other hand, the FTIR technique
substrates and HSMs on the degradation process under moisture shows to be a simple and effective technique to detect this kind of
exposure. interactions.

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 7

Please do not adjust margins


Please
Physical do not adjust
Chemistry margins
Chemical Physics Page 8 of 9
View Article Online
DOI: 10.1039/C6CP01265E
ARTICLE Journal Name

Acknowledgements 27 R. E. Wasylishen, O. Knop and J. B. Macdonald, Solid State


Commun., 1985, 56, 581 – 582.
We thank Regional Goverment of Andalusia for financial 28 N. J. Jeon, J. H. Noh, W. S. Yang, Y. C. Kim, S. Ryu, J. Seo and S. I.
support via grants FQM1851 and FQM2310 and Ministerio de Seok, Nature, 2015, 517, 476–480.
Economia y Competitividad under grant MAT2013-47192-C3-3- 29 L. K. Ono, S. R. Raga, S. Wang, Y. Kato and Y. Qi, J. Mater. Chem.

Physical Chemistry Chemical Physics Accepted Manuscript


R and MAT2013-42900-P. A, 2015, 3, 9074–9080.
30 G. E. Eperon, S. D. Stranks, C. Menelaou, M. B. Johnston, L. M.
Herz and H. J. Snaith, Energy Environ. Sci., 2014, 7, 982–988.
Notes and references 31 C. Quarti, G. Grancini, E. Mosconi, P. Bruno, J. M. Ball, M. M.
Lee, H. J. Snaith, A. Petrozza and F. D. Angelis, J. Phys. Chem.
1 B. O’Regan and M. Gratzel, Nature, 1991, 353, 737–740. Lett., 2014, 5, 279–284.
2 M. Grätzel, Acc. Chem. Res., 2009, 42, 1788–1798.
32 R. Gottesman, L. Gouda, B. S. Kalanoor, E. Haltzi, S. Tirosh, E.
3 A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Rosh-Hodesh, Y. Tischler, A. Zaban, C. Quarti, E. Mosconi and F.
Soc., 2009, 131, 6050–6051.
D. Angelis, J. Phys. Chem. Lett., 2015, 6, 2332–2338.
4 J.-H. Im, C.-R. Lee, J.-W. Lee, S.-W. Park and N.-G. Park, 33 M. Ledinský, P. Löper, B. Niesen, J. Holovský, S.-J. Moon, J.-H.
Nanoscale, 2011, 3, 4088–4093.
Yum, S. D. Wolf, A. Fejfar and C. Ballif, J. Phys. Chem. Lett., 2015,
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

5 U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissortel, J. Salbeck,


6, 401–406.
H. Spreitzer and M. Gratzel, Nature, 1998, 395, 583–585.
34 N. Onoda-Yamamuro, T. Matsuo and H. Suga, J. Phys. Chem.
6 J. M. Ball, M. M. Lee, A. Hey and H. Snaith, Energy Environ. Sci.,
Solids, 1990, 51, 1383 – 1395.
2013.
35 E. Mosconi, C. Quarti, T. Ivanovska, G. Ruani and F. De Angelis,
7 M. Liu, M. B. Johnston and H. J. Snaith, Nature, 2013, 501, 395–
Phys Chem Chem Phys, 2014, 16, 16137–16144.
398.
36 M. A. Pérez-Osorio, R. L. Milot, M. R. Filip, J. B. Patel, L. M. Herz,
8 J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M.
M. B. Johnston and F. Giustino, J. Phys. Chem. C, 2015, 119,
K. Nazeeruddin and M. Grätzel, Nature, 2013, 499, 316–319.
25703–25718.
9 H. Zhou, Q. Chen, G. Li, S. Luo, T. Song, H.-S. Duan, Z. Hong, J.
37 T. Glaser, C. Müller, M. Sendner, C. Krekeler, O. E. Semonin, T. D.
You, Y. Liu and Y. Yang, Science, 2014, 345, 542–546.
Hull, O. Yaffe, J. S. Owen, W. Kowalsky, A. Pucci and R. Lovrinčić,
10 N. J. Jeon, J. H. Noh, Y. C. Kim, W. S. Yang, S. Ryu and S. I. Seok,
J. Phys. Chem. Lett., 2015, 6, 2913–2918.
Nat Mater, 2014, 13, 897–903.
38 C. Müller, T. Glaser, M. Plogmeyer, M. Sendner, S. Döring, A. A.
11 N. Ahn, D.-Y. Son, I.-H. Jang, S. M. Kang, M. Choi and N.-G. Park,
Bakulin, C. Brzuska, R. Scheer, M. S. Pshenichnikov, W.
J. Am. Chem. Soc., 2015, 137, 8696–8699.
Kowalsky, A. Pucci and R. Lovrinčić, Chem. Mater., 2015, 27,
12 M. A. Green, K. Emery, Y. Hishikawa, W. Warta and E. D. Dunlop,
7835–7841.
Prog. Photovolt. Res. Appl., 2015, 23, 805–812.
39 A. A. Bakulin, O. Selig, H. J. Bakker, Y. L. A. Rezus, C. Müller, T.
13 D. Bi, G. Boschloo, S. Schwarzmuller, L. Yang, E. M. J. Johansson
Glaser, R. Lovrincic, Z. Sun, Z. Chen, A. Walsh, J. M. Frost and T.
and A. Hagfeldt, Nanoscale, 2013, 5, 11686–11691.
L. C. Jansen, J. Phys. Chem. Lett., 2015, 6, 3663–3669.
14 J. Dong, Y. Zhao, J. Shi, H. Wei, J. Xiao, X. Xu, J. Luo, J. Xu, D. Li, Y.
40 A. Listorti, E. J. Juarez-Perez, C. Frontera, V. Roiati, L. Garcia-
Luo and Q. Meng, Chem. Commun., 2014, 50, 13381–13384.
Andrade, S. Colella, A. Rizzo, P. Ortiz and I. Mora-Sero, J. Phys.
15 D. Liu, M. K. Gangishetty and T. L. Kelly, J. Mater. Chem. A, 2014,
Chem. Lett., 2015, 6, 1628–1637.
2, 19873–19881.
41 A. Matas Adams, J. M. Marin-Beloqui, G. Stoica and E.
16 K. Mahmood, B. S. Swain and H. S. Jung, Nanoscale, 2014, 6,
Palomares, J Mater Chem A, 2015, 3, 22154–22161.
9127–9138.
42 J. Zhang and T. Pauporté, J. Phys. Chem. C, 2015, 119, 14919–
17 M. H. Kumar, N. Yantara, S. Dharani, M. Graetzel, S. Mhaisalkar,
14928.
P. P. Boix and N. Mathews, Chem. Commun., 2013, 49, 11089–
43 H.-S. Kim, I.-H. Jang, N. Ahn, M. Choi, A. Guerrero, J. Bisquert
11091.
and N.-G. Park, J. Phys. Chem. Lett., 2015, 6, 4633–4639.
18 J. Zhang, P. Barboux and T. Pauporté, Adv. Energy Mater., 2014,
44 A. G. Vega-Poot, M. Macías-Montero, J. Idígoras, A. Borrás, A.
4, n/a–n/a.
Barranco, A. R. Gonzalez-Elipe, F. I. Lizama-Tzec, G. Oskam and J.
19 J. Zhang, E. J. Juarez-Perez, I. Mora-Sero, B. Viana and T.
A. Anta, ChemPhysChem, 2014, 15, 1088–1097.
Pauporte, J. Mater. Chem. A, 2015, 3, 4909–4915.
45 L. González-García, A. Barranco, A. M. Páez, A. R. González-Elipe,
20 Z.-L. Tseng, C.-H. Chiang and C.-G. Wu, Sci. Rep., 2015, 5, 13211. M.-C. García-Gutiérrez, J. J. Hernández, D. R. Rueda, T. A.
21 D.-Y. Son, J.-H. Im, H.-S. Kim and N.-G. Park, J. Phys. Chem. C,
Ezquerra and D. Babonneau, ChemPhysChem, 2010, 11, 2205–
2014, 118, 16567–16573. 2208.
22 O. Malinkiewicz, C. Roldán-Carmona, A. Soriano, E. Bandiello, L.
46 A. Barranco, A. Borras, A. R. Gonzalez-Elipe and A. Palmero,
Camacho, M. K. Nazeeruddin and H. J. Bolink, Adv. Energy Prog. Mater. Sci., 2016, 76, 59–153.
Mater., 2014, 4, n/a–n/a.
47 T. Ma, M. Cagnoni, D. Tadaki, A. Hirano-Iwata and M. Niwano, J.
23 A. Guerrero, J. You, C. Aranda, Y. S. Kang, G. Garcia-Belmonte, H. Mater. Chem. A, 2015, 3, 14195–14201.
Zhou, J. Bisquert and Y. Yang, ACS Nano, 2015.
48 J. Bellanato, Spectrochim. Acta, 1960, 16, 1344–1357.
24 D. B. Mitzi, K. Chondroudis and C. R. Kagan, IBM J Res Dev, 2001, 49 X. Dong, H. Hu, B. Lin, J. Ding and N. Yuan, Chem. Commun.,
45, 29–45.
2014, 50, 14405–14408.
25 K. T. Butler, J. M. Frost and A. Walsh, Mater. Horiz., 2015, 2, 50 S. Pathak, A. Sepe, A. Sadhanala, F. Deschler, A. Haghighirad, N.
228–231.
Sakai, K. C. Goedel, S. D. Stranks, N. Noel, M. Price, S. Hüttner,
26 J. M. Frost, K. T. Butler and A. Walsh, APL Mater, 2014, 2, N. A. Hawkins, R. H. Friend, U. Steiner and H. J. Snaith, ACS
081506.
Nano, 2015, 9, 2311–2320.

8 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins


Page 9 of 9 Please
Physical do not adjust
Chemistry margins
Chemical Physics
View Article Online
DOI: 10.1039/C6CP01265E
Journal Name ARTICLE

51 M. K. Gangishetty, R. W. J. Scott and T. L. Kelly, Nanoscale, 2016,


-.
52 J. Yang, B. D. Siempelkamp, D. Liu and T. L. Kelly, ACS Nano,
2015, 9, 1955–1963.
53 J. A. Christians, P. A. Miranda Herrera and P. V. Kamat, J. Am.

Physical Chemistry Chemical Physics Accepted Manuscript


Chem. Soc., 2015, 137, 1530–1538.
54 B. Hailegnaw, S. Kirmayer, E. Edri, G. Hodes and D. Cahen, J.
Phys. Chem. Lett., 2015, 6, 1543–1547.
55 A. Halder, D. Choudhury, S. Ghosh, A. S. Subbiah and S. K.
Sarkar, J. Phys. Chem. Lett., 2015, 6, 3180–3184.
56 E. Mosconi, J. M. Azpiroz and F. De Angelis, Chem. Mater., 2015,
27, 4885–4892.
57 S. N. Habisreutinger, T. Leijtens, G. E. Eperon, S. D. Stranks, R. J.
Nicholas and H. J. Snaith, Nano Lett., 2014, 14, 5561–5568.
58 I. C. Smith, E. T. Hoke, D. Solis-Ibarra, M. D. McGehee and H. I.
Karunadasa, Angew. Chem. Int. Ed Engl., 2014, 53, 11232–
Published on 12 April 2016. Downloaded by test 3 on 16/04/2016 09:20:48.

11235.
59 Y. Han, S. Meyer, Y. Dkhissi, K. Weber, J. M. Pringle, U. Bach, L.
Spiccia and Y.-B. Cheng, J Mater Chem A, 2015, 3, 8139–8147.
60 G. Niu, W. Li, F. Meng, L. Wang, H. Dong and Y. Qiu, J Mater
Chem A, 2014, 2, 705–710.

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 9

Please do not adjust margins

You might also like