Vous êtes sur la page 1sur 6

PCCP

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

COMMUNICATION

Cite this: Phys. Chem. Chem. Phys.,


2014, 16, 1327

View Article Online


View Journal | View Issue

WO3a-Fe2O3 composite photoelectrodes with


low onset potential for solar water oxidation
Peng Zhao,a Coleman X. Kronawitter,b Xiaofang Yang,b Jie Fua and Bruce E. Koel*b

Received 5th August 2013,


Accepted 11th November 2013
DOI: 10.1039/c3cp53324g
www.rsc.org/pccp

The physical and photoelectrochemical properties of a composite


oxide photoelectrode comprised of a-Fe2O3 and WO3 crystals is
investigated. The composite films exhibit a water oxidation photocurrent onset potential as low as 0.43 V vs. RHE, a value considerably lower than that of pure a-Fe2O3 photoanodes prepared in
comparable synthesis conditions. This result represents one of the
lowest onset potentials measured for hematite-based PEC water
oxidation systems. Compositional analysis by X-ray Photoelectron
Spectroscopy and Energy Dispersive Spectroscopy indicates the
composition of the films diers between the surfaces and bulk, with
tungsten found to be concentrated in the surface region. Post-reaction
Raman spectroscopy characterization demonstrates that water interacts with surface WO3 crystals, an event that is associated with the
formation of a hydrated form of the oxide.

Composite oxide films represent an interesting class of materials


for photoelectrochemical (PEC) applications because they can
be designed to utilize the often complementary physical properties of dissimilar metal oxides. This characteristic is made
possible by the diversity of optical, chemical, and electronic
properties of transition metal oxides. For solar water oxidation,
a light-driven reaction of critical technological importance for
solar fuel development, hematite (a-Fe2O3) is an especially
promising electrode constituent material due to its optimal
band gap (2.12.2 eV), chemical stability in aqueous non-acidic
electrolytes, abundance, and nontoxicity.16 Additionally, the
suitably anodic position of its valence band edge on the electrochemical scale permits its implementation as a photoanode to drive
the water oxidation reaction.7 However, the use of a-Fe2O3 for this
application is especially challenging because it is associated with a
very short hole diffusion length (on the order of a few nanometers),
a

Department of Chemistry, Princeton University, Princeton, New Jersey 08544, USA.


E-mail: pztwo@princeton.edu; Tel: +1-609-712-0829
b
Department of Chemical and Biological Engineering, Princeton University,
Princeton, New Jersey 08544, USA. E-mail: bkoel@princeton.edu;
Fax: +1-609-258-0211; Tel: +1-609-258-4524
Electronic supplementary information (ESI) available. See DOI: 10.1039/
c3cp53324g

This journal is the Owner Societies 2014

which hinders efficient collection of carriers via interfacial


charge-transfer reactions.8
Recently researchers have focused on improving the photoconversion eciency of a-Fe2O3 through design of composite
and nanostructured films.911 In addition, it is frequently found
that photocatalysts or photoelectrodes containing heterojunctions
of two types of materials show enhanced performance when
compared to their constituent phases alone.1215 Among other
reported benefits, it has been shown that heterojunctions
facilitate charge separation and prolong the lifetime of photogenerated carriers. A similar approach for increasing the
efficiency of otherwise promising photoanode materials is to
design composite oxide films, which have the potential to
utilize the often complementary properties of the constituent
oxides. The success of these electrodes depends in large part on
the direction and energetics of charge transport among the
oxide phases. Also, narrow-bandgap semiconductors have been
used to couple with a wide-bandgap oxide (e.g. TiO2, ZnO) to
harvest visible light and to enhance separation efficiency of
carriers.1619 Thus, mixed oxides are promising materials to
solve problems we mentioned above.
As an n-type semiconductor with a wide bandgap, WO3 in many
ways possesses electronic and optical properties complementary to
those of a-Fe2O3.20,21 Here, we report a WO3a-Fe2O3 composite
electrode possessing a low onset potential for water photo-oxidation,
and discuss the physical properties of the electrode that are likely to
relate to the origins of this beneficial property.
Films were deposited onto transparent conducting FTO-coated
glass substrates by simultaneous sputter-deposition of Fe2O3 and
WO3 targets under high vacuum conditions (106 Torr with a
3 : 1 O2 and Ar mixture as the sputtering gas). As deposited,
the films were amorphous, and then were annealed at 700 1C in
1-atm O2 for 4 h to crystallize.
Fig. 1a and b show the morphology of the WO3a-Fe2O3
films. These SEM images reveal that the films are comprised
of arrays of nanocolumnar structures oriented perpendicular
to the substrate surface. In Fig. 1b, the top half of the picture
displays the deposited film and bottom shows the polycrystalline

Phys. Chem. Chem. Phys., 2014, 16, 1327--1332 | 1327

View Article Online

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

Communication

PCCP

Fig. 2 Photoelectrochemical characterization of WO3a-Fe2O3, pure


WO3 and pure a-Fe2O3 films (all annealed for 4 h at 700 1C) under AM
1.5 solar illumination. (a) Currentpotential plots in 0.5 M Na2SO4 in the
dark and under AM 1.5 illumination plotted against the reversible hydrogen
electrode (RHE). (b) UV-Vis absorption spectra. (c) Incident photon to
current efficiency (IPCE) measured at several wavelengths.
Fig. 1 (a, b) SEM images of WO3a-Fe2O3 samples annealed at 700 1C.
Both images are side views with a 651 tilt. (c) EDS of WO3a-Fe2O3 film.
(d) XRD patterns of WO3a-Fe2O3 film and pure a-Fe2O3 film. (Ref. trigonal
Fe2O3 PDF#33-0664, orthorhombic WO3 PDF#20-1324, monoclinic WO3
PDF#24-0747, tetragonal SnO2 PDF#33-1374.)

FTO substrate. The individual pillars are 100200 nm in diameter


and 400500 nm in height, while the spacings between pillars
range from 50 to 100 nm. Pure a-Fe2O3 films deposited in
identical conditions lacked the columnar morphology and
instead possessed a denser porous structure (Fig. S1, ESI).
These films have a much smaller surface area and when applied
as photoelectrodes are expected to be associated with longer
required hole transport distances to the surface compared with
nanocolumnar films. Fig. 1c and d show compositional and
structural characterization of the WO3a-Fe2O3 films by energy
dispersive spectroscopy (EDS) and X-ray diffraction (XRD),
respectively, as discussed further below.
PEC measurements were performed in a glass cell fitted with
a quartz window, using a three-electrode configuration with a
Ag/AgCl reference electrode, and a platinum wire counter
electrode separated from the FTO working electrode by a glass
frit. The electrolyte was an aqueous 0.5 M Na2SO4 solution at
pH E 7. Photocurrent densities were measured as a function of
applied voltage under simulated 1 sun AM 1.5 solar irradiation.
Our previous work showed that photoelectrochemical performance depends strongly on electrode surface structure, deposition substrate temperature, source flux, deposition angle, and
the post-deposition annealing treatment.
In this study, we found that post-deposition annealing temperature strongly aects the PEC performance of WO3a-Fe2O3.
We observed a noticeable initial rise in the current density
when the annealing temperature was increased from 400 to
500 1C (Fig. S2, ESI). This significant improvement may be
attributed to the increased crystalline order associated with
high temperature annealing, which should facilitate charge
carrier transport rather than recombination (a major limiting
factor for PEC performance of pure a-Fe2O3). The photocurrent
density reached the highest value when electrodes were

1328 | Phys. Chem. Chem. Phys., 2014, 16, 1327--1332

annealed to 700 1C, (the glass substrate loses stability at


temperatures higher than 800 1C) and so these conditions were
chosen for discussion of PEC characterization in the main text.
Fig. 2a shows dark and photocurrent densitypotential curves
for WO3, a-Fe2O3 and WO3a-Fe2O3 photoanodes under front-side
solar-simulated illumination. The WO3 and a-Fe2O3 electrode
yielded photocurrent curves that starts to rise (the onset potential,
the potential at which electrochemical water oxidation starts) at
0.54 and 0.73 V vs. RHE (reversible hydrogen electrode), and are
associated with current densities of 0.5 and 1.0 mA cm2 at
1.4 VRHE (black curve), respectively. The WO3a-Fe2O3 photoanodes are associated with an onset potential of 0.43 VRHE and
a plateau in the current density at 0.84 mA cm2 (green curve).
By comparing the dark and photo current of the a-Fe2O3
samples, we obtained a 0.94 V shift in the onset potential from
1.67 VRHE (red curve) to 0.73 VRHE (black curve), which is
consistent with most pure hematite photoanodes.22,23 The
performance of WO3 film alone is comparable to results
reported before with an onset potential of 0.50 VRHE and a
current density of 0.7 mA cm2 at 1.4 VRHE (in Na2SO4 + H2SO4
electrolyte).24 However, for the WO3a-Fe2O3 films, the dark
current onset potential is 1.65 VRHE and the photocurrent onset
potential is 0.43 VRHE, which corresponds to a shift of 1.22 V.
In these PEC experiments, an external bias is required to
raise the free energy of electrons suciently and convert this
energy into additional photovoltage to carry out the water
splitting reaction. However, in order to ensure compatibility
with promising tandem devices for solar water splitting,25
which require no external power supply, a significant cathodic
shift in the hematite electrode onset potential is required. Our
observation of a large negative shift in the onset potential is
therefore relevant to the design of unassisted water splitting
devices.
The WO3a-Fe2O3 films showed photocurrent densities that
were similar to those of a-Fe2O3 samples under visible light
illumination. This is a surprising result since their associated
nanocolumnar morphology might be advantageous for water
splitting performance, by reducing the required diffusion

This journal is the Owner Societies 2014

View Article Online

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

PCCP

Communication

distances for holes to reach the film/electrolyte interface. In the


photoanodic system examined here, the majority of photoexcited electrons are lost in electronhole recombination processes, and only holes generated near the interface can be used
for oxidation. Consequently previous studies have reported
higher photocurrents associated with columnar structures.23
However, the general idea of reducing feature dimensions to
match carrier diffusion lengths is now ubiquitous in the literature of this field, and the benefit of nanostructuring electrodes is
in fact material dependent. Below we examine some physical
properties of the electrodes to help further elucidate the nature
of the composite oxide electrode.
Fig. 2b shows UV-Vis absorbance measured on a-Fe2O3 and
WO3a-Fe2O3 films. The UV-Vis spectra show an absorption
onset at 400450 nm for WO3a-Fe2O3 samples and an onset just
below 600 nm for the a-Fe2O3 samples. The nanocolumnar films
cannot absorb visible light at wavelengths larger than 500 nm
well, which decreases the total portion of photons absorbed and
leads to a relatively small maximum current density. The incident photon to current efficiency (IPCE) of our films for water
splitting can be tested by studying the spectral photoresponse;
the IPCE value is given by:26
IPCE%

1240  jph
 100
lI

(1)

in which, jph is the steady state photocurrent density, l is the


wavelength of the incident light, and I is the light power
intensity at the film surface. The IPCE value for pure a-Fe2O3
and WO3a-Fe2O3 films was measured at 0.40 VAg/AgCl (1.01 VRHE),
as shown in Fig. 2c. In the UV region below 300 nm, IPCE values
for WO3a-Fe2O3 films are relatively low compared to that for
WO3, but higher than that for a-Fe2O3. In the UV region from
300 to 400 nm, WO3a-Fe2O3 samples have better performance
(B20% IPCE) than the constituent oxides alone. In the visible
light region, a-Fe2O3 has the highest IPCE values, which is
consistent with expectations considering the higher absorption
observed for a-Fe2O3 in this spectral range (Fig. 2b). The shapes
of IPCE spectra for WO3a-Fe2O3 and a-Fe2O3 are similar to
that of their respective UV-Vis spectra. In other words, the
presence of WO3 in the films improves the photoresponse of
the electrodes in the near UV region.
X-ray photoelectron spectroscopy (XPS) and EDS measurements were carried out for further compositional analysis at the
surface and in the bulk of the electrodes. Analysis of the EDS
measurements shown in Fig. 1c indicates the bulk film composition contains 35% W and 65% Fe. XPS measurements,
which probe the surface of the WO3a-Fe2O3 film, are shown in
Fig. 3a. These measurements establish that the surface contains Fe, W, O, and Sn, with the presence of Sn attributed to
uncovered regions of the FTO substrate. Narrow scans of the Fe
and W core levels were obtained to investigate the metals
oxidation states. The Fe 2p3/2 and 2p1/2 peaks at 710.9 and
724.5 eV binding energy (BE) are indicative of the Fe3+ in
Fe2O3.27 The W 4f7/2 and 4f5/2 peaks at 31.4 and 33.6 eV BE
are assigned as due to W6+ in WO3.28 Assuming a homogeneous
distribution of Fe and W in the near surface region, and using

This journal is the Owner Societies 2014

Fig. 3 XPS spectra of WO3a-Fe2O3 film: (a) broad scan, (b) Fe 2p region,
(c) W 4f region.

appropriate sensitivity factors,29 we calculate that there is 12%


Fe and 88% W at the surface. In a previous study,30 comparing
the morphology of undoped, W6+-doped, and WO3-modified
a-Fe2O3 nanorod films, the different morphologies indicated
that W6+ can easily diffuse into the crystal lattice. However, for
WO3-modified samples, W6+ was not doped into the lattice, and
instead existed as WO3 on the surface. Our work is consistent
with these findings; WO3 in our co-deposited films is more
concentrated near the a-Fe2O3 surface.
X-ray diraction (XRD) patterns of both films with and
without modification have similar peaks superimposed on
the FTO substrate pattern (Fig. 1d). The intensity distribution
in the XRD pattern of the a-Fe2O3 (red lines) corresponds to the
reference pattern of hematite (trigonal Fe2O3, PDF#33-0664).
There are three strong peaks corresponding to the (110), (104),
and (012) planes, along with weaker peaks, namely the (113),
(024), and (116) reflections. No evidence was found for other
possible phases of iron oxide. Compared with a-Fe2O3 films, an
increase in the relative peak intensities of the hematite (104)
and (012) planes were observed in the WO3a-Fe2O3 films,
while there was a small decrease in the peak intensity for the
(110) plane. Based on previous studies,31,32 enhancement in the
(110) orientation of films facilitates collections of photogenerated electrons, due to the higher conductivity of planes
perpendicular to the crystals hexagonal c-axis. The decrease in
this orientation in our films may contribute to the unexpectedly
similar photocurrent densities among the samples. WO3 in
the WO3a-Fe2O3 film is best assigned as orthorhombic WO3
rather than monoclinic WO3. The patterns of the two phases

Phys. Chem. Chem. Phys., 2014, 16, 1327--1332 | 1329

View Article Online

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

Communication

are similar, although the peak near 451 is distinct to the


orthorhombic phase.
It is important to consider the stability of the WO3a-Fe2O3
films, since WO3 is most stable in acidic media (pH o 3), which
are conditions where a-Fe2O3 is unstable. We performed a
series of XPS measurements studying the surface concentration
of W after various durations of PEC reaction; the results of
these experiments are shown in Fig. S3 (ESI). The intensity of
the W signal decreased 8% after 30 min reaction and 10% after
1 h reaction. After times longer than 1 h (up to 3 h), there was
no additional change that would suggest further corrosion had
occurred. In all experiments, no photocurrent decrease larger
than 5% was observed and no onset potential shift was
observed. This finding motivated additional characterization
of the WO3a-Fe2O3 films after PEC reactions.
Information on the interaction of the electrolyte with the
films during PEC reactions can be obtained by Raman spectroscopy, which can monitor the vibrational modes of compounds
formed from the reaction. Raman spectra of the a-Fe2O3 film,
WO3a-Fe2O3 film, and WO3 film deposited on an FTO glass
substrate are shown in Fig. 4. The spectrum for the a-Fe2O3 film
shows peaks at 224, 244, 293, 409, 612, and 659 cm1, almost all
of which can be assigned to modes present in hematite, a-Fe2O3.26
The broad band present at 659 cm1 has been reported by a
number of studies and is commonly attributed to the presence of
defects.3335 The reduction in the space symmetry group and the
appearance of this forbidden Raman mode would lead to a
significant lattice distortion. Bersani also correlated crystalline
size with the intensity of the 659 cm1 peak, which suggests
that this peak may be related in particular to surface and grain
boundary disorder.36
The spectrum of the pure WO3 has well-defined Raman peaks
at 273, 327, 434, 715, and 807 cm1. These peaks are very close to
the wavenumbers commonly observed in crystalline tungsten
oxide.37 The peaks at 715 and 807 cm1 correspond to OWO
stretching modes (n(OWO)) of bridging oxygen, and the peaks
at 273, 327, and 434 cm1 are bending modes.37 Compared to
the WO3 and a-Fe2O3 spectra, the spectrum for WO3a-Fe2O3 has
reduced intensity for all Raman peaks. The 224, 293, 409, and
612 cm1 peak of a-Fe2O3 can still be easily observed. For the
composite film, Raman peaks at 170, 210, 253, 320, 377, 434,
805, and 948 cm1 are observed and associated with band
broadening compared to results of tungsten trioxide hydrates
in the ref. 37. This broadening suggests that the crystalline size
of WO3 is much smaller than that of a-Fe2O3. Among these
Raman peaks, with the exception of that at 434 cm1, all can be
assigned to tungsten trioxide hydrates, including WO31/3H2O,
WO3H2O and WO32H2O.37
It has been reported that WO3 will dissolve in basic solutions,
and that its phase can transition to that of a hydrated oxide in
neutral solutions.37 Our observations here are consistent with this
phenomenon: XPS results indicate WO3 is concentrated in the
surface region and post-reaction Raman measurements suggest
the formation of a hydrated WO3 in neutral reaction solution.
Another interesting observation is that for the composite film
there is no peak associated with a-Fe2O3 defects (near 659 cm1).

1330 | Phys. Chem. Chem. Phys., 2014, 16, 1327--1332

PCCP

Fig. 4 Raman spectra of a WO3a-Fe2O3 film, pure WO3 film, and a-Fe2O3
film deposited on a FTO glass substrate. Numbers and arrows in the figure
indicate values from reference data from previous reports that were used for
identification.26,3337

It is conceivable that the small WO3 crystallites exist on the


surface or grain boundaries of a-Fe2O3, reducing the density of
defect states.
The mechanism by which the incorporation of WO3 into
the electrode causes a large cathodic shift in onset potential
compared to pure a-Fe2O3 is not known, although the character
of the photoelectrochemical response provides some additional
information. In our consideration of why and how the addition of
WO3 to a-Fe2O3 improves the onset potential, we note previous work
that studied carrier dynamics when CoOx and Ga2O3 were deposited
onto a a-Fe2O3 film.38 In these studies there was a negative shift in
the photocurrent onset, an observation that was correlated with the
appearance of long-lived photoholes. This relationship was attributed to a cathodic shift in the development of electron depletion
conditions in the a-Fe2O3. Several recent studies have also suggested
that the negative shift in the onset potential by treatment with oxide
overlayers may result from passivation of surface recombination
sites,39,40 thereby partially suppressing surface recombination. Thus,
a possible explanation of our results is that WO3 treatment,
as carried out here, leads to surface site passivation and/or
reduces the electron density in the a-Fe2O3 photoelectrode,

This journal is the Owner Societies 2014

View Article Online

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

PCCP

thereby retarding electronhole recombination and increasing


the yield of long-lived holes at the electrode surface.
It is worth speculating on other possible explanations for our
results. For example, the observed cathodic shift might be related
to the distinct electronic structure of WO3. This is relevant here
because the measured onset potential in an operating photoelectrochemical cell is intimately related to the quasi-Fermi level
of electrons in the photoelectrode. In addition, the cathodic shift
in onset potential for a-Fe2O3 photoanodes has been observed to
be associated with a number of system features, including modification of the electrode surfaces with water oxidation catalysts41 as
well as the presence of hole scavengers in the electrolyte.42
Typically these features are suggested to reduce the energy barrier
for hole injection from the photoelectrode to the electrolyte, which
manifests as a reduction in the required applied overpotential for
water oxidation. Thus, in the system presently investigated, it is
possible that the large concentration of tungsten measured at the
surface in some way reduces this energy barrier. Then, one could
speculate that the improvement in onset potential is related to the
distinct electronic structure of WO3, whose valence band is
comprised primarily of O 2p character. The superior kinetics of
charge injection into O 2p bands compared to the Fe 3d bands of
a-Fe2O3 has been observed previously through inference of the rate
constant for electron transfer at surfaces. The rate constant for
electron transfer was determined to be orders of magnitude
greater at a WO3 surface than at an a-Fe2O3 surface.8

Conclusions
In summary, we find that the physical and photoelectrochemical
properties of a-Fe2O3 films are favorably influenced by modification
with WO3. In this work, a nanocolumnar WO3a-Fe2O3 composite
film was synthesized and characterized in a photoelectrochemical
cell for water oxidation. UV-Vis absorption measurements show that
the WO3a-Fe2O3 film has an increase in absorption below 400 nm
and a decrease in absorption in longer visible wavelengths when
compared with a-Fe2O3. Current-potential and IPCE measurements
in neutral aqueous electrolytes showed that the WO3a-Fe2O3
electrode performs better (20% IPCE) than Fe2O3 and WO3 in the
near-UV region, and in addition is associated with a reduced onset
potential under solar light illumination. The onset potential for
water oxidation was 0.43 VRHE, and the maximum photocurrent
density of the photoanode was 0.84 mA cm2. Post-reaction Raman
spectroscopy measurements suggest the photoelectrochemical
reaction is associated with the formation of a surface tungsten
oxide hydrate phase. Our future studies to further understand this
composite oxide electrode and the shift of the onset potential
involve identifying flat band potentials, probing the carrier
dynamics on short timescales, as well as surface science measurements on the interaction of water with surface adsorption sites.

Acknowledgements
This material is based upon work supported by the Addy/ISN
North American Low Carbon Emission Energy Self-Suciency

This journal is the Owner Societies 2014

Communication

Fund of the Andlinger Center for Energy and the Environment


(ACEE) and by the Grand Challenges Program at Princeton
University.

References
1 H. W. Gao, C. Liu, H. E. Jeong and P. D. Yang, ACS Nano,
2012, 6, 234240.
2 K. Sivula, F. Le Formal and M. Gratzel, ChemSusChem, 2011,
4, 432449.
3 Y. J. Lin, G. B. Yuan, S. Sheehan, S. Zhou and D. W. Wang,
Energy Environ. Sci., 2011, 4, 48624869.
4 N. T. Hahn, H. C. Ye, D. W. Flaherty, A. J. Bard and
C. B. Mullins, ACS Nano, 2010, 4, 19771986.
5 I. Cesar, K. Sivula, A. Kay, R. Zboril and M. Graetzel, J. Phys.
Chem. C, 2009, 113, 772782.
6 C. Sanchez, K. D. Sieber and G. A. Somorjai, J. Electroanal.
Chem., 1988, 252, 269290.
7 A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253278.
8 M. P. Dareedwards, J. B. Goodenough, A. Hamnett and
P. R. Trevellick, J. Chem. Soc., Faraday Trans. 1, 1983, 79,
20272041.
9 H. W. Tang, W. J. Yin, M. A. Matin, H. L. Wang, T. Deutsch,
M. M. Al-Jassim, J. A. Turner and Y. F. Yan, J. Appl. Phys.,
2012, 111, 073502.
10 W. D. Chemelewski, N. T. Hahn and C. B. Mullins, J. Phys.
Chem. C, 2012, 116, 52565262.
11 N. T. Hahn and C. B. Mullins, Chem. Mater., 2010, 22,
64746482.
12 H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri
and J. H. Ye, Adv. Mater., 2012, 24, 229251.
13 V. B. R. Boppana and R. F. Lobo, ACS Catal., 2011, 1,
923928.
14 S. F. Chen, X. L. Yu, H. Y. Zhang and W. Liu, J. Hazard.
Mater., 2010, 180, 735740.
15 Z. Y. Wang, B. B. Huang, Y. Dai, X. Y. Qin, X. Y. Zhang,
P. Wang, H. X. Liu and J. X. Yu, J. Phys. Chem. C, 2009, 113,
46124617.
16 M. Sun, G. D. Chen, Y. K. Zhang, Q. Wei, Z. M. Ma and
B. Du, Ind. Eng. Chem. Res., 2012, 51, 28972903.
17 F. L. Zhou, X. J. Li, J. Shu and J. Wang, J. Photochem.
Photobiol., A, 2011, 219, 132138.
18 L. L. Xu, J. G. Guan, L. A. Gao and Z. G. Sun, Catal. Commun.,
2011, 12, 548552.
19 G. S. Li, D. Q. Zhang and J. C. Yu, Environ. Sci. Technol.,
2009, 43, 70797085.
20 A. Mao, J. K. Kim, K. Shin, D. H. Wang, P. J. Yoo, G. Y. Han
and J. H. Park, J. Power Sources, 2012, 210, 3237.
21 K. Sivula, F. Le Formal and M. Gratzel, Chem. Mater., 2009,
21, 28622867.
22 S. D. Tilley, M. Cornuz, K. Sivula and M. Gratzel, Angew.
Chem., Int. Ed., 2010, 49, 64056408.
23 I. Cesar, A. Kay, J. A. G. Martinez and M. Gratzel, J. Am.
Chem. Soc., 2006, 128, 45824583.
24 Z. H. Jiao, J. M. Wang, L. Ke, X. W. Sun and H. V. Demir,
ACS Appl. Mater. Interfaces, 2011, 3, 229236.

Phys. Chem. Chem. Phys., 2014, 16, 1327--1332 | 1331

View Article Online

Published on 21 November 2013. Downloaded by University of California - Davis on 12/10/2016 04:58:08.

Communication

25 J. Brillet, J. H. Yum, M. Cornuz, T. Hisatomi, R. Solarska,


J. Augustynski, M. Graetzel and K. Sivula, Nat. Photonics,
2012, 6, 823827.
26 A. Duret and M. Gratzel, J. Phys. Chem. B, 2005, 109,
1718417191.
27 A. P. Grosvenor, B. A. Kobe, M. C. Biesinger and N. S.
McIntyre, Surf. Interface Anal., 2004, 36, 15641574.
28 O. Y. Khyzhun, J. Alloys Compd., 2000, 305, 16.
29 M. P. Seah, I. S. Gilmore and S. J. Spencer, J. Electron
Spectrosc. Relat. Phenom., 2001, 120, 93111.
30 S. H. Shen, C. X. Kronawitter, J. G. Jiang, S. S. Mao and
L. J. Guo, Nano Res., 2012, 5, 327336.
31 J. A. Glasscock, P. R. F. Barnes, I. C. Plumb and N. Savvides,
J. Phys. Chem. C, 2007, 111, 1647716488.
32 A. Kay, I. Cesar and M. Gratzel, J. Am. Chem. Soc., 2006, 128,
1571415721.
33 S. Saremi-Yarahmadi, K. G. U. Wijayantha, A. A.
Tahir and B. Vaidhyanathan, J. Phys. Chem. C, 2009, 113,
47684778.

1332 | Phys. Chem. Chem. Phys., 2014, 16, 1327--1332

PCCP

34 J. A. Glasscock, P. R. F. Barnes, I. C. Plumb, A. Bendavid and


P. J. Martin, Thin Solid Films, 2008, 516, 17161724.
35 C. P. Leon, L. Kador, M. Zhang and A. H. E. Muller, J. Raman
Spectrosc., 2004, 35, 165169.
36 D. Bersani, P. P. Lottici and A. Montenero, J. Raman Spectrosc.,
1999, 30, 355360.
37 M. F. Daniel, B. Desbat, J. C. Lassegues, B. Gerand and
M. Figlarz, J. Solid State Chem., 1987, 67, 235247.
38 M. Barroso, C. A. Mesa, S. R. Pendlebury, A. J. Cowan,
T. Hisatomi, K. Sivula, M. Gratzel, D. R. Klug and J. R. Durrant,
Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 1564015645.
39 R. L. Spray, K. J. McDonald and K. S. Choi, J. Phys. Chem. C,
2011, 115, 34973506.
40 S. H. Kang, J. Y. Kim, Y. Kim, H. S. Kim and Y. E. Sung,
J. Phys. Chem. C, 2007, 111, 96149623.
41 B. Klahr, S. Gimenez, F. Fabregat-Santiago, J. Bisquert and
T. W. Hamann, J. Am. Chem. Soc., 2012, 134, 1669316700.
42 H. Dotan, K. Sivula, M. Gratzel, A. Rothschild and
S. C. Warren, Energy Environ. Sci., 2011, 4, 958964.

This journal is the Owner Societies 2014

Vous aimerez peut-être aussi