Sie sind auf Seite 1von 12

Reports

arsenide, GaAs, solar cells exhibit Voc


of 1.11 V and an optical band gap of
1.4 eV giving a difference of approximately 0.29 eV (2). For dye-sensitized
and organic solar cells, this difference
is usually on the order of 0.7 to 0.8 eV
(2, 9). For organic solar cells, such
losses are predominantly caused by
their low dielectric constants. Tightly1
1
2
2,3
Michael M. Lee, Jol Teuscher, Tsutomu Miyasaka, Takurou N. Murakami,
bound excitons form that require a
1
Henry J. Snaith *
heterojunction with an electron accep1
2
tor with a significant energy offset to
Clarendon Laboratory, Department of Physics, University of Oxford, Oxford OX1 3PU, UK. Graduate
enable ionization and charge separation
School of Engineering, Toin University of Yokohama, 1614 Kurogane, Aoba, Yokohama 225-8503, Japan.
3
Research Center for Photovoltaic Technologies, National Institute of Advanced Industrial Science and
(10, 11). Likewise, dye-sensitized solar
Technology, Central 5, 1-1-1 Higashi, Tsukuba, Ibaraki 305-8565, Japan.
cells (DSSCs) have losses, both from
electron transfer from the dye (or ab*To whom correspondence should be addressed. E-mail: h.snaith1@physics.ox.ac.uk
sorber) into the TiO2 which requires a
certain driving force and from dye
The energy cost associated with separating tightly-bound excitons, photoregeneration from the electrolyte which
generated electron-hole pairs, and extracting free charges from highly disordered
requires an over potential. Efforts
low mobility networks represent fundamental losses for many low-cost photovoltaic
have been made to reduce such losses
technologies. We report a low-cost, solution-processable solar cell based on a
in the DSC by moving from a multihighly crystalline perovskite absorber with intense visible-to-near-infrared
electron iodide/tri-iodide redox couple
absorptivity that has a power conversion efficiency of 10.9% in a single junction
to one-electron outer-sphere redox
device under simulated full sunlight. This meso-superstructured solar cell (MSSC)
couples, such as a cobalt complexes or
exhibits exceptionally few fundamental energy losses illustrated by generating
a solid-state hole-conductor (1, 4, 12,
open-circuit photovoltages of over 1.1 volts, despite the relatively narrow absorber
13).
band gap of 1.55 electron volts. The functionality arises from the use of
Recently, there has been mounting
mesoporous alumina as an inert scaffold which structures the absorber and forces
interest
in inorganic semiconductor
electrons to reside in and be transported through the perovskite.
sensitized solar cells, (14, 15). An
extremely thin absorber (ETA) layer, 2
to 10 nm in thickness is coated upon
An efficient solar cell must absorb over a broad spectral range, from the the internal surface of a mesoporous TiO2 electrode and then contacted
visible to near infrared region (350 to ~ 950 nm), and convert the inci- with an electrolyte or solid-state hole-conductor. These devices have
dent light effectively into charges. The charges must be collected at a achieved power conversion efficiencies of up to 6.3% (15). However, the
high voltage with suitable current in order to do useful work (18). A ETA concept suffers from rather low Voc ; the problem may lie in the
simple measure of solar cell effectiveness at generating voltage is the electronically disordered, low mobility n-type TiO2 (16). Perovskites are
difference in energy between the optical band gap of the absorber and relatively underexthe open-circuit voltage (Voc) generated by the solar cell under simulated plored alternatives
Fig. 1. (A) Three-dimensional schematic
AM1.5G 100 mW cm2 solar illumination (9). For instance, gallium (Fig. 1A) that prorepresentation of perovskite structure,
ABX3 where A = CH3NH3, B = Pb, and X
= Cl, I (left) with two-dimensional
schematic illustrating the perovskite unit
cell (right). (B) UV-Vis absorbance
spectra of the photoactive layer in the
solar cell (mesoporous oxide; perovskite
absorber;
spiro-OMeTAD)
sealed
between two sheets of glass in nitrogen
and exposed to simulated AM1.5G
2
sunlight at 100 mW cm irradiance for up
to 1000 hours. No additional UV filtration
was used for the solar irradiance. Inset is
the extracted optical density at 500 nm as
a function of time. (C) Schematic
representation of full device structure,
where the mesoporous oxide is either
Al2O3 or anatase TiO2 (left), Crosssectional SEM image of a full device
incorporating mesoporous Al2O3 (right).
Scale bar represents 500 nm. SEM
measurements were taken using JEOL
JSM 7500F.

/ http://www.sciencemag.org/content/early/recent / 4 October 2012 / Page 1/ 10.1126/science.1228604

Downloaded from www.sciencemag.org on October 13, 2012

Efficient Hybrid Solar Cells Based on


Meso-Superstructured Organometal
Halide Perovskites

vide a framework for binding the organic and inorganic components into
a molecular composite. Given careful consideration of the interaction
between organic-inorganic elements and by controlling the size-tunable
crystal cell (17) it is possible to create new and interesting materials
using rudimentary wet chemistry. Era, Mitzi, and co-workers have
shown that layered perovskites based on organometal halides demonstrate excellent performance as light-emitting diodes (18, 19) and transistors with mobilities comparable to amorphous silicon (20).
Organometal halide perovskites have been used as sensitizers in liquid
electrolyte based photo-electrochemical cells with conversion efficiencies between 3.5 to 6.5% (21, 22). Recently, a CsSnI3 perovskite has
been shown to function efficiently as a hole-conductor in solid-state dyesensitized solar cells, delivering up to 8.5% power conversion efficiency
(23, 24).
We report on a solution-processable solar cell that overcomes the
fundamental losses of organic absorbers and disordered metal oxides.
We followed the ETA approach and used a perovskite absorber and
mesoporous TiO2 as the transparent n-type component, and 2,2(7,7(tetrakis-(N,N-di-pmethoxyphenylamine)9,9(-spirobifluorene))) (spiroOMeTAD) as the transparent p-type hole conductor. These devices exhibited power conversion efficiencies near 8%. Remarkably, we also
found replacing the mesoporous n-type TiO2 with insulating Al2O3, improved the power conversion efficiency. The Al2O3 is a wide band gap
(7 to 9eV) insulator and purely acts as a scaffold upon which the
perovskite is coated. We observe that electron transport through the
perovskite layer is much faster than through the n-type TiO2. In addition,
we observe a few hundred millivolt increase in Voc moving from the
TiO2 to the insulating Al2O3 scaffold, and a record power conversion
efficiency of 10.9% under simulated Air Mass (AM)1.5 full sun illumination.
The specific perovskite we used is of mixed-halide form:
methylammonium lead iodide chloride, (CH3NH3PbI2Cl), which was
processed from a precursor solution in N,N-dimethylformamide via spincoating in ambient conditions. X-ray diffraction analysis for
CH3NH3PbI2Cl prepared on glass (fig. S1) (25), showed diffraction
peaks at 14.20, 28.58, 43.27 we assigned as the (110), (220), and (330)
planes, respectively, of a tetragonal perovskite structure with lattice parameters a = 8.825 , b = 8.835 , c = 11.24 , similar to the
CH3NH3PbI3 previously reported (21). The extremely narrow diffraction

peaks suggest that the films have long range crystalline domains (greater
than 200 nm, peak width limited by instrument broadening) and are
highly oriented with the a-axis (21, 26). In contrast to the
methylammonium tri-halogen plumbates previously reported in solar
cells (i.e., CH3NH3PbI3) (21, 22), this iodide-chloride mixed-halide
perovskite was remarkably stable to processing in air. In Fig. 1B, the
absorption spectra demonstrated good light harvesting capabilities over
the visible to near-infrared (near-IR) spectrum, and was also stable to
prolonged light exposure, as demonstrated by 1000 hours constant illumination under full sunlight. Negligible change in absorbance is illustrated by the inset to Fig. 1B, where the absorbance of the film at 500 nm
remained around 1.8 throughout the entire measurement period (absorbance of 1.8 corresponds to 98.4% absorbance). We note that the scale is
optical density, where absorbance of ~ 0.5 at 700 nm corresponds to
around 70% attenuation in a single pass, and in the solar cell, there are
two passes of light leading to approximately 91% absorbance at this
wavelength.
The solar cells were fabricated on semi-transparent fluorine doped
tin oxide (FTO) coated glass, coated with a compact layer of TiO2 that
acts as an anode. The porous oxide films were fabricated from sol-gel
processed sintered nanoparticles. The perovskite precursor solution was
infiltrated into the porous oxide mesostructure via spin-coating, and was
dried at 100C which enabled the perovskite to form via self-assembly of
the constituent ions. Dark coloration was only observed after this drying
step.
To elaborate upon the perovskite coating process, there has been extensive work done on investigating how solution-cast materials infiltrate
into mesoporous oxides (2732). If the concentration of the solution is
low enough, and the solubility of the cast material high enough, the material will completely penetrate the pores as the solvent evaporates. Typically, the material forms a wetting layer upon the internal surface of
the mesoporous film that uniformly coats the pore walls throughout the
thickness of the electrode (2831). The degree of pore-filling can be
controlled by varying the solution concentration (2932). If the concentration of the casting solution is high, then maximum pore filling occurs,
and any excess material forms a capping layer on top of the filled
mesoporous oxide.
For the optimum perovskite precursors concentrations we used, there
was no appearance of a capping layer, which implies that the perovskite

/ http://www.sciencemag.org/content/early/recent / 4 October 2012 / Page 2/ 10.1126/science.1228604

Downloaded from www.sciencemag.org on October 13, 2012

Fig. 2. (A) IPCE action spectrum of an Al2O3 based and perovskite sensitized TiO2 solar cell, with device structure: FTO/CompactTiO2/Mesoporous Al2O3 (red trace with crosses) or mesoporous TiO2 (black trace with circles)/ CH3NH3PbI2Cl/Spiro-OMeTAD/Ag
2
(B) Current-Voltage characteristics under simulated AM1.5G 100 mW cm illumination for Al2O3 based cells, one cell exhibiting
high efficiency (red solid trace with crosses) and one exhibiting greater than 1.1 V VOC (red dashed line with crosses), perovskite
TiO2 sensitized solar cell (black trace with circles), and a planar-junction diode with a structure FTO/compact TiO2/
CH3NH3PbI2Cl/Spiro-OMeTAD/Ag (purple trace with squares).

was predominantly formed within the mesoporous film. We verified that


the perovskite was within and uniformly distributed throughout the
mesoporous oxide films by performing cross-sectional scanning electron

/ http://www.sciencemag.org/content/early/recent / 4 October 2012 / Page 3/ 10.1126/science.1228604

Downloaded from www.sciencemag.org on October 13, 2012

Fig. 3. (A) Cartoon illustrating the charge transfer and


transport in a perovskite sensitized TiO2 solar cell, (left) and a
non-injecting Al2O3 based perovskite solar cell, (right). (B)
Photo-induced absorbance (PIA) spectra of the mesoporous
TiO2 (black open circles) and Al2O3 (red crosses) films coated
with perovskite with (solid lines) and without (dashed lines)
spiro-OMeTAD hole transporter. ex = 496.5 nm, repetition rate
23 Hz. (C) Charge transport lifetime determined by small
perturbation transient photocurrent decay of perovskite
sensitized TiO2 (circles with black line to aid the eye) and
Al2O3 cells (red crosses with line to aid the eye). Inset shows
normalised photocurrent transients for Al2O3 (red trace with
th
crosses every 7 point) and TiO2 (black trace with circles
th
2
every 7 point) cells, set to generate 5 mA cm photocurrent
from the background light bias.

microscopy (SEM) with elemental mapping via energy dispersive x-ray


(EDX) analysis (fig. S2) (25). To complete the photoactive layer, the
perovskite coated porous electrode was further filled with the holetransporter, spiro-OMeTAD, via spin-coating (Fig. 1C). It is apparent in
Fig. 1C, that the spiro-OMeTAD has formed a capping layer, which
ensure selective collection of holes at the silver electrode.
In Fig. 2A, the incident photon-to-electron conversion efficiency
(IPCE) action spectrum is shown for the devices that use mesoporous
TiO2 and Al2O3, exhibiting spectral sensitivity spanning from the visible
to the near-IR (400 to 800 nm) with a peak IPCE of over 80% for both
oxides. The slight difference in shape arose from the slightly different
perovskite concentrations in the optimized devices. In Fig. 2B, we show
current-voltage (J-V) curves measured under simulated AM1.5 100
mW/cm2 illumination. The sensitized TiO2 solar cell exhibited a shortcircuit photocurrent (Jsc) of 17.8 mA cm2, a Voc of 0.80 V with a fill
factor of 0.53 yielding an overall power conversion efficiency () of
7.6%. We present two different J-V curves for the Al2O3 based device.
The most efficient device exhibited a Jsc of 17.8 mA cm2, a Voc of 0.98
V with a fill factor of 0.63 yielding an of 10.9%. The third curve,
dashed trace, shows a device with respectable Jsc of 15.4 mA cm2 and
very high Voc of 1.13 V but a low fill factor of 0.45, yielding an overall
power conversion efficiency of 7.8%. In the SOM we show histograms
of device performance parameters for the TiO2 and Al2O3 based devices
(fig. S3) (25). The general trend is that the Al2O3 cells generated over
200 mV higher open-circuit voltages than the sensitized TiO2 solar cells,
with comparable short-circuit currents and slightly lower fill factors.
From the solar cell measurements on alumina-based devices, it was apparent that the perovskite layer could function as both absorber and ntype component, transporting electronic charge out of the device. We
further illustrate the semiconducting nature of the perovskite by the
construction of a planar-junction diode with a structure FTO/compact
TiO2/ CH3NH3PbI2Cl/Spiro-OMeTAD/Ag. The perovskite film was
approximately 150 nm thick in this configuration, and the solar cell generated Jsc of 7.13 mA cm2 Voc of 0.64 V fill factor of 0.4 and of 1.8%.
If we take the optical band gap of CH3NH3PbI2Cl to be 1.55eV from
the IPCE onset at 800nm (33), and the open-circuit voltage to be 1.1V,
this represents a difference in energy of only 0.45eV, competitive with
the best thin film technologies (2). To understand why we observed such
a dramatic increase in voltage over the TiO2 cells, we need to consider
the operational mode of the two concepts (Fig. 3A). We would expect
for sensitized TiO2 devices, after light absorption in the perovskite, electrons would be transferred to the TiO2 with subsequent electron transport
to the FTO electrode through the TiO2, and holes to the spiro-OMeTAD
with subsequent transport to the silver electrode. For Al2O3 based cells,
the electrons must remain in the perovskite phase (34), until they are
collected at the planar TiO2 coated FTO electrode, and must hence be
transported throughout the film thickness in the perovskite. Hole-transfer
from the photoexcited perovskite should occur to the spiro-OMeTAD in
much the same way as in the sensitized device. We confirm that Al2O3
does not act as an n-type oxide in DSSCs in fig. S4 (25).
In order to examine the charge generation in these devices, we performed photoinduced absorption (PIA) spectroscopy on the oxide films
coated with the perovskite, both with and without the addition of spiroOMeTAD. For the mesoporous TiO2 film coated with perovskite, the
PIA spectrum revealed features in the near infrared assigned to the free
electrons in the titania (35), confirming effective sensitization of the
titania by the perovskite. In contrast, films made of Al2O3 coated with
perovskite exhibit no PIA signal, confirming the insulating role of alumina. After addition of spiro-OMeTAD, we could efficiently monitor the
oxidized species of spiro-OMeTAD created after photoexcitation of the
perovskite. They had absorption features at 525 and 750 nm, as well as a
broad band around 1200 nm, assigned to the hole located on the
triarylamine moieties (28, 36), which dominate the spectra in both the

electrode and the perovskite absorber. A thick capping layer of p-type


spiro-OMeTAD readily resolves this issue, however; spiro-OMeTAD is
less conductive (~105 S cm1) so a thicker capping layer results in high
series resistance, thus we are presented with a compromise.
In summary, we have evolved the solid-state sensitized solar cell into a new concept with low fundamental losses. The application of a
mesostructured insulating scaffold upon which extremely thin films of ntype and p-type semiconductors are assembled, termed the mesosuperstructured solar cell (MSSC), has proven to be extraordinarily effective with an n-type perovskite, delivering over 10.9% power conversion efficiency under full sun illumination. Further advances in overall
power conversion efficiency are expected by extending the absorption
onset toward 940 nm, through the implementation of new perovskites or
broadening this concept to other solution-processable semiconductors.
Enhancing the light absorption near the band edge through carefully
engineered mesostructures or better photon management would lead to
increased photocurrent. Reduced series resistance through the utilization
of higher mobility hole-transporters, or better control over the capping
layer thickness would improve the fill factor. Finally, extending this
system to multi-junction devices, notably without the requirement for
lattice matching, would further enhance performance.
References and Notes
1. B. ORegan, M. Grtzel, A low-cost, high-efficiency solar cell based on dyefilms.
Nature
353,
737
sensitized
colloidal
TiO2
(1991). doi:10.1038/353737a0
2. M. A. Green, K. Emery, Y. Hishikawa, W. Warta, E. D. Dunlop, Solar cell
efficiency tables (version 39). Prog. Photovolt. Res. Appl. 20, 12
(2012). doi:10.1002/pip.2163
3. L. Han et al., High-efficiency dye-sensitized solar cell with a novel coadsorbent. Energy Environ. Sci. 5, 6057 (2012). doi:10.1039/C2EE03418B
4. A. Yella et al., Porphyrin-sensitized solar cells with cobalt (II/III)-based redox
electrolyte exceed 12 percent efficiency. Science 334, 629
(2011). doi:10.1126/science.1209688 Medline
5. G. Yu, J. Gao, J. C. Hummelen, F. Wudl, A. J. Heeger, Polymer photovoltaic
cells: Enhanced efficiencies via a network of internal donor-acceptor
heterojunctions.
Science
270,
1789
(1995). doi:10.1126/science.270.5243.1789
6. J. J. M. Halls et al., Efficient photodiodes from interpenetrating polymer
networks. Nature 376, 498 (1995). doi:10.1038/376498a0
7. A. H. Ip et al., Hybrid passivated colloidal quantum dot solids. Nature Nano. 7,
577 (2012). doi:10.1038/nnano.2012.127
8. T. K. Todorov, K. B. Reuter, D. B. Mitzi, High-efficiency solar cell with earthabundant
liquid-processed
absorber.
Adv.
Mater.
22,
E156
(2010). doi:10.1002/adma.200904155
9. H. J. Snaith, Estimating the maximum attainable efficiency in dye-sensitized
solar cells. Adv. Funct. Mater. 20, 13 (2010). doi:10.1002/adfm.200901476
10. G. Dennler, M. C. Scharber, C. J. Brabec, Polymer-fullerene bulkheterojunction
solar
cells.
Adv.
Mater.
21,
1323
(2009). doi:10.1002/adma.200801283
11. B. E. Hardin, H. J. Snaith, M. D. McGehee, The renaissance of dye-sensitized
solar cells. Nat. Photonics 6, 162 (2012). doi:10.1038/nphoton.2012.22
12. U. Bach et al., Solid-state dye-sensitized mesoporous TiO2 solar cells with
high photon-to-electron conversion efficiencies. Nature 395, 583
(1998). doi:10.1038/26936
13. J. Burschka et al., Tris(2-(1H-pyrazol-1-yl)pyridine)cobalt(III) as p-type
dopant for organic semiconductors and its application in highly efficient solidstate dye-sensitized solar cells. J. Am. Chem. Soc. 133, 18042
(2011). doi:10.1021/ja207367t Medline
14. Y. Itzhaik, O. Niitsoo, M. Page, G. Hodes, Sb2S3-sensitized nanoporous TiO2
solar cells. J. Phys. Chem. C 113, 4254 (2009). doi:10.1021/jp900302b
15. J. A. Chang et al., Panchromatic photon-harvesting by hole-conducting
materials in inorganic-organic heterojunction sensitized-solar cell through the
formation of nanostructured electron channels. Nano Lett. 12, 1863
(2012). doi:10.1021/nl204224v Medline
16. J. Nelson, Continuous-time random-walk model of electron transport in
electrodes.
Phys.
Rev.
B
59,
15374
nanocrystalline
TiO2

/ http://www.sciencemag.org/content/early/recent / 4 October 2012 / Page 4/ 10.1126/science.1228604

Downloaded from www.sciencemag.org on October 13, 2012

TiO2 and Al2O3 based samples. These results indicate that hole-transfer
is highly effective from the photoexcited perovskite to spiro-OMeTAD,
and specifically that a hole-conductor is required to enable long lived
charge species within the perovskite coated on the Al2O3. We note that
the PIA signal depended both on the concentration and lifetime of the
species monitored, hence from this measurement alone quantification of
the relative charge generated yield is not possible.
In order to probe the effectiveness of the perovskite layer at transporting electronic charge out of the device, we performed small perturbation transient photocurrent decay measurements (37). The solar cells
were exposed to simulated sun light and flashed with a small red light
pulse, the decay rate of the transient photocurrent signal is approximately proportional to the rate of charge transport out of the photoactive layer
(37). In Fig. 3C, we observed over tenfold faster charge collection in the
Al2O3 based devices than in the TiO2 based sensitized devices indicating
faster electron diffusion through the perovskite phase, than through the
n-type TiO2.
Because there was no n-type oxide in the Al2O3 based cells, the devices were no longer sensitized solar cells, but are now a twocomponent hybrid solar cell. As designed, the Al2O3 is simply acting as
a meso-scale scaffold upon which the device is structured, we term
this concept a meso-superstructured solar cell (MSSC). The above
measurements demonstrate that long-lived charge carriers can be generated via hole-transfer from the perovskite to spiro-OMeTAD, and that
the perovskite layer is faster at transporting electronic charge than the
mesoporous TiO2. However, they do not explain the increase in Voc: The
Voc is generated by the buildup of electrons in the n-type material and
holes in the p-type material, resulting in splitting of the quasi Fermi levels for both electrons and holes. For mesoporous TiO2 there exist sites in
the tail of the density of states which extend into the band gap (38). These fill with electrons under illumination, and result in the quasi Fermi
level for electrons (EFn*) being further from the conduction band for any
given charge density, than would the case be if these states did not exist,
i.e., in a crystalline semiconductor. The increased charge storing capacity of materials with a high density of sub-band gap states is termed
chemical capacitance (38). There is, in essence, no chemical capacitance of the Al2O3, and for the MSSCs all the electronic charge resides
in the perovskite, moving the EFn* in this material nearer the conduction
band for the same charge density. The higher voltage indicates that there
are fewer surface and sub-band gap states in the perovskite films than in
the mesoporous TiO2. Hence, the increased voltage is caused by a substantial reduction of the chemical capacitance of the solar cell. We used
a compact layer of TiO2 as the electron selective anode, but the chemical
capacitance of this extremely thin (50 to 100nm) TiO2 layer was very
low because of the low surface area (i.e., flat). In addition, the compact
layer deposited via spray pyrolysis has a donor density of around 1018
cm3 (39), and the sub-band gap sites responsible for the chemical capacitance may be full.
A central question is whether the MSSC is excitonic or a distributed
p-n junction. The perovskites tend to form layered structures, with continuous 2D metal halide planes perpendicular to the z-axis, and the lower
dielectric organic components (methyl amine) between these planes. The
quasi 2D confinement of the excitons results in an increased exciton
binding energy which can be up to a few hundred meV (40). The reasonably high photocurrents from the planar junction solar cells (Fig. 2B)
could be explained by either moderately delocalized and highly mobile
excitons being quenched at the perovskite spiro-OMeTAD interface, or
the generation of free charges in the bulk of the perovskite films with
reasonably good electron and hole-migration out of the devices.
The key limitation in performance of the MSSC at present is a balance between series and shunt resistance. The perovskite absorber is
reasonably conductive, measured to be on the order of 103 S cm3, thus
short-circuiting of the device occurs if contact exists between the silver

40. T. Ishihara, J. Takahashi, T. Goto, Optical properties due to electronic


transitions in two-dimensional semiconductors (CnH2n+1NH3)2PbI4. Phys.
Rev. B 42, 11099 (1990). doi:10.1103/PhysRevB.42.11099
Acknowledgments: This work was supported by European Research Council
(ERC) HYPER project no. 279881 and Strategic International Research
Cooperative Program (SCIP) of the Engineering and Physical Sciences
Research Council (EPSRC) and Japan Science and Technology Agency (JST).
T.M. thanks the funding program for World-Leading Innovative R&D on
Science and Technology (FIRST Program), Japan, for hybrid solar cell
research. We thank the NEDO projects for support. M.M.L. is grateful for
support from the Simms Bursary granted by Merton College, Oxford. We
thank S. K. Pathak for assistance with x-ray diffraction measurements and
analysis. We also thank Agnese Abrusci, James Ball, Pablo Docampo,
Andrew Hey, Tomas Leijtens and Nakita Noel, for valuable discussions. The
University of Oxford has filed 3 patents related to this work.
Supplementary Materials
www.sciencemag.org/cgi/content/full/science.1228604/DC1
Materials and Methods
Supplementary Text
Figs. S1 to S4
9 August 2012; accepted 7 September 2012
Published online 4 October 2012
10.1126/science.1228604

/ http://www.sciencemag.org/content/early/recent / 4 October 2012 / Page 5/ 10.1126/science.1228604

Downloaded from www.sciencemag.org on October 13, 2012

(1999). doi:10.1103/PhysRevB.59.15374
17. D. B. Mitzi, C. A. Field, W. T. A. Harrison, A. M. Guloy, Conducting tin
halides with a layered organic-based perovskite structure. Nature 369, 467
(1994). doi:10.1038/369467a0
18. K. Chondroudis, D. B. Mitzi, Electroluminescence from an organicinorganic
perovskite incorporating a quaterthiophene dye within lead halide perovskite
layers. Chem. Mater. 11, 3028 (1999). doi:10.1021/cm990561t
19. M. Era, T. Tsutsui, S. Saito, Polarized electroluminescence from oriented psexiphenyl vacuum-deposited film. Appl. Phys. Lett. 67, 2436
(1995). doi:10.1063/1.114599
20. C. R. Kagan, D. B. Mitzi, C. D. Dimitrakopoulos, Organic-inorganic hybrid
materials as semiconducting channels in thin-film field-effect transistors.
Science 286, 945 (1999). doi:10.1126/science.286.5441.945 Medline
21. A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, Organometal halide
perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem.
Soc. 131, 6050 (2009). doi:10.1021/ja809598r Medline
22. J. H. Im, C. R. Lee, J. W. Lee, S. W. Park, N. G. Park, 6.5% efficient
perovskite quantum-dot-sensitized solar cell. Nanoscale 3, 4088
(2011). doi:10.1039/c1nr10867k Medline
23. I. Chung, B. Lee, J. He, R. P. H. Chang, M. G. Kanatzidis, All-solid-state dyesensitized solar cells with high efficiency. Nature 485, 486
(2012). doi:10.1038/nature11067 Medline
24. H. J. Snaith, How should you measure your excitonic solar cells? Energy
Environ. Sci. 5, 6513 (2012). doi:10.1039/C2EE03429H
25. See supplementary materials on Science Online.
26.
A.
Poglitsch,
D.
Weber,
Dynamic
disorder
in
methylammoniumtrihalogenoplumbates (II) observed by millimeter-wave
spectroscopy. J. Chem. Phys. 87, 6373 (1987). doi:10.1063/1.453467
27. T. Leijtens et al., Hole transport materials with low glass transition
temperatures and high solubility for application in solid-state dye-sensitized
solar cells. ACS Nano 6, 1455 (2012). doi:10.1021/nn204296b Medline
28. H. J. Snaith et al., Charge collection and pore filling in solid-state dyesensitized solar cells. Nanotechnology 19, 424003 (2008). doi:10.1088/09574484/19/42/424003 Medline
29. J. Melas-Kyriazi et al., The effect of hole transport material pore filling on
photovoltaic performance in solid-state dye-sensitized solar cells. Adv. Energy
Mater. 1, 407 (2011). doi:10.1002/aenm.201100046
30. A. Abrusci et al., Facile infiltration of semiconducting polymer into
mesoporous electrodes for hybrid solar cells. Energy Environ. Sci. 4, 3051
(2011). doi:10.1039/C1EE01135A
31. I.-K. Ding et al., Pore-filling of spiro-OMeTAD in solid-state dye sensitized
solar cells: Quantification, mechanism, and consequences for device
performance.
Adv.
Funct.
Mater.
19,
2431
(2009). doi:10.1002/adfm.200900541
32. P. Docampo et al., Pore filling of spiro-OMeTAD in solid-state dye-sensitized
solar cells determined via optical reflectometry. Adv. Funct. Mater. n/a
(2012). doi:10.1002/adfm.201201223 Medline
33. D. A. R. Barkhouse, O. Gunawan, T. Gokmen, T. K. Todorov, D. B. Mitzi,
Device characteristics of a 10.1% hydrazine-processed Cu2ZnSn(Se,S)4 solar
cell. Prog. Photovolt. Res. Appl. 20, 6 (2012). doi:10.1002/pip.1160
34. A. Kojima, M. Ikegami, K. Teshima, T. Miyasaka, Highly luminescent lead
bromide perovskite nanoparticles synthesized with porous alumina media.
Chem. Lett. 41, 397 (2012). doi:10.1246/cl.2012.397
35. G. Rothenberger, D. Fitzmaurice, M. Graetzel, Spectroscopy of conduction
band electrons in transparent metal oxide semiconductor films: Optical
determination of the flatband potential of colloidal titanium dioxide films. J.
Phys. Chem. 96, 5983 (1992). doi:10.1021/j100193a062
36. G. Boschloo, A. Hagfeldt, Photoinduced absorption spectroscopy as a tool in
the study of dye-sensitized solar cells. Inorg. Chim. Acta 361, 729
(2008). doi:10.1016/j.ica.2007.05.040
37. P. Docampo et al., Control of solid-state dye-sensitized solar cell performance
by block-copolymer-directed TiO2 synthesis. Adv. Funct. Mater. 20, 1787
(2010). doi:10.1002/adfm.200902089
38. J. Bisquert, Chemical capacitance of nanostructured semiconductors: Its
origin and significance for nanocomposite solar cells. Phys. Chem. Chem.
Phys. 5, 5360 (2003). doi:10.1039/b310907k
39. L. Kavan, M. Grtzel, Highly efficient semiconducting TiO2 photoelectrodes
prepared by aerosol pyrolysis. Electrochim. Acta
40, 643
(1995). doi:10.1016/0013-4686(95)90400-W

www.sciencemag.org/cgi/content/full/science.1228604/DC1

Supporting Online Material for


Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal
Halide Perovskites
Michael M. Lee, Jol Teuscher, Tsutomu Miyasaka, Takurou N. Murakami, Henry J. Snaith*

*To whom correspondence should be addressed. E-mail: h.snaith1@physics.ox.ac.uk

Published 4 October 2012 on Science Express


DOI: 10.1126/science.1228604

This PDF file includes:


Materials and Methods
Supplementary Text
Figs. S1 to S4

Materials and Methods1


Organometal mixed-halide perovskite synthesis
Methylamine (CH3NH2) solution 33 wt% in absolute ethanol was reacted with hydroiodic acid
(HI) 57 wt % in water with excess methylamine under nitrogen atmosphere in ethanol at room
temperature. Typical quantities were 24 mL methylamine, 10 mL hydroiodic acid and 100 mL
ethanol. Crystallization of methylammonium iodide (CH3NH3I) was achieved using a rotary
evaporator; a white colored powder was formed indicating successful crystallization.
Methylammonium iodide (CH3NH3I) and lead (II) chloride (PbCl2) was dissolved in anhydrous
N,N-Dimethylformamide at a 3:1 molar ratio of CH3NH3I to PbCl2, to produce a mixed halide
perovskite precursor solution. In spin-coated and dried films, the formed perovskite of
methylammonium lead iodide chloride (CH3NH3PbI2Cl) had a final I to Cl ratio of
approximately 2:1 as determined by energy dispersive X-Ray analysis. The excess
methylammonium iodide and chlorine is assumed to be lost via evaporation during the drying
process.
Aluminium oxide paste synthesis
Aluminium oxide dispersion 10 wt% in water was washed by centrifuging at 7500 RPM for 6
hours and redispersing in absolute ethanol with an ultrasonic probe at a duty cycle of 50 % (2
seconds on, 2 seconds off) for a duration of 5 minutes. The washing process was repeated for
three cycles. For every 10 g of the original dispersion (1 g total Al2O3) the following was added:
3.33 g of -terpineol and 5 g of a 50:50 mix of ethyl-cellulose 10 cP and 46 cP in ethanol 10 wt
%. After the addition of each component, the mix was stirred for 2 minutes and sonicated with
the aforementioned sonication program for 1 minute. Finally, the resulting mixture was
introduced into a rotary evaporator to remove excess ethanol in order achieve the required paste
consistency suitable for doctor blading, spin-coating or screen-printing.
Perovskite-Sensitized TiO2 and Mesoporous-Superstructured Solar Cell fabrication
Fluorine-doped tin oxide (F:SnO2) coated glass (Pilkington TEC 15) 15 / was patterned by
etching with Zn powder and 2 M HCl diluted in milliQ water. The etched substrate was then
cleaned with 2% hellmanex diluted in milliQ water, rinsed with milliQ water, acetone and
ethanol and dried with clean dry air. The substrate underwent an oxygen plasma treatment for 5
minutes prior to spray pyrolysis of compact titanium dioxide (TiO2). A thin layer of compact
anatase TiO2 of roughly 50 nm in thickness was formed through spray pyrolysis of titanium
diisopropoxide bis(acetylacetonate) diluted in anhydrous ethanol at a volumetric ratio of 1:10
using N2 as a carrier gas. Either 0.5 m thick mesoporous Al2O3 layer was deposited by spincoating Al2O3 paste diluted further in anhydrous ethanol at 1:1.5 by weight at 2000 RPM, or 0.5
m thick mesoporous TiO2 layer was deposited by spin-coating TiO2 paste (Dyesol 18NR-T)
diluted in anhydrous ethanol at 1:2.5 by weight at 2000 RPM. The layers were then sintered in
air at 550 C for 30 minutes. Once cooled and cut down to device size 25 l of 20 wt%
perovskite precursor solution was dispensed onto each pre-prepared mesoporous electrode film
spin-coating at 1500 RPM for 30 seconds in air. The coated films were then placed on a hot
plate set at 100 C for 45 minutes in air. During the drying procedure at 100 C, the coated
electrode changed color from light yellow to dark brown, indicating the formation of the
1

Unless stated otherwise all chemicals were purchased from Sigma-Aldrich and were anhydrous if available.
2

perovskite film. 25 l of a chlorobenzene solution containing 68 mM Spiro-OMeTAD, 55 mM


tert-butylpyridine and 9 mM lithium bis(trifluoromethylsyfonyl)imide salt was cast onto the
perovskite coated substrate and spun at a rate of 2000 RPM for 45 seconds. Cells were left in the
dark in air overnight prior to thermal evaporation of 200 nm Ag electrodes to complete the solar
cells.
Solar Cell Characterization
Current-Voltage characteristics were measured
(2400 Series SourceMeter, Keithley
Instruments) under simulated AM 1.5G sunlight at 100 mWcm-2 irradiance; generated using an
AAB ABET technologies Sun 2000 solar simulator and calibrated using an NREL calibrated
silicon reference cell with a KG5 filter to minimise spectral mismatch (the mismatch factor was
calculated to be less than 1%). The solar cells were masked with a metal aperture to define the
active area which was typically 0.09 cm2 and measured in a light-tight sample holder to
minimize any edge effects.
Photovoltaic action spectra were measured (2400 Series SourceMeter, Keithley Instruments)
with chopped monochromatic light incident which were biased with white light-emitting diodes
(LED) at an equivalent solar irradiance of 10 mWcm-2. The monochromatic light intensity for
the incident photon-to-electron conversion efficiency (IPCE) was calibrated with a UV-enhanced
silicon photodiode. The solar cells were masked with a metal aperture to define the active area
which was typically 0.09 cm2 and measured in a light-tight sample holder to minimize any edge
effects.
Photo-induced absorption (PIA) spectra were obtained from perovskite/TiO2 or Al2O3 films
deposited on FTO coated glass. The films were excited with an Argon ion laser tuned at 496.5
nm with a maximum fluence of 50 mW cm-2 and chopped at a frequency of 23 Hz. The detection
is made with a continuous white light probe (halogen bulb) of around 1 sun intensity onto the
sample. After passing through the sample, the probe beam enters a monochromator (SpectraPro2300i, Acton Research Corporation) coupled to diode for detection in the visible (PDA10A,
Thorlabs) and in the NIR (ID-441-C, Acton Research Corporation). Acquisition is made by a
lock in amplifier locked at the light modulation frequency (SR830, Stanford Research Systems)
and a NI USB-6008 (National Instruments) acquisition card. A computer running LabView
(National Instruments) controls the setup and record spectra. No further treatment is applied to
the data.

SOM Text
Determining the presence of mixed-halide perovskite in mesoporous Al2O3
To confirm the presence of the CH3NH3PbI2Cl throughout the mesoporous Al2O3, we have
performed energy dispersive x-ray (EDX) analysis on a cross section of a device. Fig. S2 shows
a cross-sectional scanning electron micrograph (SEM) of a ~3 m thick device (left), and
energy-dispersive x-ray (EDX) spectroscopy with elemental mapping of (right) aluminium, lead,
chlorine, iodine. We show the uniform distribution of Al, Pb, Cl, and I throughout a ~3 m thick
film. We note that for the optimized devices the mesoporous alumina films are only ~0.5 m
thick, but we have used a 3 m due to spatial resolution plimitations of the EDX technique. The
elemental composition determined by the EDX gives an approximately 2:1 I:Cl ratio.
Comparison between sensitised versus meso-superstructure device architectures
Fig. S3 shows histogram plots of key solar cell performance parameters: fill factor, open-circuit
voltage, short-circuit current and photo-conversion efficiency for a large batch of perovskite cells
measured under simulated AM 1.5 100 mW cm-2 illumination. The main observation is that the
fill factor is generally lower but the open-circuit voltage is significantly higher, giving overall a
slightly higher efficiency for the meso-superstructure solar cells (MSSC) as compared to the
TiO2 sensitized devices. The distribution is reasonably wide, reducing this spread in performance
is central to the commercialization of this technology.
Confirming the insulating nature of Al2O3
To confirm the insulating nature of alumina (Al2O3), solid-state dye-sensitized solar cells were
made with either mesoporous titania (TiO2) or mesoporous alumina (Al2O3). Fig. S4 Currentvoltage characteristics under AM 1.5 (100 mWcm-2) simulated sunlight (solid) and in dark
(dashed) to red trace- TiO2, and bottom black trace- Al2O3. The devices were constructed in the
same manner as the perovskite-sensitized or the meso-superstructured solar cells (using spiroOMeTAD as the hole-transporing medium); however, a prototypical ruthenium based dye
(termed N719) was used in lieu of the perovskite absorber. The film thicknesses of the
mesoporous TiO2 and Al2O3 were around 500nm, accounting for the relatively low photocurrent
of the TiO2 sensitized device.

Figures

Fig. S1. X-ray Diffraction (XRD) spectra of CH3NH3PbI2Cl spin-coated on a glass slide
and heated to 100C before measuring.

Fig. S2. Cross-sectional scanning electron micrograph (SEM) of a ~3 m thick device


(left), and energy-dispersive x-ray (EDX) spectroscopy with elemental mapping of (right)
aluminium, lead, chlorine, iodine.

Fig. S3. Histogram plots of solar cell performance parameters: open-circuit voltage
(VOC), short-circuit current (JSC) fill factor (FF) and photo-conversion efficiency () for a
large batch of CH3NH3PbI2Cl sensitized TiO2 photovoltaic devices (black, total of 262
devices) and meso-superstructured solar cells (Al2O3 based) (red, total of 767 devices).
The Gaussian fits have been added to aid the eye.

Fig. S4. Current-voltage characteristics measured under AM 1.5 100mWcm-2 simulated


sunlight (solid) and in dark (dashed) for solid-state dye-sensitised solar cells with TiO2
(red trace with crosses) and Al2O3. (black trace with circles).

Das könnte Ihnen auch gefallen