A Perspective On Solar-Driven Water Splitting With All-Oxide Hetero-Nanostructures

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Energy &

Environmental Science

View Article Online / Journal Homepage / Table of Contents for this issue

Dynamic Article Links <

Cite this: Energy Environ. Sci., 2011, 4, 3889

PERSPECTIVE

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

www.rsc.org/ees

A perspective on solar-driven water splitting with all-oxide


hetero-nanostructures
Coleman X. Kronawitter,ab Lionel Vayssieres,c Shaohua Shen,d Leijin Guo,d Damon A. Wheeler,e Jin Z. Zhang,e
Bonnie R. Antounf and Samuel S. Mao*ab
Received 20th July 2011, Accepted 18th August 2011
DOI: 10.1039/c1ee02186a
A perspective on the design of all-oxide heterostructures for application in photoelectrochemical cells for
solar water splitting is provided. Particular attention is paid to those structures which possess nanoscale
feature dimensionality, as structures of this type are most likely to utilize the benefits afforded by the
formation of oxide heterojunctions and likely to show functional behavior relating to the interfacial
region. In the context of this discussion, a novel hetero-nanostructure array, based on quantum-confined
and visible light-active iron(III) oxide nanostructures and their surface modification with tungsten(VI)
oxide, is introduced. The heterostructure architecture is designed to combine the functionality of the
consituent phases to address the primary requirements for electrodes enabling the efficient generation of
hydrogen using solar energy: visible light activity, chemical stability, appropriate bandedge
characteristics, and potential for low-cost fabrication. Photoelectrochemical characterization for solar
hydrogen/oxygen generation indicates the presence of unexpected minority carrier transfer dynamics
within the oxide hetero-nanostructures, as observed additionally by ultrafast transient absorption
spectroscopy.
a
University of California at Berkeley, Department of Mechanical
Engineering, 6141 Etcheverry Hall, Berkeley, CA, 94720, USA. E-mail:
[email protected]; Fax: +510-486-7303; Tel: +510-486-7038
b
Lawrence Berkeley National Laboratory, Environmental Energy
Technologies Division, 1 Cyclotron Rd MS 70-108B, Berkeley, CA,
94709, USA
c
National Institute for Materials Science, International Center for
Materials NanoArchitectonics, Namiki 1-1, Tsukuba, Ibaraki 305-0044,
Japan
d
Xian Jiaotong University, State Key Laboratory of Multiphase Flow in
Power Engineering, Xian, Shaanxi, 710049, China
e
University of California at Santa Cruz, Department of Chemistry and
Biochemistry, Physical Sciences Building, Steinhart Way, Santa Cruz,
CA, 95064, USA
f
Sandia National Laboratories, P.O. Box 969 MS 9042, Livermore, CA,
94551-0969, USA

1. Introduction
Metal oxide heterostructures can be engineered to possess
functionality that results both from the bulk properties of their
constituent phases as well as from the emergent properties that
relate directly to the electronic and atomic character of their
interfaces. The study of oxide heterostructures has led to the
discovery of a number of new and technologically promising
interfacial phenomena, including magnetism from non-magnetic
materials,1 electronic conductivity from insulators,2 and emergent ionic conductivity.3
In addition to their prospective application as electronic and
magnetic device components, oxide heterostructures have been

Broader context
Metal oxide nanomaterials possess a wide range of functional optoelectronic properties, which relates to the diversity of cation
oxidation states, crystal structures, and electronic configurations associated with oxides and especially with the oxides of transition
metals. Utilizing this diversity by combining dissimilar oxides in hetero-nanostructure materials allows for design of photoelectrodes
specialized to perform the optical, electronic, and chemical functions required of photoelectrochemical (PEC) solar water splitting
cells. This article provides a perspective on the recent use of all-oxide hetero-nanostructures in photoanodes whose function is to
oxidize water using solar energy. In the context of these recent studies, experimental results that are representative of the all-oxide
hetero-nanostructure photoelectrode design strategy are presented. Arrays of alpha-phase iron(III) oxide quantum rods forming
a heterojunction with nanostructured tungsten(VI) oxide are studied as photoanodes in PEC cells. These results as well as the broad
set of phenomena associated with the fabrication of oxide heterostructures are discussed in terms of the design of PEC devices
enabling the solar-driven generation of hydrogen fuel.
This journal is The Royal Society of Chemistry 2011

Energy Environ. Sci., 2011, 4, 38893899 | 3889

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

applied as photoactive components of photoelectrochemical


(PEC) devices devoted to the clean and sustainable generation of
hydrogen from water using sunlight. The motivation for these
applications relates to the diversity of cation oxidation states,
crystal structures, and electronic configurations, and to the
concomitant material properties, associated with oxides and
especially with the oxides of the transition metals. Utilizing this
diversity by combining dissimilar oxide materials allows for
design of photoelectrodes specialized to perform the optical,
electronic, and chemical functions required of PEC solar water
splitting cells.
This perspective focuses on the design of all-oxide heterostructure photoanodes whose function is to enable the photooxidation of water, a critical and performance-limiting half
reaction in the overall splitting of water to produce hydrogen and
oxygen gas.4 Particular attention is paid to those heterostructures
with nanoscale dimensionality, as structures of this type are most
likely to utilize the benefits afforded by the formation of oxide
heterojunctions and likely to show functional behavior relating
to the interfacial region. For detailed analyses of alternative
designs, readers are referred to the comprehensive and everexpanding literature of this field, which most recently includes
books,5 general and comprehensive reviews,6,7 including progress
in a-Fe2O3 electrodes,8,9 discussions of composite systems
incorporating catalytic phases,10 and summaries of the use of
nanomaterials,11 among others.

2. Motivation for water oxidation at nanoscale oxide


heterostructure photoanodes
2.1 Oxide nanostructures
The in-depth consideration of physical processes that occur on
nanometer length scales is the defining principle guiding
advances in the next generation of solar energy conversion
technologies.12 The application of nanoscience to solar energy
conversion appears to represent a pivotal step toward implementing cost-effective energy conversion schemes.5,6,1318 Recent
notable contributions in this area include the study of Si nanowire arrays for photovoltaics19,20 the fabrication of efficient
dendritic a-Fe2O3 nanostructures for water oxidation,8 and the
use of plasmonic resonance modes to enhance PEC performance.21 Most important to the success of these endeavors is the
design of new materials, structures, or architectures that can
utilize a fast-growing understanding of nanoscale optoelectronic
phenomena, and which can potentially bypass or surpass certain
constraints that arise from consideration of conventional
materials for this application.
Metal oxide nanomaterials in particular have been extensively
investigated for application to solar energy conversion for their
diversity of optical properties, electronic and ionic transport
properties, chemical and thermal stability, ease of fabrication,
and relative low cost. When integrated into solar energy
conversion devices, single-crystalline one-dimensional (1-D)
metal oxide structures optimize crucial operation processes by
increasing optical path lengths and providing direct electronic
carrier transport pathways. Aligned oxide nanostructures can
now be fabricated onto large-area substrates using controllable
and inexpensive techniques.22
3890 | Energy Environ. Sci., 2011, 4, 38893899

Photoactive oxide nanomaterials - those with suitable bandgaps to absorb visible light - are stable when irradiated and
biased in most chemical environments, and are therefore
compatible with a number of PEC device configurations for solar
electricity and renewable fuel production. This includes semiconductor-based PEC cells for water splitting or electricity
generation23 as well as dye-24 and quantum dot-sensitized12 solar
cells. However, no single photoactive metal oxide material has
proven to possess the required optoelectronic properties to efficiently convert solar light energy into electrical or chemical
energy (although recent promising results indicate progress in
this direction25).
2.2.

Oxide heterostructures

A summary of the various motivations for design of oxide


heterostructures for solar water oxidation is presented in Fig. 1.
As has been established in the decades of research in tandem
photovoltaics, placing in series multiple optical absorbers has the
potential to dramatically increase the overall conversion efficiencies of solar energy conversion devices. Because of the known
thermodynamic requirements7 for the photo-oxidation of water,
it is likely a successful heterostructure will involve successive
absorption of UV and visible (blue, green, and yellow) light.
Optical absorption enhancements effected by multiple
bandgaps translate to conversion efficiency enhancements only if
the band structure alignments permit efficient utilization of
photoexcited charges. In solid-state nanostructured photovoltaic
devices, heterostructures comprised of photoactive semiconductors exhibiting Type II band offsets are utilized to
promote the efficient separation of photogenerated charges
under irradiation.26 Similarly, in dye- and quantum dot-sensitized solar cells, heterojunctions of transparent oxides with
differing electron affinities are utilized to suppress detrimental
back reactions (or charge recombination). By directing charges
to lower energy states physically separated from the photoexcited
dye or quantum dots, this technique has led to efficiency
enhancements in both nanoparticulate27 and nanorod28
configurations.

Fig. 1 Illustrations depicting the primary processes responsible for


performance enhancements in oxide heterostructure photoelectrodes.

This journal is The Royal Society of Chemistry 2011

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

Band offsets among photoactive oxides can promote charge


separation if the constituent phases and their junctions provide
the necessary conduction pathways for efficient operation of the
electrochemical cell. Establishing efficient conduction paths in
oxide heterostructure photoanodes is complicated generally by
the orbital composition of most relevant oxides occupied electronic structures: valence band edges are primarily of O 2p
character, resulting in little deviation in ionization potential
among the oxides.29 Their common orbital character creates
difficulty in establishing valence band offsets that facilitate hole
separation during operation of the oxide heterostructure photoanode. However, because the oxides conduction bands are
generally cationic in character,29 this observation contrasts with
that found for electron transfer across heterojunctions. The
recent development of highly efficient Cu2O-based heterostructure photocathodes30 highlights the additional versatility
provided by energy bands whose chemical character is primarily
cationic.
There is considerable evidence available suggesting that the
surface modication of oxide photoanodes with ultrathin oxide
coatings has the potential to effect enhanced kinetics for water
oxidation at the oxide-electrolyte interface. Ultrathin coatings
permit electron tunneling, passivate dangling bonds, or possess
electronic structures that differ from corresponding bulk structures. These configurations potentially alleviate restrictions
associated with incompatible band alignments.
Finally, it is suggested here that the fabrication of oxide
hetero-nanostructures possessing high interfacial areas is likely
to enable the development of materials with entirely new electronic structures, with associated electronic and optical properties and thus technological potential. As will be elaborated
upon later in this report, the powerful combination of nanoscience and oxide interface engineering may potentially lead to
development of new interface-property-driven electrode
structures.

3. Discussion of selected recent systems


There is considerable content in the literature that suggests the
benefits described above indeed are associated with performance
enhancements for oxide photoanodes in PEC cells. Table 1

provides a summary of selected recent systems from the literature


that highlight these findings. The table also includes the charge
transfer among oxide phases suggested in the study. This
summary is by no means comprehensive and is meant solely as an
introduction to the recent research efforts dedicated to the design
of oxide heterostructures for this application.
3.1 Nanoscale architectures
Various heterostructure architectures possessing nanoscale
feature dimensionality show promise to effectively utilize the
benefits described above. The concept of extremely-thinabsorbers,31 as established for photovoltaic devices, has been
successfully applied in designs for oxide heterostructure photoanodes. In short, by placing the absorbing phase in intimate
contact with a nanostructured conductive substrate, this technique permits the use of materials with small carrier diffusion
lengths but large optical absorption lengths. Glasscock32 fabricated and characterized core-shell nanorod ZnO-a-Fe2O3
heterostructures using a combination of aqueous chemical
synthesis and vapor phase deposition. Their characterization
revealed the presence of a-Fe2O3 / ZnO electron transfer
during irradiation, and showed some long wavelength conversion enhancements associated with the architecture. The analysis
suggested a significant energy barrier for electron injection from
a-Fe2O3 into ZnO, which is consistent with expectations
considering ZnOs greater electron affinity. Sivula et al.33 utilized
a two-step atmospheric pressure CVD process to produce WO3a-Fe2O3 heterostructures, which showed PEC enhancements at
weakly absorbing long wavelengths, which was directly attributed to the presence of the nanostructured WO3 substrate.
Oblique-angle physical vapor deposition, which enables fabrication of nanostructured films using ballistic species transport
and self-shadowing effects, has recently been used to fabricate
WO3-TiO2 core-shell nanostructures.34 Charge transfer between
the phases was observed and was associated with overall efficiency enhancements. A related nanoscale architecture was
previously achieved by coating TiO2 nanotubes with WO3 by
electrochemical deposition.35
WO3 and BiVO4 were combined in a recent work by Su et al.36
to efficiently utilize both the complimentary optical absorption
of these materials and charge separation ability associated with

Table 1 Summary of selected recent oxide heterostructure systems discussed in this report
Suggested charge transfer
Heterostructure

Description

a-Fe2O3CoAl2O4
a-Fe2O3TiO2
a-Fe2O3WO3
WO3a-Fe2O3
WO3BiVO4
WO3BiVO4

Nanoparticle composite
Bilayer film
Bilayer film
Nanostructured composite
Core-shell nanostructure arrays
Bilayer film

WO3SrTiO3
WO3TiO2
WO3TiO2
ZnOa-Fe2O3

Fabrication methods

Doctor blading
Thermal growth; CVD
Solgel
APCVD
Solvothermal technique; spin coating
Spin coating; polymer-assisted direct
deposition
Bilayer film
Spin coating
Core-shell nanostructure arrays Oblique-angle deposition
Coated nanotube arrays
Electrochemical anodization and
deposition
Core-shell nanostructure arrays Aqueous chemical synthesis; filtered arc
deposition

This journal is The Royal Society of Chemistry 2011

Electron

Hole

Reference

a-Fe2O3 / CoAlO4
TiO2 / a-Fe2O3
WO3 / a-Fe2O3
a-Fe2O3 / WO3
BiVO4 / WO3
BiVO4 / WO3

CoAlO4 / a-Fe2O3
a-Fe2O3 / TiO2
none
none
WO3 / BiVO4
WO3 / BiVO4

37
42
41
33
36
38,39

SrTiO3 / WO3
TiO2 / WO3
none

WO3 / SrTiO3
WO3 / TiO2
none

40
34
35

a-Fe2O3 / ZnO

none

32

Energy Environ. Sci., 2011, 4, 38893899 | 3891

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

their junction. Because the valence band edge of BiVO4 contains


contributions from Bi s states, it is situated at higher energies
than that of WO3, and therefore provides an efficient mechanism
for the separation of photogenerated holes during photanode
operation. Their results showed considerable enhancements are
associated with the heterostructure configuration.
Nanoparticle heterostructures comprised of a-Fe2O3 and
CoAl2O4 nanoparticles have shown PEC performance enhancements over bare p-type CoAl2O4.37 It was speculated that the
differing conductivity types of the nanoparticles, organized in
three dimensions, facilitated charge separation under irradiation.
3.2 Bilayers
In systems not limited by charge transport, simple bilayer
heterostructures provide an effective means to utilize the
complimentary functionalities of metal oxides. WO3 and BiVO4
bilayers have been studied comprehensively,38,39 and in general
these configurations permit the simultaneous utilization of the
transport properties of WO3 as well as the optical absorption
properties of BiVO4.38 Similarly, WO3 has been combined with
the transparent wide-bandgap insulator SrTiO340 for the purpose
of charge separation, with consistent enhancements evident.
WO3-a-Fe2O3 bilayer films show enhancements associated with
formation of their junction.41 Interestingly, PEC characterization
of an electrode comprised of a 120 nm TiO2 surface layer on
thermally grown a-Fe2O3 was shown to exhibit energetic
contributions from both phases.42 In general, the simple geometries of bilayer heterostructures provide systems that are
suitable for fundamental analyses of prospective photoanode
materials and especially of the potential benefits of junction
formation.
3.3 Ultrathin coatings
In a disucssion of oxide heterostructures with nanoscale dimensionality it is worthwhile to comment on the influence of ultrathin surface coatings, and especially on their influence on
a-Fe2O3-based photoanodes. The poor water oxidation kinetics
of a-Fe2O3 focuses attention on the modification of surface
states, toward increasing rate constants associated with electron
transfer from hydroxyl ions in solution. It was shown that the
deposition of tungsten oxide species on the surface of a-Fe2O3
films increases the faradic rate constant for water oxidation.43
More recently the a-Fe2O3 surfaces were passivated with ultrathin Al2O344 and Ga2O345 layers, which show consistent and
positive effects on photoelectrochemical performance, including
a cathodic shift of the photocurrent onset of a-Fe2O3. In general
it can be expected that ultrathin coatings passivate surface states
and therefore reduce the probability of electron-hole recombination associated with their presence. This is especially true for
those oxides that are isostructural to a-Fe2O3.45

4. The a-Fe2O3-WO3 system


4.1 Introduction to material system
The above discussion above describes the motivation and
framework for developing effective oxide heterostructures with
nanoscale dimensionality for efficient solar water oxidation. The
3892 | Energy Environ. Sci., 2011, 4, 38893899

present section introduces and discusses in this provided context


recent experimental work from the authors laboratories.
Tungsten(VI) oxide (tungsten trioxide, WO3) and its slightly
reduced form WO3-x are intensely studied as solar energy materials for their significant photoactivity under UV and blue light
irradiation. In PEC devices, nanostructured WO3 films produce
high photocurrents for water oxidation, and are compatible with
a number of tandem designs based on multiple photosystems.46,47
The materials ability to produce high, stable photocurrents for
sea water splitting is considered an important discovery toward
sustainable and economical solar fuel production.48
Alpha-phase iron(III) oxide (hematite, a-Fe2O3) is highly
abundant in Earths crust and is widely utilized by a number of
established and global industries. Nanostructured a-Fe2O3 in
particular has proven to be a highly functional metal oxide for
a number of energy applications. It has received considerable
attention as a solar energy conversion material because its
bandgap is well-suited for solar photon absorption and its high
ionization potential yields a valence band appropriately positioned to accept charges toward the oxidation of water. The
extreme low cost and chemical stability of a-Fe2O3 make its
implementation into solar energy conversion devices highly
attractive for the development of economical solar electricity and
fuel production.
Fig. 2 provides an illustration depicting the hetero-nanostructure design. Depending on the fabrication conditions and
application at hand, these anisotropic composites might be
described as nanoparticle-sensitized, core-shell, or radial-junction structures. In the present case, both materials are photoactive under solar irradiation when applied as electrodes in
a PEC cell, as indicated in the illustration and described later in
this report. The arrays were fabricated through application of
a combination of aqueous chemical growth49 and physical vapor
deposition techniques (see Methods for details).
4.2 Physical and optical characterization
The morphology of the arrays is examined through scanning
electron microscopy and is presented in Fig. 3a-e. Fig. 3a shows
the a-Fe2O3 nanorods produced by the hydrolysis-condensation
and dehydration process.49 The width of these rod bundles is
approximately 50 nm and their length is approximately 800 nm.
As indicated in the HRTEM image in Fig. 2, they consist of
35 nm-diameter single crystals whose electrons are quantumconfined in the lateral dimension, which manifests as a cathodic
shift (vs SHE) of the conduction band and a bandgap blueshift.50 They provide a very efficient pathway for both majority
carriers (direct pathway along the vertically oriented rods) and
minority carriers (hole diffusion length match with the quantum
rod lateral dimensions). The structures after modification with
WO3 are displayed in Fig. 3b-d. The large internal surface area is
maintained and the diameter of the structures increases with
increasing deposition time indicating that the rods are coated
with a thin layer of WO3 during exposure to the deposition
source. Similar coatings resulting from nanorod exposure to laser
ablation plasmas have previously been observed for a ZnO-ZnSe
system.51 Fig. 3e shows a cross sectional view of the structures
after extended exposure. There is a slight gradient in thickness
along the length of the rod, which is attributed to
This journal is The Royal Society of Chemistry 2011

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

Fig. 2 Illustration depicting the design of metal oxide hetero-nanostructure arrays and their application as photoanodes for photoelectrochemical cells.
The HRTEM image shows an individual single-crystalline a-Fe2O3 quantum rod. The picture on the right shows the as-prepared sample (Scale bar
indicates 1 cm).

shadowing effects commonly encountered in vacuum deposition


techniques.
Optical and structural properties of the hetero-nanostructures
are provided in Fig. 3f and 3g. All samples are deep red and
transparent to lower-energy visible light (see representative
sample photograph in Fig. 2). The UV-vis-near IR spectrum of
the hetero-nanostructures is presented in Fig. 3h. The main
absorption transition around 600 nm matches closely to those of
the a-Fe2O3 nanorods,52 which is consistent with expectations
considering the larger bandgap of WO3. X-ray diffractograms
(Fig. 3h) indicate the presence of three distinct crystal phases:
orthorhombic WO3, trigonal Fe2O3 (hematite), and the tetragonal SnO2 (cassiterite) substrate.
4.3 Photoelectrochemistry
The samples have been applied as photoelectrodes in PEC cells
for solar hydrogen generation. For this application, the interest
in this system lies in the complimentary functionality offered by
the two components. When the electrode is applied as an anode
in a PEC cell for water photoelectrolysis, photogenerated holes
participate in the water oxidation reaction and electrons travel to
the back contact and enter the external circuit. The faradic rate
constant for electron transfer from hydroxyl ions at the WO3
surface is orders of magnitude greater than at the a-Fe2O3
surface (rWO3 103104 cm s1; ra-Fe2O3 0.11 cm s1),53 due to
the superior kinetics of charge injection into O 2p bands. The
modification of a-Fe2O3 film surfaces with tungsten oxide species
has been shown previously to increase the faradic rate constant
for water oxidation.43 The hole diffusion length in WO3 has been
estimated to be about 150 nm,54 two orders of magnitude higher
than that estimated for a-Fe2O3, 24 nm.55 Despite these
advantages, the theoretical maximum water splitting efficiency of
WO3 remains low due to insufficient overlap between its optical
absorption and the solar spectrum. The bandgap of a-Fe2O3 on
the other hand is nearly ideal, which results in a theoretical
This journal is The Royal Society of Chemistry 2011

efficiency of 12.9%.56 The ultrafine a-Fe2O3 nanorods utilized in


this study were first designed and fabricated by Vayssieres et al.
with the small hole diffusion length of a-Fe2O3 in mind; their
photoelectrochemistry has been documented.52,57
The hetero-nanostructure arrays and the bare a-Fe2O3 electrodes were studied in a two-electrode configuration for solar
hydrogen generation. Fig. 4a shows chopped-light currentpotential behavior in an aqueous 0.5 M NaCl solution with solarsimulated 100 mW cm2 AM 1.5G-filtered (1 sun) irradiation;
Fig. 4b provides an incident photon conversion efficiency (IPCE)
spectrum. The sodium chloride solution is chosen for its similarity to the composition of filtered sea water; under these
conditions the thermodynamically favorable electrochemical
reactions are the hydrogen, oxygen (and potentially chlorine at
high bias) gas evolution reactions. No sacrificial reagents are
employed and the solution is neutral. All PEC results presented
indicate that the arrays behave as photoanodes in these electrolytic conditions; that is, a positive photocurrent is observed at
applied anodic potential.
These PEC data suggest that a significant quantity of holes
originating from visible-light excitations is extracted before
recombining with electrons traveling to the back contact. This is
confirmed in the IPCE spectrum, which shows spectral response
beyond 550 nm. Considering the transparency of WO3 to visible
light (Egap  3 eV), these low energy excitations must occur either
in the a-Fe2O3 core, or at intra-bandgap defect, surface, or
interface states in WO3. The electrode operation is discussed
below in the context of its physical design and its potential for
possessing anomalous electronic characteristics.
The hetero-nanostructure architecture demands electron
transport through the a-Fe2O3 rods in order for the PEC cell to
register photocurrent in the external circuit. Conduction band
electron transport between the layers is therefore a topic of
interest. Results in previous reports of a-Fe2O3-WO3 composite
photoanodes suggest electron transfer occurs in either direction
under applied anodic potential.33,41 Understanding of the band
Energy Environ. Sci., 2011, 4, 38893899 | 3893

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

Fig. 3 (a) SEM images of a-Fe2O3 arrays as fabricated by aqueous


chemical growth. (b)-(d) SEM images of arrays modified by WO3 for
increasing deposition times, indicating the evolution of the structure
morphology. (e) SEM Cross-sectional view of hetero-nanostructure array
after longest deposition time. All scale bars indicate 1 mm. Insets provide
cross-sectional views at identical magnification. (f) UV-vis-near IR
optical absorptance of a-Fe2O3-WO3. (g) X-ray diffraction pattern of the
hetero-nanostructure array indicating the presence of orthorhombic WO3
(blue), trigonal Fe2O3 (red), and tetragonal SnO2 substrate (*).

alignment in the present system is complicated by the quantum


confinement effect in the a-Fe2O3 rods (Ref. 50 indicates the
electron affinity is near 0 V vs. SHE).
Simultaneously, the oxidation reaction must occur in order to
limit the recombination of free carriers, and therefore holes must
be present at the hetero-nanostructure-liquid interface to accept
charges from electrolyte species. From the viewpoint of bulk
semiconductor physics, it is expected that there is an energy
barrier of ca. 0.5 eV58 for hole injection from a-Fe2O3 into WO3,
due to the energetic difference between O 2p orbitals, of which
the valence band of WO3 is comprised, and the Fe d levels at
which holes arrive at the a-Fe2O3 surface (it is likely that holes
arrive at the surface in a band of eg levels, which are expected to
be the highest occupied orbitals in a-Fe2O353). The observation
of photoactivity originating from visible light excitations could
involve phenomena directly related to the nanoscale architecture.
Indeed, at the length scales considered here, the interfacial region
is expected to play an important role in the electronic band
construction. In transition metal oxide heterostructures, interface electronic orbital reconstruction has been reported by
a number of mechanisms, depending on the precise nature of the
3894 | Energy Environ. Sci., 2011, 4, 38893899

Fig. 4 Photoelectrochemical characterization in aqueous 0.5 M NaCl


solution: (a) Photocurrent-potential curve under chopped (0.2 s1),
100 mW cm2 AM 1.5G-filtered solar-simulated irradiation. Potential
applied vs. a Pt counter electrode. (b) IPCE spectrum with +1 V applied
versus a Pt counter electrode. Data correspond to arrays fabricated as
those in Fig. 2c.

chemical bonds at the interface.59 Considering the very high


density of interface states in this system, owing to the nanoscale
particle dimensions, it is feasible the valence band consists of
hybridized orbitals with character resulting from both metal
oxides and their surface configurations. The electronic character
of the a-Fe2O3-WO3 interface is expected to be extremely
complex, and depend on several interface system properties,
including charge transfer and orbital hybridization. Understanding these interface properties requires intense experimental
investigations, the strategies for which are being studied in the
authors laboratories.60
Alternatively, due to the physical dimensions of the a-Fe2O3
structures (ca. 5 nm lateral dimension50) and thin WO3 layers,
quantum-mechanical tunneling processes must be considered for
hole extraction from the core. The transfer of holes from a-Fe2O3
to an electrolyte through TiO2, which also has a valence band
This journal is The Royal Society of Chemistry 2011

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

edge derived from O 2p orbitals, was reported as early as 198242


and was attributed to a tunneling process. Ref. 42 also suggests
that band banding at the semiconductor-liquid interface could
potentially extend through both phases in heterostructure photoelectrodes. If this is the case during operation of the present
electrodes, depletion effects and upward band banding could
assist in overcoming energy barriers for charge transfer from aFe2O3 to the electrolyte. It is unknown to what extent band
banding exists in the a-Fe2O3-WO3 electrodes; numerous
previous studies suggest these effects are absent in nanostructured electrodes in contact with electrolytes.6163 Exciting
work is currently being conducted in this area: a recent study
from Augustynski and coworkers64 suggests that during operation proton intercalation induces structural changes in mesoporous WO3 electrodes, which results in a core-shell structure.
In the PEC configurations studied above, the electrode
photocurrent response is remarkably similar to those found in
previous studies of a-Fe2O3 nanorod arrays.52,57,65 In addition,
the photoelectrochemical characterization instrumentation
employed was not precise enough to identify differences among
samples with varying WO3 thickness. Indeed, these PEC
performance similarities most likely result from the currentlimiting effect of the undoped a-Fe2O3 core, which is expected to
be highly resistive.66 These observations suggest that more indepth optoelectronic characterization techniques are required to
elucidate the complex charge transfer processes present in
quantum-confined metal oxide hetero-nanostructures containing
d-shell electrons. To this end the hetero-nanostructures carrier
dynamics were examined by ultrafast transient absorption
spectroscopy.

The samples were pumped with 540 nm (2.29 eV) 130 fs pulses,
which excite the a-Fe2O3 core near the weak Fe 3d/3d transition.5 The transient absorbance spectrum of a representative
a-Fe2O3-WO3 sample is presented in Fig. 5a at 500 fs after
excitation. Absorption difference signals at various probe energies reflect the occupancies of electronic states with corresponding absorption energies after excitation. A broad
distribution of absorbance signals over the studied wavelength
range is evident at this delay time. Spectra of this type have been
previously observed in femtosecond relaxation studies of iron
oxide particulate suspensions.69 These studies concluded that
because the electron relaxation processes possess both a fast (1
ps) and slower (10 ps) decay over the entire probe spectrum,
they may correspond to a common physical process - for
instance, the capture of electrons by oxygen-deficient centers of
Fe3+.69 As observed below, all relaxation processes exhibit both
a sub-picosecond fast component as well as a longer component
extending to hundreds of picoseconds.
Fig. 5be show the normalized transient absorption signals for
a-Fe2O3 and a-Fe2O3-WO3 at two probe wavelengths, 579.31 nm
(2.14 eV) and 674.23 nm (1.84 eV), in two time regimes. For both

4.4 Ultrafast transient absorption spectroscopy


The ultrafast carrier dynamics of a-Fe2O3 thin films, single
crystals, and nanoparticle suspensions have been comprehensively examined.6770 It is observed that intraband carrier thermalization occurs on extremely fast times scales (< 150 fs;67 < 75
fs69). Transient absorption signals at longer times are therefore
attributed to relaxation of excited electrons to a dense manifold
of electronic states within the bandgap and valence band,
resulting from both intrinsic defects and surface states.67
However, the assignment of transient absorption signals to
specific physical correlates in this complex material system must
be conducted in the context of all available information. This
includes the many studies performed over the past few
decades,6771 including several recent advances,72,73 with
numerous signal assignments proposed. In general the analysis is
complicated by the d-d transitions present over spectral ranges
commonly probed. Analyses of carrier dynamics of a-Fe2O3
nanomaterials are made difficult in addition by high densities of
surface states, which may introduce new decay pathways.
Of particular relevance to the present analysis is a recent report
from Durrant and colleagues,73 which describes the use of a holescavenging sacrificial chemical to determine the correspondence
of the transient absorption signal at 580 nm to unoccupied
valence band states (holes) in a-Fe2O3. The following discussion
is presented in the context of this and other studies of excited
carrier dynamics in related material systems.
This journal is The Royal Society of Chemistry 2011

Fig. 5 Ultrafast transient absorption spectroscopy for hetero-nanostructure electrodes: (a) Transient difference absorbance spectra for
a-Fe2O3-WO3 before excitation (black squares) and 500 fs after excitation (blue circles). Transient absorption signals for a-Fe2O3 (black) and
a-Fe2O3-WO3 (red) at the picosecond time scale for 579.31 nm probe (b)
and 674.23 nm probe (c) and the sub-nanosecond time scale for
579.31 nm probe (d) and 674.23 nm probe (e).

Energy Environ. Sci., 2011, 4, 38893899 | 3895

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

energies, there is an ultrafast rise in signal intensity limited by the


laser pulse. Examination of the transient absorption at the
picosecond time scale (Fig. 5b,c) reveals the fast components of
the kinetics for both probe energies are very similar before and
after modification with WO3. Because the time signatures of the
signals rise and decay overlap at this scale, the corresponding
fast physical processes apparently occur irrespective of the
presence of surface modification.
Additional information is provided by the transient absorption at longer time scales (Fig. 5d,e). For all probe wavelengths
examined, following the initial fast process, there is a slow
process that persists to beyond 100 ps. The most notable feature
of these data is that the normalized transient absorption signal at
579.31 nm decays at a faster rate for a-Fe2O3-WO3 than for aFe2O3. Considering that the transient absorption signal signifies
the occupation of an electronic state at a given delay time,
a faster destruction of the absorption signal at 579.31 nm may
suggest a faster destruction (removal) of unoccupied states
(holes) from the probed system. It is observed that the accelerated decay is nonexistent for the lower energy transition recorded
at 674.23 nm. Alternatively, it is possible the decays represent
a trapping process, and the introduction of the WO3 phase
enhances the trapping rate of electrons within a particular range
of corresponding probe energies. Trapped electrons in this case
possess a low extinction coefficient, which manifests as an
attenutation of the associated probe signal.
Given the knowledge that a-Fe2O3 nanomaterials possess
many midgap states, associated with both the d levels and high
surface areas, the wavelength dependence could describe migration of photoexcited electrons from conduction bands into
shallow levels and deep traps. In this case, the associated
recombinations evolve over the spectral range probed.
Upon examination of the decay kinetics over the complete
probe spectrum, it is found that the signals attenuation difference upon creation of a a-Fe2O3-WO3 interface is determined by
the energetic position of the corresponding electronic state.
Larger probe energies, near the bandgap of a-Fe2O3, exhibit the
largest signal difference upon creation of an a-Fe2O3-WO3
interface. Indeed, the aforementioned study on operating aFe2O3-based PEC cells73 indicates that the removal of surfacetrapped holes from a-Fe2O3 corresponds to decay in 580 nm
absorption after excitation. Adoption of this assignment in the
present study suggests that modification of a-Fe2O3 nanostructures with WO3 promotes the extraction of surface-trapped
holes on picosecond time scales.
These observations, when connected to the PEC results and
band structure descriptions provided above, indicate the possible
existence of anomalous carrier transport dynamics upon
photoexcitation of the hetero-nanostructures.

5. Electronic structure of heterostructure interfaces


A number of the device performance enhancements associated
with the use of nanomaterials relate directly to high surface areas
and high densities of surface electronic states. Oxide heteronanostructures possess high interfacial areas, which can similarly
be expected to contribute significantly to their electronic structures and therefore performance of electronic devices into which
they are integrated.
3896 | Energy Environ. Sci., 2011, 4, 38893899

The mechanisms by which oxide heterostructure interfaces are


found to possess emergent functionality has been recently discussed74 in a summary of the variety of physical phenomena
associated with their fabrication. The summary proposes several
strategies to tune the interface systems degrees of freedom,
including charge transfer, epitaxial strain, symmetry breaking,
electrostatic coupling, and frustration.74 In addition to those
effects isolated at interfaces, it is noted that several phenomena
extend well away from interfaces. These phenomena include
unscreened Coulomb interactions as well as relaxations toward
bulk properties away from the interface, which result in gradients
of structure, polarization, and magnetization properties.
In the oxides of the transition metals, the d orbital character
and occupancy critically influences the interaction of metal ions,
and by extension the physical properties of these materials.
Recent experiments conducted in the authors laboratories60
indicate that the formation of heterostructures comprised of
titanium oxide and iron oxide nanomaterials, both of which are
intensely investigated for photoelectrochemical applications,
results in a number of modifications to the orbital character of
the interface. Most importantly, it was found that the d orbitals
in the interfacial region were enriched with electrons, which has
a number of implications for devices utilizing transitions to and
from their associated electronic states. In addition, through
examination of the oxygen site environment, these studies
provided evidence that the interface possessed an emergent
degree of p-d orbital hybridization that was directly associated
with formation of the heterojunction. These results provide
confirmation of applicability of the principles of transition metal
oxide interface engineering to the design of oxide heterostructure
photoelectrodes for solar water splitting cells.

Conclusions
A perspective on the design of oxide heterostructure photoanodes for solar water oxidation has been presented. Performance enhancements can be associated with increased optical
absorption, efficient charge separation across heterojunctions,
kinetic advantages, and potentially with the engineering of
interface electronic characteristics.
The demonstration of the successful combination of largescale chemical and vapor phase synthesis techniques for the
fabrication of vertically oriented metal oxide hetero-nanostructures without template, surfactant or applied fields toward
solar energy conversion is of direct relevance to optimized
(doped) nanostructured oxide materials. These results suggest
that through this novel nanoscale architecture it is possible to
engineer interesting or unexpected interfacial minority carrier
transport properties for PEC applications of metal oxide heteronanostructures. The constituent oxides are highly functional
nanomaterials for a number of applications and as such the
hetero-nanostructure array is expected to display further functionalities, as well as to pave the way to cost-effective, large scale,
hetero-nanostructure devices for solar energy conversion.

Methods
The fabrication procedure of arrays in this work consists of
application of a combination of solution chemistry and physical
This journal is The Royal Society of Chemistry 2011

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

vapor deposition techniques. The a-Fe2O3 nanorod arrays were


first fabricated onto SnO2:F-coated glass (FTO; Pilkington
TEC7) following the procedure for hydrolysis-condensation
described in detail in ref. 49. For PEC characterization, the
samples were annealed in air 700  C. The WO3 layer was fabricated by deposition of species from pulsed laser ablation of
a rotating (10 rpm) pressed polycrystalline WO3 target in O2
ambient (reactive pulsed laser deposition) using a Lambda
Physik LPX 200 KrF Excimer Laser (248 nm). A stainless steel
chamber was evacuated to a base pressure about 104 torr, after
which O2 gas (industrial grade) was flowed through the evacuated chamber for 30 min at 18 sccm, as set by a mass flow
controller, to replace residual gases. The a-Fe2O3/FTO sample
was mounted on a substrate holder using conductive Ag paste.
Deposition occurred at a laser fluence of  1 J cm2 per pulse
with pulse frequency of 9 Hz for various times. Fig. 2b-2e show
samples from deposition times of 5, 10, 16, and 16 min, respectively. The O2 pressure during deposition was approximately 4 
103 torr as measured by a pirani pressure gauge. Before deposition the target was ablated for a several minutes at the deposition laser parameters to remove any contaminants from the
target surface, during which the entire substrate was masked with
a mechanical shutter. Throughout deposition the sample was
maintained at 100  C using a resistance heater and thermocouple
embedded in the substrate holder. After deposition the chamber
was sealed and the sample was allowed to cool to room
temperature in the quiescent low-pressure O2 ambient. After the
sample was removed it was annealed in air at 400  C for 2 h (the
ramp rate to 400  C was approximately 1.5  C/min).
Physical characterization
Scanning electron microscopy (SEM) images were obtained with
a Hitachi environmental field emission scanning electron
microscope (Model S-4300SE/N) with an accelerating voltage of
35 kV, operating in secondary electron detection mode.
Spectral transmittance and reflectance measurements were
taken on the a-Fe2O3-WO3/SnO2:F/glass samples with a Perkin
Elmer Lambda Spectrophotometer fitted with an integrating
sphere over the wavelength range 375 nm to 900 nm at a wavelength interval of 2 nm. The sample was irradiated from the aFe2O3-WO3 surface. The spectral absorptance was obtained by
solution of the equation Al (100  Rl  Tl)  Al,FTO.
X-ray diffraction data were obtained with Cu Ka radiation
from a commercial diffractometer.
Photoelectrochemical characterization
The electrolyte was prepared with 18.1 MU-cm water with 0.5 M
NaCl (Sigma Aldrich; 98%). An electrical contact was made to
the SnO2:F electrode with a conductive silver paste and copper
wire. Photoelectrochemical measurements were made in an open
pyrex cell fitted with 3 mm thick quartz window. A 1 cm2
masked-off, sealed area of the sample was irradiated with a 300
W Xe bulb solar simulator with adjustable power settings
through an AM 1.5G filter (Oriel; 81092). For current-potential
measurements, the light intensity at the sample location in the
photoelectrochemical cell was 100 mW cm2 as measured by
a power detector (Newport; 70284). No correction was made for
This journal is The Royal Society of Chemistry 2011

the optical absorption of the 4 cm of electrolyte between the


quartz window and sample location. A potentiostat (Pine
Instruments Bipotentiostat) was used to measure electrochemical
data in a two-electrode setup using a coiled Pt wire counter
electrode. N2 gas was continuously bubbled in solution and
directly over the Pt counter electrode before and during the
experiment to remove any dissolved O2 and therefore suppress
the reduction of dissolved O2 at the counter electrode. For
current-potential measurements, the potential scan was anodic
(in the positive direction) and at a rate of 5 mV s1, with the light
mechanically chopped every ten seconds. To measure the IPCE
spectra, a monochromator was used to select wavelengths (10 nm
interval) from the chopped (5 Hz) output of a 70 W Xe bulb solar
simulator. The incident photon flux at each wavelength was
measured by diverting a portion of the beam with a beam splitter
into a calibrated silicon photodiode.

Ultrafast transient absorption spectroscopy


The ultrafast laser system is based on a Quantronix femtosecond
laser system,75 consisting of an Er-doped fiber oscillator,
a regenerative/multi-pass amplifier, and a diode-pumped, Qswitched, second harmonic Nd:YLF pump laser (527 nm, 10 W
capacity). Before injection into the amplifier, chirped pulse
amplification is performed to temporally stretch, amplify, and
recompress the initial short pulse, resulting in a pulse near its
original duration, albeit with a vastly higher energy level (sub-nJ
raised to mJ). After amplification, the as-generated fundamental
(795 nm) is beam-split to generate both a white light continuum
(WLC) probe pulse as well as feeding a tunable optical parametric amplifier (OPA) consisting of two delay stages: a signal
pre-amplification stage and a power amplification stage. The
output of the OPA was passed through wavelength separators in
order to achieve a tunable pump wavelength. The system operates at 750 Hz repetition rate.
The final output was ca. 130 fs pulses centered at 540 nm
excitation wavelength which was attenuated with neutral density
filters. The pump beam was overlapped spatially and temporally
with the WLC probe beam at the sample. The time delay between
the pump and probe beams was controlled by a translation stage
with 1 mm resolution. Each sample was tested for four cycles to
achieve a lower-noise average response.

Acknowledgements
This research has been partially supported by the U.S. Department of Energy, Office of Energy Efficiency and Renewable
Energy. C.X.K. and B.R.A. were supported by Sandia National
Laboratories. Sandia National Laboratories is a multi-program
laboratory managed and operated by Sandia Corporation,
a wholly owned subsidiary of Lockheed Martin Corporation, for
the U.S. Department of Energys National Nuclear Security
Administration under contract DE-AC04-94AL85000. Additional support was provided by MEXT Japan; the National
Natural Science Foundation of China; the National Basic
Research Program of China; and the China Scholarship Council.
J.Z.Z. is grateful to the Basic Energy Sciences Division of the
U.S. DOE (DE-FG02-ER46232) for support. D.A.W. was
Energy Environ. Sci., 2011, 4, 38893899 | 3897

View Article Online

supported in part by the W.M. Keck Center for Nanoscale


Optofluidics at UCSC.

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

Notes and references


1 A. Brinkman, M. Huijben, M. van Zalk, J. Huijben, U. Zeitler,
J. C. Maan, W. G. van der Wiel, G. Rijnders, D. H. A. Blank and
H. Hilgenkamp, Nat. Mater., 2007, 6, 493496.
2 A. Ohtomo and H. Y. Hwang, Nature, 2004, 427, 423426.
3 J. Garcia-Barriocanal, A. Rivera-Calzada, M. Varela, Z. Sefrioui,
E. Iborra, C. Leon, S. J. Pennycook and J. Santamaria, Science,
2008, 321, 676680.
4 L. Duan, L. Tong, Y. Xu and L. Sun, Energy Environ. Sci., 2011,
DOI: 10.1039/C1EE01276B, Advance Article.
5 On Solar Hydrogen & Nanotechnology, Ed: L. Vayssieres, John Wiley
& Sons, Singapore, 2009.
6 X. Chen, S. Shen, L. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503
6570.
7 M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi,
E. A. Santori and N. S. Lewis, Chem. Rev., 2010, 110, 6446
6473.
8 Y. Lin, G. Yuan, S. Sheehan, S. Zhou and D. Wang, Energy Environ.
Sci., 2011, DOI: 10.1039/C1EE01850G, Advance Article.
9 K. Sivula, F. Le Formal and M. Gratzel, ChemSusChem, 2011, 4,
432444.
10 J. Sun, D. K. Zhong and D. R. Gamelin, Energy Environ. Sci., 2010, 3,
12521261.
11 Y. Lin, G. Yuan, R. Liu, S. Zhou, S. W. Sheehan and D. Wang,
Chem. Phys. Lett., 2011, 507, 209215.
12 A. J. Nozik, Nano Lett., 2010, 10, 27352741.
13 H. A. Atwater and A. Polman, Nat. Mater., 2010, 9, 205213.
14 P. V. Kamat, J. Phys. Chem. C, 2007, 111, 28342860.
15 Nanostructured and Photoelectrochemical Systems for Solar Photon
Conversion, Eds: M. D. Archer, A. J. Nozik, Imperial College Press,
London, 2008.
16 X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 28912959.
17 H. G. Park and J. K. Holt, Energy Environ. Sci., 2010, 3, 1028
1036.
18 S. Shen, J. Shi, P. Guo and L. Guo, Int. J. Nanotech., 2011, 8, 523
591.
19 M. C. Putnam, S. W. Boettcher, M. D. Kelzenberg, D. B. TurnerEvans, J. M. Spurgeon, E. L. Warren, R. M. Briggs, N. S. Lewis
and H. A. Atwater, Energy Environ. Sci., 2010, 3, 1037
1041.
20 M. D. Kelzenberg, S. W. Boettcher, J. A. Petykiewicz, D. B. TurnerEvans, M. C. Putnam, E. L. Warren, J. M. Spurgeon, R. M. Briggs,
N. S. Lewis and H. A. Atwater, Nat. Mater., 2010, 9, 239
244.
21 I. Thomann, B. A. Pinaud, Z. Chen, B. M. Clemens, T. F. Jaramillo
and M. L. Brongersma, Nano Lett., 2011, 11, 34403446.
22 L. Vayssieres, A. Hagfeldt and S. -E. Lindquist, Pure Appl. Chem.,
2000, 72, 4752.
23 N. S. Lewis, J. Electroanal. Chem., 2001, 508, 110.
24 M. Gr
atzel, Inorg. Chem., 2005, 44, 68416851.
25 X. Chen, L. Liu, P. Y. Yu and S. S. Mao, Science, 2011, 331, 746
750.
26 H. Zhong, Y. Zhou, Y. Yang, C. Yang and Y. Li, J. Phys. Chem. C,
2007, 111, 65386543.
27 H. J. Snaith and C. Ducati, Nano Lett., 2010, 10, 12591265.
28 N. O. V. Plank, I. Howard, A. Rao, M. W. B. Wilson, C. Ducati,
R. S. Mane, J. S. Bendall, R. R. M. Louca, N. C. Greenham,
H. Miura, R. H. Friend, H. J. Snaith and M. E. Welland, J. Phys.
Chem. C, 2009, 113, 1851518522.
29 H. H. Kung, H. S. Jarrett, A. W. Sleight and A. Ferretti, J. Appl.
Phys., 1977, 48, 24632469.
30 A. Paracchino, V. Laporte, K. Sivula, M. Gratzel and E. Thimsen,
Nat. Mater., 2011, 10, 456461.
31 K. Ernst, A. Belaidi and R. K
onenkamp, Semicond. Sci. Technol.,
2003, 18, 475479.
32 J. A. Glasscock, PhD Thesis, University of New South Wales,
Australia, 2008, 1220.
33 K. Sivula, F. Le Formal and M. Gratzel, Chem. Mater., 2009, 21,
28622867.

3898 | Energy Environ. Sci., 2011, 4, 38893899

34 W. Smith, A. Wolcott, R. C. Fitzmorris, J. Z. Zhang and Y. Zhao, J.


Mater. Chem., 2011, 21, 1079210800.
35 J. H. Park, O. O. Park and S. Kim, Appl. Phys. Lett., 2006, 89,
163106.
36 J. Su, L. Guo, N. Bao and C. A. Grimes, Nano Lett., 2011, 11, 1928
1933.
37 K.-S. Ahn, Y. Yan, M.-S. Kang, J.-Y. Kim, S. Shet, H. Wang,
J. Turner and M. Al-Jassim, Appl. Phys. Lett., 2009, 95, 022116.
38 S. J. Hong, S. Lee, J. S. Jang and J. S. Lee, Energy Environ. Sci., 2011,
4, 17811787.
39 P. Chatchai, Y. Murakami, S.-y. Kishioka, A. Y. Nosaka and
Y. Nosaka, Electrochim. Acta, 2009, 54, 11471152.
40 Y. Wang, T. Yu, X. Chen, H. Zhang, S. Ouyang, Z. Li, J. Ye and
Z. Zou, J. Phys. D: Appl. Phys., 2007, 40, 39253930.
41 W. Luo, T. Yu, Y. Wang, Z. Li, J. Ye and Z. Zou, J. Phys. D: Appl.
Phys., 2007, 40, 10911096.
42 F.-T. Liou, C. Y. Yang and S. N. Levine, J. Electrochem. Soc., 1982,
129, 342345.
43 T. Kishi and M. Aritsuka, Surf. Coat. Technol., 1988, 34, 345
353.
44 F. Le Formal, N. Tetreault, M. Cornuz, T. Moehl, M. Gratzel and
K. Sivula, Chem. Sci., 2011, 2, 737743.
45 T. Hisatomi, F. Le Formal, M. Cornuz, J. Brillet, N. Tetreault,
K. Sivula and M. Gratzel, Energy Environ. Sci., 2011, 4, 2512
2515.
46 J. Augustynski, G. Calzaferri, J. C. Courvoisier and M. Gratzel, in
Proc. of the 11th World Hydrogen Energy Conference, Eds: T. N.
Veziroglu, C. J. Winter, J. P. Baselt, G. Kreysa, Stuttgart, 1996, p.
2379.
47 E. L. Miller, R. E. Rocheleau and X. M. Deng, Int. J. Hydrogen
Energy, 2003, 28, 615623.
48 J. Augustynski, R. Solarska, H. Hagemann and C. Santato in Proc.
SPIE, 6340, Ed. L. Vayssieres, SPIE Press, Bellingham, 2006,
63400J-1.
49 L. Vayssieres, N. Beermann, S.-E. Lindquist and A. Hagfeldt, Chem.
Mater., 2001, 13, 233235.
50 L. Vayssieres, C. Sathe, S. M. Butorin, D. K. Shuh, J. Nordgren and
J. Guo, Adv. Mater., 2005, 17, 23202323.
51 K. Wang, J. Chen, W. Zhou, Y. Zhang, Y. Yan, J. Pern and
A. Mascarenha, Adv. Mater., 2008, 20, 32483253.
52 N. Beermann, L. Vayssieres, S.-E. Lindquist and A. Hagfeldt, J.
Electrochem. Soc., 2000, 147, 24562461.
53 M. P. Dare-Edwards, J. B. Goodenough, A. Hamnett and
P. R. Trevellick, J. Chem. Soc., Faraday Trans. 1, 1983, 79, 2027
2041.
54 M. A. Butler, J. Appl. Phys., 1977, 48, 19141920.
55 J. H. Kennedy and K. W. Frese, J. Electrochem. Soc., 1978, 125, 709
714.
56 A. B. Murphy, P. R. F. Barnes, L. K. Randeniya, I. C. Plumb,
I. E. Grey, M. D. Horne and J. A. Glasscock, Int. J. Hydrogen
Energy, 2006, 31, 19992017.
57 T. Lindgren, H. Wang, N. Beermann, L. Vayssieres, A. Hagfeldt and
S.-E. Lindquist, Sol. Energy Mater. Sol. Cells, 2002, 71, 231
243.
58 M. Gratzel, Nature, 2001, 414, 338344.
59 J. Chakhalian, W. Freeland, H.-U. Habermeier, G. Cristiani,
G. Khaliullin, M. van Veenendaal and B. Keimer, Science, 2007,
318, 11141117.
60 C. X. Kronawitter, J. R. Bakke, D. A. Wheeler, W.-C. Wang,
C. Chang, B. R. Antoun, J. Z. Zhang, J. Guo, S. F. Bent,
S. S. Mao and L. Vayssieres, Nano Lett., 2011, DOI: 10.1021/
nl201944h.
61 J. Nelson, Phys. Rev. B: Condens. Matter, 1999, 59, 15374
15380.
62 F. Cao, G. Oskam, G. J. Meyer and P. C. Searson, J. Phys. Chem.,
1996, 100, 1702117027.
63 I. Cesar, K. Sivula, A. Kay, R. Zboril and M. Gratzel, J. Phys. Chem.
C, 2009, 113, 772782.
64 A. Kr
olikowska, P. Barczuk, R. Jurczakowski and J. Augustynski, J.
Electroanal. Chem., 2011, DOI: 10.1016/j.jelechem.2011.07.003.
65 Y. Ling, G. Wang, D. A. Wheeler, J. Z. Zhang and Y. Li, Nano Lett.,
2011, 11, 21192125.
66 F. J. Morin, Phys. Rev., 1954, 93, 11951199.
67 N. J. Cherepy, D. B. Liston, J. A. Lovejoy, H. Deng and J. Z. Zhang,
J. Phys. Chem. B, 1998, 102, 770776.

This journal is The Royal Society of Chemistry 2011

Published on 06 September 2011. Downloaded by University of California - Davis on 12/10/2016 05:00:17.

View Article Online

68 A. G. Joly, J. R. Williams, S. A. Chambers, G. Xiong, W. P. Hess and


D. M. Laman, J. Appl. Phys., 2006, 99, 053521.
69 V. A. Nadtochenko, N. N. Denisov, V. Y. Gak, F. E. Gostev,
A. A. Titov, O. M. Sarkisov and V. V. Nikandrov, Russ. Chem.
Bull., 2002, 51, 457461.
70 Y. P. He, Y. M. Miao, C. R. Li, S. Q. Wang, L. Cao, S. S. Xie,
G. Z. Yang, B. S. Zou and C. Burda, Phys. Rev. B: Condens.
Matter Mater. Phys., 2005, 71, 125411.
71 N. M. Dimitrijevic, D. Savic, O. I. Micic and A. J. Nozik, J. Phys.
Chem., 1984, 88, 42784283.

This journal is The Royal Society of Chemistry 2011

72 A. J. Cowan, C. J. Barnett, S. R. Pendlebury, M. Barroso, K. Sivula,


M. Gratzel, J. R. Durrant and D. R. Klug, J. Am. Chem. Soc., 2011,
133, 1013410140.
73 S. R. Pendlebury, M. Barroso, A. J. Cowan, K. Sivula, J. Tang,
M. Gratzel, D. Klug and J. R. Durrant, Chem. Commun., 2011, 47,
716718.
74 P. Zubko, S. Gariglio, M. Gabay, P. Ghosez and J.-M. Triscone,
Annu. Rev. Condens. Matter Phys., 2011, 2, 141165.
75 R. J. Newhouse, H. Wang, J. K. Hensel, D. A. Wheeler, S. Zou and
J. Z. Zhang, J. Phys. Chem. Lett., 2011, 2, 228235.

Energy Environ. Sci., 2011, 4, 38893899 | 3899

You might also like