You are on page 1of 36

Review

pubs.acs.org/CR

TiO2 Nanoparticles as Functional Building Blocks


Lixia Sang,† Yixin Zhao,‡ and Clemens Burda*,§

Key Laboratory of Enhanced Heat Transfer and Energy Conservation, Ministry of Education and Key Laboratory of Heat Transfer
and Energy Conversion, Beijing Municipality, College of Environmental and Energy Engineering, Beijing University of Technology,
Beijing 100124, China

School of Environmental Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
§
Center for Chemical Dynamics and Nanomaterials Research, Department of Chemistry, Case Western Reserve University, 10900
Euclid Avenue, Cleveland, Ohio 44106, United States
7.2.1. Improving Charge Separation with
Metal NPs 9302
7.2.2. Core−Shell Metal@TiO2 9303
7.2.3. Increasing UV and Visible Light Absorp-
tion with Plasmonics NPs 9303
7.3. TiO2−Semiconductor Nanoheterostructures 9304
8. TiO2 NPs as Charge Separation Centers 9306
CONTENTS 8.1. Photoinduced Charge Separation in TiO2
NPs 9306
1. Introduction: The Development of the TiO2 8.2. Charge Extraction for Photocatalytic Redox
Nanoparticle Research Field 9283 Reactions 9306
2. Structural Evolution of Molecule−Cluster−Nano- 8.2.1. Charge Extraction for Photocatalytic
particles 9284 Purification of Water and Air 9306
3. Crystal Phases and Phase Transformation 9285 8.2.2. Charge Extraction for Solar Water
4. Synthesis and Characterization of TiO2 NPs 9287 Splitting 9307
4.1. Novel Aspects to the Synthesis Methods of 8.2.3. Charge Extraction for Photocatalytic
TiO2 NPs 9287 Reduction of CO2 9308
4.1.1. Chemical Synthesis of Sub-2-nm TiO2 8.3. Charge Separation and Injection in Grätzel-
NPs 9287 type Solar Cells 9309
4.1.2. Low-Temperature Synthesis of Sub-5- 9. Outlook 9310
nm TiO2 Nanoparticles 9287 Author Information 9310
4.1.3. Nonaqueous Preparation of TiO2 NPs 9288 Corresponding Author 9310
4.1.4. Controlling Shapes and Crystal Facets of Notes 9310
TiO2 9289 Biographies 9310
4.2. Characterization Approaches 9290 Acknowledgments 9311
4.2.1. Structural Properties via XRD and References 9311
Raman Spectroscopy 9290
4.2.2. Optoelectronic Properties via UV-DRS,
XPS, EPR, IR, and PL Spectroscopy 9290 1. INTRODUCTION: THE DEVELOPMENT OF THE TIO2
4.2.3. Laser Spectroscopic Characterization of NANOPARTICLE RESEARCH FIELD
Charge Carrier Dynamics 9292
TiO2 has been commercially manufactured by the millions of
5. TiO2 NPs as Growth Centers: Evolution of TiO2-
tons to be widely utilized as pigment, paint additive, and
Based Nanomaterials 9294
sunscreen to name a few uses, due to its photostability and light
5.1. One-Dimensional (1-D) Growth 9294
dispersion, yet simultaneously strong UV light filtering. TiO2
5.2. TiO2 with Higher Organized Architectures 9296
has also been one of most investigated engineering materials
5.3. Breaking One-Dimensional TiO2 Arrays into
during recent decades, especially in the arena of energy and
Zero-Dimensional TiO2 NPs 9296
environmental applications. The momentum of this research
6. Chemical and Physical Modifications of TiO2 NPs 9297
and its historical development has been significantly impacted
6.1. TiO2 NPs as Hosts for Metal and Nonmetal
by two milestone research reports in 1972. In 1972, Fujishima
Dopants 9297
and Honda published the finding of photocatalytic splitting of
6.2. Codoping 9298
water on a TiO2 electrode under ultraviolet (UV) light.1 The
6.3. Photonic Crystals Constructed from TiO2 NPs 9299
discovery of this phenomenon spurred a tremendous amount of
6.4. Ti3+ Self-Doped TiO2 9300
7. TiO2 NP Heteronanostructures 9301
7.1. Mixed Phase TiO2 9301 Special Issue: 2014 Titanium Dioxide Nanomaterials
7.2. TiO2−Metal Nanostructures 9302 Received: November 5, 2013
Published: May 20, 2014

© 2014 American Chemical Society 9283 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

research related to TiO2 photocatalysis including solar fuels and an unusual scientific explosion of this filed since the beginning
environmental remediation. Second, and also in 1972, of this millennium. Accordingly, we focused this Review on
Tributsch demonstrated the idea of a dye-sensitized solar cell work published past the year 2000 and particularly on work
(DSSC), fabricating a chlorophyll-sensitized zinc oxide (ZnO) within the last 5 years. Astonishing progress has been reported
electrode to convert visible light radiation into an electric since 2007 warranting a fresh look at the field as a whole.
current by charge injection from the excited dye molecules into Hereby we attempted to emphasize newer research topics. We
the wide band gap metal oxide.2 Although in this first DSSC start with the structural evolution of TiO2 NPs (section 2),
concept report the author used ZnO, not TiO2, as the wide current knowledge about TiO2 phases and phase tranforma-
band gap semiconductor, TiO2 was soon to become the most tions at the nanoscale (section 3), and novel insights into the
popular wide band gap semiconductor material for later DSSCs, preparation and characterization of TiO2 NPs (section 4). The
mainly due to its better photostability. second half of this review (sections 5−8) will summarize
These two reports mark the start of a remarkable modifications and functionalizations recently undertaken on
development toward the promise of utilizing TiO2 as a TiO2 nanocrystals. We review how TiO2 NPs have been used
photoactive catalyst for solar fuels, photovoltaics, and environ- to prepare larger, more complex stuctures (section 5), how
mental remediation. These three applications are still now the chemical and physical modifications were used to enhance the
main research areas for TiO2. In all these applications, TiO2 has properties of TiO2 (section 6), and the recent use of TiO2 in
been used mainly in three important functions: the first is to heteronanostructures (section 7). In section 8 we highlight the
work as a photochemical energy conversion material due to use of TiO2 NPs as charge separation centers. A brief outlook
their electronic structure in the case of photocatalyts; the in section 9 concludes this review, keeping in mind that more
second is to function as the photosensitizer substrate due to its in-depth treatments to more focused topics follow in this
surface area and surface stability in photosynthesis applications; Thematic Issue on TiO2.
and the third is to serve as the electron transport scaffold in
photovoltaic applications due to its electrical properties. The 2. STRUCTURAL EVOLUTION OF
research findings derived from these investigations have then MOLECULE−CLUSTER−NANOPARTICLES
been utilized for other advanced applications such as sensing, Compared to the bulk crystalline phases of TiO2, nanosized
biomedicine, and electronics.3−11 particles are smaller and often composed of strained lattices due
Future advanced uses of TiO2 clearly depend on the to the large surface-to-volume ratio. In addition, mixed phases
development of more complex materials with advanced and interfaces incorporating a number of defects are quite
functionalities. Materials that are purpose-built to perform common. Clusters, on the other hand, are composed of far
specific tasks are the targets of such current research. For fewer atoms, in a size range that can actually be computed and
example, for photocatalysis TiO2 has the advantage of good understood atom by atom. Global minima structures for the
photocatalytic activity and photostability and low cost. discrete (TiO2)n clusters were predicted as stable or low-energy
However, one limitation of TiO2 as a photocatalyst is its metastable structures.12 Numerous experimental and theoretical
wide band gap, making TiO2 only sensitive to the UV light studies investigated isolated titanium oxide clusters to correlate
which covers less than 5% of the solar spectrum. To overcome their structures and properties with those of the bulk
this intrinsic disadvantage, two strategies have been pursued. phases.12−16 The TinO2n and TinO2n+1 clusters were found to
The first is to broaden the active spectrum of TiO2 by a variety be the most stable neutral clusters, while TinO2n‑1 and TinO2n‑2
of chemical modifications that adjust its electronic structure. clusters were formed by fragmentation.17 Titanium oxide
The other strategy, aimed at improving TiO2 utility, is to create cluster cations, TinO2n‑m+ (n = 1−8; m = 0−4), were observed
high surface area TiO2. In photocatalysis, the photocatalytic by sputtering titanium foil exposed to oxygen, and for n = 1−7
conversion scales with the surface area. and m = 1−3 by sputtering titanium dioxide powder.18 As initial
For the DSSC application, an ideal TiO2 photoelectrode parameters, the experimental and computational data used for
should have a large enough surface area to load sufficient TiO2 molecules are a Ti−O bond length of ∼1.62 Å and a O−
sensitizer dye onto the TiO2 anode for efficient solar energy Ti−O angle of ∼110°. The structural motifs of the (TiO2)n
conversion. The breakthrough in DSSC efficiency was reported nanoparticles are found to substantially differ from those of the
by Grätzel and co-workers in 1991, reporting work in which the bulk TiO2 (Figure 1).19
mesoporous TiO2 NP photoelectrode replaced a bulk TiO2 The average Ti−O bonding length in these clusters is smaller
photolectrode.3 Since then, a tremendous number of variations than in the bulk. They have a more compact structure because
of nanostructured TiO2 electrodes have been developed to of the high surface-to-bulk ratio. They are not molecular
achieve high efficiency DSSCs. In search of ever-improving structures anymore and still lack the strict periodicity of the
performance for different applications, nanostructured TiO2 bulk crystals. The changes in the coordination of the atoms, the
became one of the most investigated solid-state materials of the size and shape of the cluster, and the amount of (TiO2) units all
past 10 years. Zero-dimensional (0-D) TiO2 NPs are the most bring about sensitive changes in the electronic structure and the
basic nanostructures of TiO2 suitable for large-scale production. energy gap.20 A spectral blue-shift and effective band gap
They can be utilized as starting points for more complex broadening are observed when the size of the semiconductor
materials with more specified and enhanced performance particles becomes smaller than the exciton radius of ∼1 nm,
parameters. resulting in a confined bound state of the hole and electron.21,22
In the following Review, we show why TiO2 NPs are central Through density functional theory (DFT) calculation of
for the development of future designer materials involving (TiO2)n clusters (Figure 2), one finds a mixture of O(2p) and
titania, how they are currently grown and incorporated, and Ti(3d) atomic orbitals for all frontier molecular orbitals of the
how material properties are affected by TiO2 NPs. smaller clusters. However, in the larger clusters, there is a clear
While the start of this development was initiated by spatial separation, whereby the HOMO and HOMO-1 are
milestones of research in 19721,2 and 1991,3 there has been composed of O(2p) orbitals and the LUMO and LUMO+1 are
9284 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

dramatically from an extended-semiconducting structure to


nanometer-sized small clusters of titanium oxide.25

Figure 1. Relaxed geometries of (TiO2)n clusters (a) (TiO2), (b)


(TiO2)2, (c) (TiO2)3, (d) (TiO2)4, (e) (TiO2)5, (f) (TiO2)6, (g)
(TiO2)7, (h) (TiO2)10, and (i) (TiO2)13. The purple spheres represent Figure 3. Illustration of the electronic structure changes in titanium
the Ti atoms while the red spheres represent the O atoms.19 oxide when moving from semiconducting bulk TiO2 to isolated Ti−
Reproduced with permission from ref 19. Copyright 2010 American Oxide clusters.25 Reproduced with permission from ref 25. Copyright
Chemical Society. 2012 Royal Society of Chemistry.

Qu et al.26 have studied the electronic structure and stability


of defect-free (TiO2)n NPs with n = 10−16 (similar to and
slightly larger in diameter than the 1 nm exciton Bohr radius of
titania). Even-n (TiO2)n clusters consist of only 4-coordinated
Ti(4) and 2-coordinated O(2−) atoms with compact covalent
networks that are more stable, while odd-n clusters tend to
form more ionic structures with additional Ti(6), Ti(5), O(3), and
O(4) atoms. These theoretical studies suggested that (TiO2)n
clusters with odd and small n values would help to engineer the
material with visible light photoactivity because the most stable
Figure 2. Calculated frontier molecular orbitals for (TiO2)n with n = 6 clusters with odd n exhibit small vertical excitation energies.
(a, b), n = 7 (c, d), n = 10 (e, f), and n = 13 (g, h) clusters. Purple and TiO2 heterofullerenes using Ti2O4 units as the basic building
red spheres represent the titanium and oxygen atoms. Superimposed blocks have been proposed due to their structural stability and
and semitransparent, HOMO-1 and HOMO orbitals are colored blue
appropriate growth properties.19 These defect-free nanocages
and orange, respectively, whereas, in the other panels, green and
yellow are chosen for the LUMO and the LUMO+1 orbitals, display unique frequency modes at ∼827−854 cm−1 and a
respectively. In comparison, the HOMO and LUMO orbitals for bulk larger HOMO−LUMO energy gap than bulk TiO2 materials.27
rutile (not shown) are uniformly distributed over the material, with the During the synthesis of titanium dioxide nanoparticles,
HOMO residing solely on the oxygens (2p) and the LUMO on the hydrolysis causes the individual TiO2+ or Ti(OH)22+ ions to
titanium (3d).19 Reproduced with permission from ref 19. Copyright polymerize via olation and oxolation to (TiO2)n− chains.20
2010 American Chemical Society. These chains tend to agglomerate and polymerize into small
particles of several nanometers. Zhai et al.28 have probed the
electronic structure and band gap evolution of titanium dioxide
composed of Ti(3d) orbitals. Thus, direct excitation from the clusters, (TiO2)n− (n = 1−10), using photoelectron spectros-
HOMO to the LUMO would have small oscillation strength. copy (PES) (Figure 4). Electron detachment energies between
Furthermore, Shevlin et al.19 conclude that the formations of the ground state of the neutral and its first excited state (band
single valent oxygen atoms in subnanometer TiO2 clusters are gap) have been shown to be strongly size-dependent for n < 6,
responsible for the narrowing of the energy gap. For the and to rapidly approach the bulk limit for n > 7. The broad PES
ground-state structures of the (TiO2)n (n = 1−4) clusters, a features observed have been attributed to the localized nature
common feature among the calculated energy minima is the of the extra electron in the (TiO2)n− clusters. The extra
presence of two bridge oxygen atoms between each pair of electron localized in a trivalent Ti of (TiO2)n− (n > 1) clusters
adjacent titanium centers.23 In fact, the structures of these small creates a single Ti3+ site, which can provide clusters with
clusters are closer to that of the anatase phase. This observation molecular properties suited for mechanistic studies of TiO2
is consistent with the experimental and theoretical results that surface defects and photocatalytic activity.
the anatase phase is more stable than the rutile phase when the
particle size is below ∼14 nm.24 The calculated adiabatic energy 3. CRYSTAL PHASES AND PHASE TRANSFORMATION
gaps are ∼2.2 eV for the monomer, 2.3−3.0 eV for the low- Titanium dioxide can exist in one of three bulk crystalline
lying structures of the dimer, 2.4−2.6 eV for the trimer, and forms, rutile, anatase, and brookite, all of which can be
2.1−3.5 eV for the tetramer23 Most of these gaps are below the described in terms of distorted TiO6 octahedra with different
band gap of the bulk material. Therefore, properly designing symmetries or arrangements. The anatase structure consists of
the particle size and structure is crucial to adjust and optimize edge-sharing TiO6 octahedra, while the rutile and the brookite
the optical and photocatalytic activity of TiO2 nanoparticles. frameworks exhibit both corner and edge-sharing configura-
Figure 3 highlights that the electronic properties change tions (Figure 5).29 The different characteristics of the Ti−O
9285 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 4. Photoelectron spectra (left) of (TiO2)n− (n = 5, 10) at 193 nm (6.424 eV) and 157 nm (7.866 eV). The 193 nm spectra are shown as
insets. Observed energy gaps (right) of (TiO2)n− (n = 1−10), as measured from the adiabatic detachment energy (ADE) difference between the X-
and A-bands. The TiO2 bulk limits (rutile, 3.0 eV; anatase, 3.2 eV) are shown as horizontal dashed lines.28 Reproduced with permission from ref 28.
Copyright 2007 American Chemical Society.

variations has been used to determine the temperature at the


onset of the nucleation process through the whole crystal-
lization process.40 Zhang41,42 investigated the mechanism of the
size-dependent outer/inner phase transformation in ultrafine
TiO2 particles with narrowed size distribution. It is found that
particle size is the critical parameter determining the onset
transition temperature and nucleation behavior. From the
proposed model (Figure 6), the phase transformations take

Figure 5. Representations of the TiO2 anatase, rutile, and brookite


forms.29 Anatase (tetragonal, a = 3.785 Å, c = 9.513 Å), rutile
(tetragonal, a = 4.593 Å, c = 2.959 Å), and brookite (orthorhombic, a
= 9.181 Å, b = 5.455 Å, c = 5.142 Å). Reproduced with permission
from ref 29. Copyright 2010 American Chemical Society.

bonds play an important role in the structural and electronic


features of the different phases.30 Brookite TiO2 is a more
exotic titania polymorph with a layered structure.
Diebold et al.31 have comprehensively reviewed the bulk
properties of TiO2 with different crystal structures. Phase
transformations between different phases of TiO2 have been
extensively studied from both scientific and technological Figure 6. Proposed scheme for the phase transformations of TiO2 with
points of view.32−34 Rutile is the only stable phase in the bulk particle size (a) smaller than 10 nm, (b) in the range 10−60 nm, and
form, while bulk brookite and bulk anatase are metastable and (c) larger than 60 nm (blue, anatase; red, rutile).41Reproduced with
transform irreversibly to rutile upon heating.35 However, permission from ref 41. Copyright 2006 American Chemical Society.
different phase stability can be expected at the nanoscale
structures. At temperatures ranging between 325 and 750 °C,
anatase is the most stable phase at particle sizes under 11 nm, place at lower temperatures for smaller particles (<10 nm).
brookite is the most stable phase between 11 and 35 nm, and With the increase in initial particle size, the transformation
rutile is the most stable phase at all particle sizes above 35 temperature increases. Rutile nucleates at interfaces of
nm.24 The different phase stability in TiO2 nanoparticles is thecontacting anatase grains (<60 nm). When the particle
related to their physical environment and the interaction size is larger than 60 nm, the free surface, interface, and bulk are
between TiO2 and H2O. Molecular dynamic simulations by all likely to become rutile nucleation sites. The phase transition
Koparde35 show that rutile is the most stable phase for smaller of anatase to rutile can be prevented with the addition of other
TiO2 NPs at higher temperatures in vacuum. During synthesis metal oxide layers within the TiO2 NPs, such as ZnO,43
in liquid media, the interaction between amorphous and Nd2O3,44 and Al2O3.45 For TiO2 nanoparticles modified with
anatase phases can destabilize anatase crystals and favor Nd2O3 (Nd−TiO2), Nd3+ ions highly disperse onto the Nd−
anatase-to-rutile transformation as the aging of the solution TiO2 nanoparticle surface in the form of Nd2O3 crystallites
progresses.36 Phase stability and transformation in TiO2 NPs in which can inhibit the growth in crystal size, and anatase-to-
aqueous solutions can be guided by surface energy.37,38 DFT rutile phase transformation of TiO2.44 Similar phenomena were
calculations on different phase TiO2 NPs explain that the also found in TiO2 NPs doped with rare earth elements like
stability reversal in the nanoparticle relative to the bulk phase is La,46 Ce,46,47 and Pr.48 Dopants play a significant role in the
due to the lower surface energies in anatase compared to that in selective crystallization of the anatase phase. Other than rare
rutile at the nanoscale.39 Surfaces, edges, and vertices induce a earth elements, it is found that the introduction of Fe2+ at
much higher energetic penalty in rutile than in anatase, which higher concentrations leads to an anatase-to-rutile conversion at
are relatively more abundant and can stabilize anatase at the room temperature for the seeded TiO2−xNx nanocolloids. This
nanoscale. In the nucleation and growth process of TiO2 NPs, process is associated with the detectable formation of Ti(III)
conclusive evidence of local and intraparticle ordering and Ti(II) at the surface of the titania NPs.49
9286 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 7. (A) Schematic representation of the synthesis of TiO2 NPs and the structures of 1 and 2 used in synthesis of sub-2-nm TiO2 particles.
Yellow spheres in sphere 2 denote Pd(II) ions. Adopted from ref 63. (B) Band gap measurements for sub-nanometer-sized TiO2 particles 6 × TiO2
(green), 14 × TiO2 (red), and 30 × TiO2 (blue) reported in ref 64. Reproduced with permission from refs 63 and 64. Copyright 2013 American
Chemical Society. Copyright 2008 Nature Publishing Group.

In general, rutile and anatase are the two most dominant solution-based preparation processes.61 In the following, we
polymorphs used in applications of TiO2, with anatase showing summarize more recent progress in the synthesis of TiO2 NPs.
a higher photocatalytic activity. However, nanosized rutile TiO2 4.1.1. Chemical Synthesis of Sub-2-nm TiO2 NPs. TiO2
may show high activity compared with anatase TiO2, even particles sized 10−50 nm can be easily prepared by many well-
superior activity compared to that of Degussa P25 if the developed conventional approaches. However, significant novel
synthetic route can be controlled.50 In comparison, the brookite chemical and optical applications have recently been explored
structure is not experimentally investigated as much as the in the size range of only a few nanometers.40 Although
others because brookite is a metastable phase with a theoretical studies have predicted quantum-confinement effects
complicated and low-symmetry structure; furthermore, the in TiO2 nanoclusters, it is difficult to observe the quantum
formation of brookite TiO2 is almost always accompanied by confinement in most reported TiO2 nanocrystals because the
the presence of secondary phases such as anatase and/or rutile. exciton Bohr radius of TiO2 is just ∼1 nm.62 Other than the
Nevertheless, an increasing number of works have been focused difficulty to synthesize sub-2-nm TiO2 particles, the wider band
on brookite TiO2 in recent years.51−55 The kinetic and phase gap of such TiO2 NPs due to the quantum-confinement effect
transformation mechanisms of brookite nanoparticles were will move their light absorption into the UV region, which is
investigated by studying the annealing behavior of brookite undesired for most photorelated applications. Consequently,
nanoparticles at various temperatures.56 Indeed, anatase and there are only a few reports of sub-2-nm TiO2 clusters.63,64 In
brookite are both metastable TiO2 polymorphs, which can these reports, the TiO2 nanoclusters were synthesized through
convert into rutile at high temperature. During the brookite to a metal−organic molecular synthesis approach using molecular
rutile phase transformation, anatase was observed as transitory Ti precursors rather than those common in most inorganic
phase with surface stabilization phenomena. The brookite TiO2 synthesis approaches. In addition, the resulting small
thermal stability can be improved by the use of surface TiO2 clusters are embedded in a matrix to stabilize them. One
additives.56 On the basis of the synthesis of brookite TiO2, scheme for such a synthesis is shown in Figure 7A. In another
Cozzoli et al.57 reported that brookite is formed via a direct study, the sub-nanometer-sized TiO2 particles have been
solid-state phase transformation of the initially formed c-axis- fabricated by using the phenylazomethine dendrimer (DPA
elongated anatase, accompanying homogeneous nucleation and G4) template.65 The quantum-confinement effect observed in
heterogeneous nucleation processes. As one of the most the subnanometer TiO2 particles is shown in Figure 7B. The
intriguing forms of titanium oxide, brookite TiO2 shows energy gap exhibits a blue-shift with decreasing particle size.
remarkable photocatalytic activity.58 For future technological Combined with crystal modeling, the smallest anatase or rutile
uses of brookite, it is also interesting and practical to explore cluster would be 3 × TiO2. It was revealed that crystalline form,
the dielectric performance of high-quality brookite TiO2.59 In rutile or anatase, exists in small TiO2 clusters, which has an
influence on the optical band gap.65
addition, brookite TiO2 can provide higher volumetric energy
4.1.2. Low-Temperature Synthesis of Sub-5-nm TiO2
density than comparable nanomaterials in lithium-ion bat-
Nanoparticles. For a typical TiO2 nanoparticle (NP)
teries.29
synthesis, high-temperature treatments during either the
hydro/solvo-thermal preparation or the postsynthesis annealing
4. SYNTHESIS AND CHARACTERIZATION OF TIO2 NPS are required in order to achieve crystalline TiO2 NPs. The TiO2
nanoparticle sizes are usually larger than 10 nm with these
4.1. Novel Aspects to the Synthesis Methods of TiO2 NPs
approaches with high-temperature treatments because these
In 2007, Chen60 reviewed a variety of synthetic techniques of regular annealing temperatures merge the initially small
TiO2 nanomaterials, including sol−gel, micelle, hydrothermal, nanoparticles into larger ones during the high-temperature
solvothermal, sonochemical, microwave, electrodeposition, treatment. The low-temperature preparation method therefore
physical vapor deposition, and chemical vapor deposition is of great interest for preparing small nanocrystalline
methods. The crystal sizes and crystal structures of TiO2 TiO2.66−73
strongly depend on the synthesis parameters including the TiO2 NPs with sizes around 2−10 nm can be synthesized
temperature, solvent, additives, acidity, and aging during the using a modified sol−gel method at around 50−100 °C.68−73
9287 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Nanoscale anatase TiO2 with high crystallinity has been hollow spherical TiO2 NPs and core−shell TiO2 NPs can be
successfully synthesized at a temperature as low as 100 °C by fabricated by surface modification and adjustment of the
a novel simple method using anhydrous diethyl ether as the deposition parameters.87
solvent for the hydrolysis of tetrabutyl titanate.74 This synthesis 4.1.3. Nonaqueous Preparation of TiO2 NPs. The high
showed a huge advantage over the other methods which require refractive index, strong UV absorption, and low cost make TiO2
other additives, template agents, or special equipment. High- NPs candidates to modify polymers by forming TiO2−polymer
crystallinity anatase TiO2 NPs with a size range 2−4 nm have nanocomposites. The dry TiO2 NPs powders are difficult to be
been prepared at low temperature using ethylene glycol to homogeneously redispersed in nonaqueous solutions, while
control the hydrolysis and condensation rates.75 The simplicity most polymers are usually dispersed and prepared in non-
(only three reagents are involved) and reproducibility of this aqueous media. Currently, most TiO2 NPs are synthesized from
method makes this route possibly scalable. Highly soluble aqueous methods by hydrolysis of the Ti precursor. All these
anatase nanoparticles with average diameter of 3 nm and facts make it technically difficult to prepare highly homoge-
excellent dispersity have been synthesized from TiCl4 at 60 °C neous TiO2−polymer nanocomposites. To overcome these
using tert-butyl alcohol as a new sol−gel reaction medium problems, nonaqueous synthesis schemes for TiO2 NPs have
without the need of additional stabilizing ligands.76 More sol− been developed and have recently gained additional attrac-
gel media such as ethanol, n-butanol, and hexanol have been tion.76,88−95
adapted to the synthesis of 3−5 nm sized TiO2 nanocrystals In aqueous synthesis, the Ti precursors react with H2O
(NCs) at reaction temperatures of 90−170 °C.77 The sub-5-nm through the hydrolysis-and-condensation reaction to form a
TiO2 NC suspensions are nearly transparent due to the small Ti−O−Ti network with a fast reaction rate. In nonaqueous
particle sizes, as shown in Figure 8, which is typical for small syntheses of TiO2 NPs, the Ti precursor reacts with alcohols at
size nanoparticle solutions. higher temperatures through an aldol condensation to form the
Ti−O−Ti organo-gel.92 tert-Butyl alcohol and benzene alcohol
with high boiling points are often utilized in the nonaqueous
sol−gel synthesis. Since the nonaqueous synthesis reaction
rates are much slower than corresponding aqueous synthesis,
3−5 nm monodisperse TiO2 NPs have been successfully
synthesized.76 As shown in Figure 9A, nonaqueous prepared

Figure 8. Photographs A77 and B78 of sub-5-nm TiO2 NPs suspension


prepared by low-temperature synthesis. Reproduced with permission
from refs 77 and 78. Copyright 2013 Elsevier.

Besides the use of solvents to control the hydrolysis of Ti


precursors, nitric acid has been adopted to control the
hydrolysis of titanium butoxide and the synthesis of 2−5 nm
anatase TiO2 NPs by heating to 80 °C for 20 h.78 Rutile TiO2
nanorods were synthesized through a modified sol−gel method
by refluxing at temperatures of 50−90 °C with the addition of
HCl acid catalyst to inhibit the hydrolysis and promote the Figure 9. (A) Picture of nonaqueous synthesized TiO2 NP solutions:
crystallization process.79 Fine nanocrystalline TiO2 on a large in tert-butanol/toluene (3:1 volume ratio, 3.3 wt % TiO2, left), after
flocculation with heptanes centrifugation, and redispersion in ethanol
scale could be gained by controlling the hydrolysis in
(3.3 wt %, middle), and further diluted in ethanol (0.7 wt %, right).76
supercritical CO280 and ionic liquid [Emim]Br(1-ethyl-3- (B) Photograph of 1.1 wt % TiO2 in PMMA film.95 (C) Proposed
methyl-imidazolium bromide).81 crystallization pathway for TiO2 NPs prepared from titanium
Other studies show that the low-temperature crystallization isopropoxide and benzyl alcohol.95 Fast nucleation yields small
of amorphous TiO2 precursors to nanostructured anatase could particles and a high particle concentration. Slow nucleation and a
be initiated with a slightly reducing chemical environment such low concentration of seeds consuming the same amount of monomer
as glucose at a low temperature of around 80 °C.82 In the results in large particles. Reproduced with permission from refs 76 and
presence of various amino acids, crystalline titanium dioxide 95. Copyright 2010 and 2009 American Chemical Society.
particles were obtained through hydrolysis of an aqueous TiCl4
precursor at 60 °C.83 Amino acids, the protein building units, TiO2 NPs can be well-dispersed in nonaqueous solvents, so that
can be considered as an interesting first step in bioinspired they can be processed with polymers to form homogeneous
approaches to synthesize nanoscale TiO2 with narrow size TiO2−polymer nanocomposites. From Figure 9B, it can be
distribution. Although it is an emerging nanofabrication seen that a 1.1% TiO2 poly(methyl methacrylate) film prepared
technology, in the controlled synthesis of nanoscale TiO2, via nonaqueous synthesis is transparent. This nanocomposite
biomolecules such as peptides,84 silaffin protein,85,86 and film has strong UV absorption after embedding TiO2 NPs into
apoferritin87 were proven to be very efficient in aqueous the poly(methyl methacrylate). In addition, the amorphous
solutions at ambient temperatures (below 100 °C). Employing TiO2 NPs formed in the nonaqueous synthesis need to be
apoferritin, a spherically shaped protein cage, as template, crystallized in the nonaqueous solvent with a high-temperature
9288 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 10. TEM images of TiO2 NCs synthesized using the precursor TiF4 (a, d), a mixed precursor of TiF4 and TiCl4 (b, e), and TiCl4 (c, f).105
Those depicted in a−c and d−f are synthesized in the presence of oleylamine and 1-octadecanol, respectively. Reproduced with permission from ref
105. Copyright 2012 American Chemical Society.

treatment (Figure 9C). Figure 9C also illustrates that the final ions as morphology controlling agents. Chainlike 1-D TiO2
particle size is dependent on the rates of nucleation and growth. nanostructures built from nanobipyramids were generated via
4.1.4. Controlling Shapes and Crystal Facets of TiO2. It the mechanism of oriented attachment at higher temper-
was demonstrated that TiO2 NCs with various shapes, such as atures.106 Also, the percentage of reactive (001) facets of TiO2
spherical, rhombic, truncated and elongated rhombic, dog- NCs can be tuned by the amount of water present in rapid
bone, bar, and dots, can be tailored by using a simple microwave-assisted hydrothermal synthesis. With an increasing
solvothermal route employing both oleic acid and oleic amine amount of water in the synthesis, the shapes of TiO2 NCs
as capping agents in the presence of water vapor.96 Crystalline change from nanosheets to truncated octahedral bipyramids.107
TiO2 with different morphologies/shapes exposes different Yang et al.108reported a new solvothermal method using 2-
crystal facets and exhibits different surface properties.97 propanol as a synergistic capping agent and reaction medium
The structure of the rutile TiO2 (110) surface has been well- together with HF to synthesize high-quality anatase TiO2
characterized, as it is the most commonly investigated titania single-crystal nanosheets with 64% of the (001) facets.
crystal facet.31,98,99 The (011) surface of rutile TiO2 shows a Importantly, the percentage of highly reactive (001) facets in
higher reactivity in catalytic reactions than other surfaces, which rectangular TiO2 nanosheets is very high (up to 89%) with
was determined by a combination of noncontact atomic force optimal adjustment of the amount of hydrofluoric acid and
microscopy, scanning tunneling microscopy, and density reaction temperature during the hydrothermal synthesis.109
functional calculations.100 Anatase TiO2 crystals are usually Such TiO2 nanosheets show excellent photocatalytic efficiency,
dominated by (101) facets, which are thermodynamically far exceeding that of commercially available Degussa P25.
stable.101 It has been demonstrated that the order of the Unlike anatase and rutile, it is relatively difficult to prepare
average surface energies of anatase TiO2 is 0.90 J/m2 for (001) high-quality brookite TiO2 NCs. Nevertheless, single phase
> 0.53 J/m2 for (100) > 0.44 J/m2 for (101). Due to their high brookite TiO2 NCs were obtained by a one-step hydrothermal
surface-free energy, (001) facets of anatase are generally treatment of an aqueous titanium complex solution.110 Also,
considered to be more reactive than (101) facets. However, high-quality anisotropically shaped brookite TiO2 NCs can be
(001) facets usually diminish rapidly during a crystal growth synthesized using a surfactant-assisted nonaqueous strategy.57
process. Recently, an important breakthrough in the prepara- Compared with studies on controlling shapes and crystal facets
tion of anatase TiO2 crystals with exposed (001) facets was of anatase and rutile TiO2, very few groups have addressed the
reported by Lu and co-workers.102 They demonstrated that the crystal facets of brookite TiO2. On the basis of first-principle
(001) facets can be stabilized by the use of hydrofluoric acid as studies, the (001) facet of brookite TiO2 holds the smallest
a shape-controlling agent. In Xia’s report,103 TiO2 NCs with surface formation energy of 0.62 eV, while (100) exhibits a
truncated tetragonal bipyramidal shape and 9.6% of the surface larger value of 0.88 eV. This calculation reveals that brookite
being exposed by (001) facets were synthesized in high yields crystals prefer growth along the [001] direction to produce
by controlling the hydrolysis rate of the sol−gel precursor and nonspherical morphology.111 Brookite TiO2 with many kinds of
hydrothermal treatment. Low pH values tend to eliminate the morphologies, such as nanorods,52 nanofibers,55 nanotubes,53
(001) facets by forming sharp corners while high pH values and nanoflowers,59,112 have been reported in the literature. For
favor the formation of a rodlike morphology through an the formation of TiO2 nanoflowers, brookite particles grow to
oriented attachment mechanism. Changing the relative ratio of spindle-like shapes, due to the small surface energy, and then
titanium precursor and HF during hydrothermal synthesis can these spindle-like particles can assemble into flower-like
lead to different degrees of truncation.104 Highly uniform morphology.59 The preparation conditions of brookite TiO2
anatase TiO2 NCs with tailorable morphology, preferentially including inorganic salts, organic substances, pH value, reaction
exposing the (001) facet, can be prepared through a seeded time, and temperature were investigated.55,113,114 Among them,
growth technique by using titanium(IV) fluoride (TiF4), which sodium ions (Na+) play a key role in the formation of brookite
can in situ release hydrofluoric acid (HF) during reaction TiO2, which may promote the brookite nucleation. Recently, it
(Figure 10).105 Highly crystalline TiO2 truncated tetragonal was reported that anions with different molecular structures are
bipyramidal nanocrystals enclosed by (001) and (101) facets used as absorbents to control the growth direction of TiO2 as
were fabricated via a microemulsion method employing fluorine shown in Figure 11, resulting in rutile and brookite TiO2 NCs
9289 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

crystal grain size according to the Scherrer equation. Crystallite


size is determined by measuring the broadening of a particular
peak in a diffraction pattern associated with a particular planar
reflection within the crystal unit cells. It is inversely related to
the full width at half-maximum (fwhm) of an individual peak.
The narrower the peak is, the larger the crystallite size will be.
However, crystallites of less than 3−4 nm cannot be
determined reliably due to the detection limit of XRD. This
limitation can be solved by means of Raman spectrosco-
py.117,118 Similarly, as the size of TiO2 nanomaterials decreases,
the featured Raman scattering peaks become broader (Figure
13).119
Figure 11. Scheme of anion-assisted crystal form and crystal facet In general, the XRD peaks at 25.3°, 14.2°, and 27.4° are
control of TiO2 NCs.115 Reproduced with permission from ref 115. identified as the characteristic diffraction peaks for anatase,
Copyright 2013 American Chemical Society. brookite, and rutile TiO2, respectively. From comparing the
irreducible representation of the light scattering modes with the
crystal phase symmetry, the three phases of anatase, brookite,
as well as anatase TiO2 NCs with different facets (101), (001), and rutile have 6(3Eg+ 2B1g + A1g), 36(9A1g + 9B1g + 9B2g +
and (100). From Figure 12, it can be seen that the reduction 9B3g), and 4(A1g + B1g + B2g + Eg) Raman active modes,
ability of different anatase crystal facets can be ranked as (101) respectively.120 Brookite, either natural or synthetic, shows
> (001) > (100), while the oxidation ability of different facets strong Raman peaks at 128 (A1g), 153 (A1g), 247 (A1g), 322
can be ranked as (101) ≈ (001) ≈ (100). That is, the results (B1g), 366 (B2g), and 636 (A1g) cm−1. Anatase exhibits
based on specific reactions should be analyzed when discussing characteristic Raman scattering at 146 (Eg), 396(B1g), 515(A1g),
the crystal-facet-dependent catalytic activities of TiO2 NCs.115 and 641 (Eg) cm−1, while rutile shows typical scattering at
4.2. Characterization Approaches 143(Eg), 235 (two-phonon scattering), 447(Eg), and 612 (A1g)
For nearly all studies on TiO2 NPs as functional building cm−1. Besides the extensive use in phase identification of TiO2,
blocks, electron microscopic techniques, transmission electron Raman spectroscopy serves also as an efficient tool for probing
microscopy (TEM), and scanning electron microscopy (SEM) oxygen deficiency of the TiO2 lattice. It is widely observed that
are the first and foremost characterization approaches to increased contents of oxygen vacancies in the crystal structure
identify the morphology of TiO2 nanostructures and the grain lead to higher wavenumbers of the anatase Eg mode (146 cm−1)
size of TiO2 NPs. As shown in section 4.1, high-resolution while the wavenumbers of the rutile Eg mode (447 cm−1) are
TEM (HRTEM) is one of the most powerful and versatile lowered by oxygen vacancies.121
techniques to identify the crystal facets and shapes of TiO2 4.2.2. Optoelectronic Properties via UV-DRS, XPS,
NCs. With the assistance of energy-dispersive X-ray spectros- EPR, IR, and PL Spectroscopy. Fujishima98 reviewed the
copy (EDXS) and electron energy-loss spectroscopy (EELS) electronic properties of TiO2 and recognized that the creation
complementary techniques, it allows one to reach atomic of Ti3+ sites is responsible for the electronic conductivity. No
resolution of crystal lattices and to obtain chemical and evidence for ionic conduction is found in TiO2. Convincing
electronic information at the subnanometer scale.116 While evidence has revealed that the source of electronic conductivity
TEM has made enormous contributions to nanoscience, in rutile is Ti3+ that is associated with oxygen vacancies Ov.98
section 4.2 is focused on the characterization approaches to The activation energy for the electronic conductivity is found to
structural properties, optoelectronic properties, as well as be 1.75 eV for unsintered rutile powder and 1.7 eV for sintered
charge carrier dynamics in TiO2 nanostructures. rutile powder. There are large differences in the electronic
4.2.1. Structural Properties via XRD and Raman conductivities of rutile versus anatase thin films after reduction
Spectroscopy. XRD is essential in the determination of the by heating in vacuum, which is considered to be due to the
crystal structure and the crystallinity, and in estimating the different dielectric coefficient and effective electron mass.122

Figure 12. Photocatalytic performances of TiO2 NCs with different crystal forms and crystal facets in (A) reduction of nitrobenzene to aniline after
reaction for 30 min and (B) oxidation of benzyl alcohol to benzaldehyde after reaction for 4 h.115 Rutile TiO2 nanorods (TiO2-R), octahedral anatase
TiO2 NCs (TiO2-A-O), truncated octahedral TiO2 NCs (TiO2-A-TO), TiO2 nanosheets (TiO2-A-NS), long-rod shaped anatase TiO2 NCs (TiO2-A-
LR), short-rod shaped TiO2 NCs (TiO2-A-SR). Reproduced with permission from ref 115. Copyright 2013 American Chemical Society.

9290 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

Figure 13. XRD (A) and Raman spectra(B) of TiO2 NPs at various sizes.119 Reproduced with permission from ref 119. Copyright 2012 American
Chemical Society.

Figure 14. (a) UV−vis diffusion reflectance spectra and (b) energy dependence of (F(R)*hν)n for brookite TiO2 flowers. Corresponding data for
rutile and anatase TiO2 are also shown for comparison.59 Reproduced with permission from ref 59. Copyright 2009 American Chemical Society.

Figure 15. (A) Scheme of UV-induced charge separation in TiO2. Electrons from the valence band can either be trapped (a) by defect states, which
are located close to the conduction band (shallow traps), or (b) in the conduction band where they produce absorption in the IR range.135 (B) EPR
spectra recorded when nanorutile is irradiated in vacuo at 4 K with broadband radiation. (a) Initial spectrum in the dark (the signal at 325 mT is due
to a defect in the quartz sample tube), (b) during irradiation, (c) 15 min after switching off the light, and (d) computer simulation of the Ti3+
signal.133 Reproduced with permission from refs 135 and 133. Copyright 2005 and 2012 Elsevier.

The electronic structures of TiO2 have been theoretically properties. The high-quality brookite flowers presented by Hu
investigated using DFT methods.123,124 But the UV−vis and his co-workers59 show a direct transition with a band gap
spectrum is considered as the most reliable technique to energy of 3.4 ± 0.1 eV, which is larger than those of its two
measure the band gaps of TiO2 NPs. In the UV−vis spectrum other polymorphs, that is, a direct band gap of 3.0 ± 0.1 eV for
of TiO2, the absorption peaks at 220−260 and 330−400 nm rutile and an indirect band gap of 3.2 ± 0.1 eV for anatase
correspond to the tetrahedral and octahedral coordinate Ti (Figure 14). UV−vis spectroscopy of oxygen-deficient TiO2
species, respectively.125 It is beneficial to investigate the change systems demonstrates the existence of prominent absorption
in coordination number during the growth of a crystal. The bands in the visible region.98 On the basis of recently
octahedral metal centers are the basic structure unit of both demonstrated experimental observations, it is deduced that
anatase and rutile. As is well-known, the optical properties of the spectral features of visible-light-active TiO2 photocatalysts
TiO2 depend strongly on the type of material (e.g., single originate from F-type color centers associated with oxygen
crystal, powder) and the synthesis conditions (e.g., calcination vacancies and Ti-related color centers.128
atmosphere).126 Zallen127 assigned the natural brookite crystal X-ray photoelectron spectroscopy (XPS) is a surface-
as an indirect-gap semiconductor by analyzing its optical sensitive spectroscopy to measure binding energies for the
9291 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 16. (A) Model for trap state photoluminescence in anatase (left) and rutile (right). Wavy and straight lines indicate nonradiative and radiative
transitions, respectively. The PL of anatase is considered to be a combination of both type 1 and type 2 PL involving spatially separated hole and
electron traps, respectively, while the rutile PL at ∼840 nm is type 1.139(B) PL development for TiO2 during irradiation in vacuum. The baseline
spectrum, obtained using a clean tungsten grid at the sample position, is shown by the blue line.140 Reproduced with permission from refs 139 and
140. Copyright 2008 and 2012 American Chemical Society.

individual participating elements. The photoelectrons ejected light-emitting properties of the TiO2 allotropes upon UV or X-
from Ti ions with Al Kα irradiation have a mean free path of 1− ray excitation (Figure 16A). The total photoluminescence (PL)
2 nm, which means that one probes surface atoms on the of anatase involves spatially separated trapped electrons and
nanoparticle and about 2−3 atomic layers below the surface. trapped holes, which are about 0.7−1.6 and 1.8−2.5 eV below
This technique can identify specific oxidation states of the the conduction band edge, respectively. The commonly
atoms and determine their chemical environments. The valence observed luminescence from anatase (nanocrystals, in most
band (VB) XPS spectra for TiO2 related peaks near 22−24 eV cases) is in the visible green region (∼550 nm), and the
are generally associated with the O 2s region. The region luminescence is highly sensitive to the surface properties.
between 3 and 9 eV, attributed to the O 2p orbitals in pure Rutile, on the other hand, has no visible luminescence but
TiO2, is very sensitive to the Ti−Ti and Ti−O distances.129 shows a near-IR (NIR) emission (∼800 nm) and is less surface
The surface species and charge-transfer process of TiO2 and sensitive. Knorr et al.138 found that TiO2 nanocrystalline films
modified TiO2 NPs can be identified by the combination of the containing a small amount of rutile show solvent-dependent
core level and the VB XPS data.130−132 relative PL intensities of the anatase and rutile that reveal
Electron paramagnetic resonance (EPR) and infrared carrier transport between the two phases. Ellis et al.139
spectroscopy (IR) are powerful techniques to explore the suggested that the variation in PL intensity by the adsorption
electronic structure of excited TiO2 NPs.133,134 Localized of charge-donating molecules on an n-type semiconductor
carriers such as holes trapped at oxygen anions (O−) and could alter the surface band structure of the semiconductor by
electrons trapped at coordinatively unsaturated cations (Ti3+ reducing the depletion width as donor molecules are adsorbed.
formation) are accessible to EPR spectroscopy (Figure 15). Yates et al.140 found that the PL intensity of TiO2 at 529.5 nm
During continuous UV irradiation, photogenerated electrons in (2.34 eV) increased with irradiation time as the number of
TiO2 NPs get trapped at localized sites, forming paramagnetic photon-accessible defects increased (Figure 16B). They further
Ti3+ centers.133 Similarly, EPR-detectable holes also form upon concluded that alteration of the surface potential of TiO2 by
photogeneration of active O− anions in lattice O2− dianions.135 UV light or adsorbed electron-donor/acceptor molecules
In contrast, delocalized electrons in the conduction band are results in a change in the depth of the active PL region and
EPR silent but can be traced by their IR absorption. Panayotov in the intensity of the observed PL as a result of the band-
et al.136 have demonstrated that IR radiation can excite the bending effect.140
fraction of electrons that are trapped at shallow donor levels 4.2.3. Laser Spectroscopic Characterization of Charge
0.12−0.3 eV below the conduction band minimum. The Carrier Dynamics. Time-resolved laser spectroscopic meas-
trapped electrons have a broad IR absorption with a maximum urements provide a direct and versatile approach for probing
characteristic of the donor level energy. The free conduction the electronic dynamics of nanoparticle-based systems and lead
band electrons exhibit a broad featureless absorbance that to a better understanding of the rates of carrier injection and
increases exponentially across the entire mid-IR range.136 competing relaxation pathways at the nanoparticle interface.
Since anatase and brookite TiO2 are both indirect-type The optical properties of nanomaterials can readily be studied
semiconductors, band gap emission from recombination of on time scales of tens of femtoseconds and longer. This allows
conduction band electrons with valence band holes is extremely the monitoring of the electron and hole dynamic processes in
weak at room temperature. However, localized defect states real time as they occur.
have been reported to exhibit radiative recombination of More than 30 years ago, laser-induced photoelectrochemical
trapped electrons and holes.137 The energy of the emitted effects at the TiO2 semiconductor−electrolyte interface could
photon depends thereby on the band structure and defect-site be measured by time-resolved flash techniques.141 Currently,
energies in the material. The difference in the crystal structures transient absorption (TA) spectroscopy has been used
between anatase and rutile leads to interesting differences in extensively to investigate the dynamics and mechanisms of
9292 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 17. (A) Transient absorption bands for trapped holes, trapped electrons, and bulk electrons in TiO2 nanocrystalline film.147 (B) Diagram of a
time-resolved IR absorption spectrometer for kinetic measurements of photocatalysts.142 (C) Scheme to explain the trapping and relaxation
dynamics of electrons and holes in TiO2 excited at 266 and 355 nm.147 (D) Primary reaction steps in DSSCs.143 Reproduced with permission from
refs 142, 143, and 147. Copyright 2009 and 2004 American Chemical Society. Copyright 2003 Elsevier.

photoinduced charge transfer occurring in TiO 2 NPs conduction-band electrons within the lattice and within pico- to
structures,142,143 involving adsorption dynamics of molecules nanoseconds, a deep trap level filling process with time
on the TiO2 surface, the driving force for the interfacial constants of up to microseconds, and much slower decay
electron-transfer reactions, and the electronic interaction processes corresponding to the interparticulate carrier transport
between TiO 2 and adsorbates.136,144−146 The transient and deep trapped electron−hole recombinations at time scales
absorptions of nanocrystalline TiO2 are assigned to three less than a microsecond.149 At the dye−TiO2 nanoparticle
kinds of charge carriers: trapped holes absorbing at 500 nm, interfaces, forward electron transfer occurs within femtoseconds
trapped electrons absorbing at 800 nm, and bulk electrons with to hundreds of picoseconds. The often observed inhomoge-
an increasing absorption in the IR region toward longer neous electron-transfer rates can be interpreted as the
wavelengths (Figure 17A).147 The absorption of injected interaction between an ensemble of dye molecules with a
electrons in TiO2-based DSSCs appears in the IR wavelength diverse range of TiO2 NP environments.150 The benefit of
range.144 So, sensitive absorption spectrometers with wide transient kinetics measurements is the opportunity to obtain
spectral ranges (∼400−3000 nm) are most useful for such absolute or relative efficiencies of photoinduced processes, such
studies (Figure 17B).142 For femtosecond (fs) time-resolved as the electron injection in DSSCs (Figure 17D), which can be
TA spectroscopy, the observed dynamics depend on many evaluated from the quantitative analysis of transient lifetimes. It
factors, including some instrumental ones such as the pump is found that, in the gold−TiO2 NP system, the electron
light properties. Usually Ti-sapphire lasers in combination with injection is completed within 50 fs and the electron injection
optical parametric amplifiers are in use. Pulse intensity and yield reaches 20−50%. The charge recombination decay within
wavelength, even the repetition rate of the excitation laser pulse 1.5 ns is nonexponential and is strongly dependent on the TiO2
train, have a direct effect on the observed relaxation dynamics. particle diameter.151
The detector materials (Si, InGaAs, or MCT photodetector) On the basis of TA spectroscopy studies, transient absorption
are often chosen according to the desired observation anisotropy measurements are developed and used for
wavelength range. As an example, it has been reported that mechanistic characterization of photoinduced reactions at
the rate of the charge recombination in nanocrystalline TiO2 nanostructured TiO2 surfaces.152 An effective diffusion constant
films is very sensitive to laser excitation intensity Iex.148 In weak for self-exchange hole transfer is quantified on a minutes time
355 nm excitation, the generated charge carrier density is low scale, while the anisotropy itself decays on a micro- to
enough that the second-order electron−hole recombination millisecond time scale under most experimental conditions.152
processes could be ignored.147 Photoexcited holes are trapped Thus, transient absorption anisotropy provides the first direct
within 100 fs at sites near the surface of the TiO2 NPs. evidence for lateral self-exchange hole transfer across semi-
Electrons are first trapped at shallow sites near the surface, conductor nanocrystallites after excited-state injection.
equilibrated by migrating over a nanoparticle, and then relax Similar to TA spectroscopy derived from transmission
into deeper trapping sites in the bulk with a common time experiments, femtosecond time-resolved diffuse reflectance
constant of ∼500 ps (Figure 17C).147 The photoinduced (TRDR) spectroscopy and time-resolved microwave conduc-
carrier relaxation for nanocrystalline TiO2 films consists tivity (TRMC) were employed under weak excitation
typically of three kinetic phases, that is, a rapid decay for conditions to clarify the charge separation and trapping
9293 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 18. Basis of kinetic models used to fit time-dependent photoluminescence decay data. (A) Simple exponential model with a unique and well-
defined barrier height ΔG⧧. (B) Albery model assumes a Gaussian distribution of barrier heights. (C) KWW model is comparable to the Albery
model but with an asymmetric distribution. (D) Power-law model has a most-barrier probability that falls off exponentially with increasing
energy.170Reproduced with permission from ref 170. Copyright 2012 American Chemical Society.

dynamics in TiO2 NPs systems. TRDR spectroscopy has from the UV-irradiated TiO2 surface to the gas phase was
advantages of measuring turbid or powdery samples with strong successfully detected.172 It is found that the diffusion time of
scattering or bad light transmission. This technique has been OH• radicals varies with the type of TiO2 powder and the heat
utilized to investigate the charge carrier dynamics of differently treatments of these powders.172
doped TiO2 NPs and their photocatalysis.153−156 In TRMC, the In summary, in both photocatalytic and photoelectrochem-
mobility of photogenerated charge carriers can be probed by ical reaction systems based on TiO2 NPs, controlling the charge
the interaction of the mobile charge carriers with microwave carrier dynamics, including hot carrier relaxation, trapping,
radiation. This technique investigates processes of carrier interfacial carrier transfer, and recombination, is essential for
trapping and recombination, exciton annihilation, and quench- successful energy conversion. Femtosecond laser spectroscopy
ing in TiO2 NPs. TRMC is particularly useful for studying the can provide direct insight into charge carrier dynamics at the
properties of TiO2-based solar cells and in dynamics related to nanoscale and provide a basis upon which to fine-tune these
photoelectrochemistry.157−167 optoelectronic systems with a temporal resolution of 100 fs
Time-resolved photoluminescence (TRPL) spectroscopy, (10−13 s) and better.168 Thus, femtosecond time-resolved laser
another popular femtosecond laser-based technique, can also spectroscopy is key and will pave the way to optimize the
provide direct insight into the charge carrier dynamics of energy conversion of novel and complex functional TiO2
nanomaterials. TA and PL techniques are used to probe nanoarchitectures that are being developed at this point.168
different kinetic pathways in semiconductor NCs. It is
necessary to observe the emitting signal in PL measurements. 5. TIO2 NPS AS GROWTH CENTERS: EVOLUTION OF
TA pump−probe spectroscopy only measures a low fraction of TIO2-BASED NANOMATERIALS
a few percent PL, while most of the TA signal is from
The syntheses of TiO2 NPs have been broadly studied over
nonradiative pathways, measured as excited state transient
several decades reaching a remarkable degree of sophistication.
absorption.168 TRPL is often used to clarify the dynamics of
With the development of nanomaterials chemistry, controlling
interfacial charge-transfer emission in photosensitized TiO2
the growth of the nanoparticle became feasible. The strategy of
NPs. For example, time-resolved fluorescence and fluorescence
using TiO2 NCs as growth centers became popular in
anisotropy of molecule−TiO2 nanostructures can be charac-
fabricating more complex TiO2-based nanostructures. Many
terized to describe the nature of the charge-transfer excitation
intricate nanostructured materials with fascinating properties
using a femtosecond fluorescence upconversion setup.169 It is
have inspired extensive research on the synthesis of materials
measured that interfacial charge-transfer emission lifetimes
with engineered structures.173−176 The recent realization that
ranged from <100 fs to a few hundreds of femtoseconds for nanoscale materials can be synthesized with great structural
small sensitizer molecules like catechol, dopamine, acetyl diversity provides numerous possibilities for novel materials
acetonate, salicylate, and hydroxamic acid depending on the with engineered designs and properties.177
electronic coupling of the molecule to the TiO2 nanoparticle.169
To fit the TRPL data in such systems, there are four kinetic 5.1. One-Dimensional (1-D) Growth
models discussed, also applicable to dye-sensitized TiO2 films: a One-dimensional nanostructures, such as nanowires,178 nano-
three-component exponential fit (homogeneous kinetics rods,179,180 nanofibers,181−183 nanotubes,184 and nanorib-
model), the Albery model, the Kohlrausch−Williams−Watts bons,185 show high electron mobility, excellent electron hole
(KWW) model, and power-law model (Figure 18). McNeil et separation power, and long-distance transport capability due to
al.170 found that the power-law model, in which the distribution their nanosize confinement in two dimensions while the third
of barriers falls off exponentially, provides the best fit to the dimension can be nano-, micro-, or even macroscaled.
measured PL data. They concluded that the time-dependent Therefore, 1-D structures can offer numerous possibilities for
emission decay follows a power-law profile. It is also proposed applications that are expected to be important for future
that each dye molecule injects electrons only into a small devices.
number of localized states that have well-defined energies, On the basis of peroxo−titanium complex decomposition,
rather than injection into abroad range of states. TRPL spectra the rutile TiO2 with nanorod morphology can be obtained
can also quantitatively analyze the electron-transfer event in according to the oriented attachment nanocrystal growth
TiO2 nanoheterojunctions. The PL decay data were analyzed mechanism.186 The growth process involves spontaneous self-
and fitted with an appropriate exponential model in which the organization of adjacent nanocrystals and coalescence. With
decay component (τi), the pre-exponential factor (Ai), and the oleic acid as a surfactant, TiO2 nanorods composed of length
intensity-averaged lifetime (⟨τ⟩) are derived.171 Using the laser- and diameter of approximately 35 and 4 nm, respectively, were
induced-fluorescence technique, the diffusion of OH• radicals synthesized at 100 °C.187 Interestingly, only a small portion of
9294 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 19. (A) SEM image of the as-spun TiO2-poly(vinylpyrrolidone) composite nanofibers. (B) Bright-field micrograph of nanofibers heated at
650 °C containing both anatase and rutile phases. (C) HRTEM image of fibers heated at 650 °C showing grain growth.188 Reproduced with
permission from ref 188. Copyright 2007 American Chemical Society.

Figure 20. HRTEM images of as-grown titania nanotube powder synthesized in HClO4 electrolyte (A, B) and annealed at 480 °C for 3 h in air
atmosphere (C). Parts A and B show the nanotubular structure of the resulting TiO2 nanotube powder with open mouth and closed bottom, as well
as the tube wall; part C shows the TiO2 nanotube powder with crystallized walls.191 Reproduced with permission from ref 191. Copyright 2009
American Chemical Society.

Figure 21. Illustration of the formation of TiO2 hierarchical nanostructures from a single nanoparticle and SEM image of resulting structure.196
Reproduced with permission from ref 196. Copyright 2010 American Chemical Society.

phase transformation occurred after annealing at 850 °C for 2 a template-directed method,184 and an electrochemical rapid
and 3 h. These slow phase transformations were ascribed to the breakdown anodization method.191 On the basis of recent
lowered surface free energy of the anatase phase due to the theoretical analysis,192,193 the nanostructure transformation of
small diameter of the TiO2 nanorods. Nanofibers of TiO2 are TiO2 between nanotubes, nanorods, and nanoparticles can be
often constructed by the electrospinning technique. The performed by changing the pH of the chemical solution during
anatase TiO2 nanofibers of variable diameters can be fabricated the hydrothermal synthesis.190,194,195 The TEM images in
by controlled electrospinning of a polymeric solution and Figure 20 show the nanotubular structure of the resulting TiO2
subsequent sintering.188 Figure 19 shows that the sintered nanotube powder with open mouth and closed bottom and also
fibers are polycrystalline and compose of densely packed TiO2 clearly indicate crystallization of the tube walls with an average
grains. Thus, the TiO2 nanofibers can be formed of crystallite size of ∼13 nm.191 A comprehensive DFT-based first-
nanocrystals within a rigid and robust mesoporous frame- principles calculation was performed to study the geometric,
work.188 energetic, and electronic properties of the rutile TiO2
The good-quality TiO2 nanotube powders were prepared nanowires and nanotubes. It is found that the competition
through a hydrothermal process from anatase TiO2 NPs,189,190 between the surface-bonding effect and the quantum-confine-
9295 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 22. (A) TEM images of the multineedle TiO2 particles.198 (B) SEM images of a typical individual nanotree grown from and rooted on
titanium-sputtered substrate.199 (C) Cross-sectional SEM view for TiO2 nanocoral structures.200 Reproduced with permission from refs 198,199, and
200. Copyright 2009, 2011, and 2012 American Chemical Society.

ment effect induces the direct−indirect transition of the band highly reactive (001) facets, two-dimensional (2-D) TiO2
gap of TiO2 nanowires and nanotubes.144 nanosheets oriented along the (001) direction become the
5.2. TiO2 with Higher Organized Architectures focus of the research.210 Jaroniec et al.211 reviewed the existing
strategies for the synthesis of anatase TiO2 micro- and
Starting from the 0-D nanoparticle, the hierarchical growth nanosheets with exposed high-energy (001) facets and for the
patterns can be further expanded. Hierarchical crystalline assembly of these nanosheets into various hierarchical
anatase TiO2 nanostructures have attracted great interest due structures. The fabrication of perpendicular TiO2 nanosheet
to their novel architecture and their resulting fascinating films (to the substrate) is of great interest, because it is
properties. Figure 21 illustrates the formation of hierarchical suggested that vertical large-area flat planes of the TiO2
TiO2 nanostructures during the hydrothermal process starting nanosheets are thermodynamically stable.212 In addition, the
from a central seed particle. 196 The 3-D hierarchical edge planes of the TiO2 nanosheets contain a large number of
morphologies of the nanostructures originate from the growth defects and dangling bonds as a result of their thermodynamic
of dozens of radially distributed thin sheets, each of several instability. This can make them display superhydrophilicity
nanometer thickness which leads to optimized large surface without UV irradiation.
areas.196,197 Similarly, flower-like multineedle TiO2 nanostruc-
tures were successfully fabricated by mild aqueous chemistry 5.3. Breaking One-Dimensional TiO2 Arrays into
Zero-Dimensional TiO2 NPs
(Figure 22 A).198
In addition, a general acid vapor oxidation strategy has been Nanostructured photoelectrodes for photoelectrochemical cells
developed to grow highly oriented hierarchically structured are among the most important applications of TiO2. Especially
rutile TiO2 nanotrees with tunable morphologies from titanium in DSSCs, an ideal photoelectrode should have a high electron
thin films (Figure 22B).199 A far-from-equilibrium method was mobility and simultaneously high surface area to load enough
applied to construct coral-like nanostructures of TiO2 on a sensitizer dye.213,214 Since high surface area usually leads to
variety of surfaces, which consist of anatase-based and rutile high surface defect density, it is important to balance between
nanocoral layers (Figure 22C).200 TiO2 mesocrystals, as a new the high dye-loading capacity and low surface-defect density.
class of porous TiO2 materials, are superstructures of orderly To achieve good electron transport properties, TiO2 NPs
aligned nanocrystals. Nanoplate anatase TiO2 mesocrystals with should have high crystallinity and few defects.215−217 Besides
dominant high-energy (001) facets were newly synthesized by a high-quality crystalline TiO2 NPs, a good TiO2 photoelectrode
topotactic structural transformation, which exhibited greatly should be highly homogeneous, forming a well-interconnected
increased photoconductivity and photocatalytic activity.201 nanoparticle matrix for good electron transfer. The most
For current and future applications, numerous types of TiO2 adopted strategy to prepare a TiO2 paste for this purpose is to
films based on packed 0-D, 1-D, and/or 2-D nanostructures prepare high-quality well-dispersed TiO2 NCs and assemble
have been developed.202−207 TiO2 NP coating on a variety of these nanoparticles into a highly uniform nanoparticle matrix.
substrates can be grown as large oriented TiO2-based arrays by For this purpose, the synthesized TiO2 NCs should be well-
a recently developed templateless approach.208 By this dispersed for further preparation of homogeneous TiO2 NP
approach, highly oriented TiO2 nanoarrays could be obtained films. Sol−gel chemistry combined with hydrothermal
on supporting substrates in contrast to randomly dispersed approaches is the typical method to prepare well-dispersed
nanostructures. Joshi et al. 209 reviewed the synthesis, TiO2 NCs suspensions. These well-dispersed TiO2 NPs need
organization, and device integration from the perspective of then to be fabricated into homogeneous pastes using organic
selected 1-D materials. It is believed that nanotubes will be binders.218 The larger surface area of the resulting mesoporous
superior to nanorods or nanowires for many applications TiO2 photoelectrode is necessary for good dye loading.
because of the exposure of inner as well as outer wall surfaces of However, the electron mobility is affected by grain boundary
tubes for increased interaction with the chemical environment. between TiO2 NCs.215−217
The 1-D nanotubes also can be combined with TiO2 nanorod Due to the better electron transport properties of 1-D TiO2
or TiO2 nanoring to enhance the photocatalytic activity. For nanostructures,219−223 numerous efforts have been directed to
the assembly of nanoarrays, interestingly, not only 1-D the development of 1-D TiO2 structures. However, the
nanotubes or nanowires but also 2-D nanosheet arrays can reported 1-D TiO2 structures used in dye-sensitized solar
grow on the Ti substrate. In view of anatase TiO2 with the cells missed the goal of providing better efficiencies due to a
9296 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

surface area that is too low for dye loading. Very recently, some it can only be activated under UV-light irradiation because of its
experiments were reported to combine the advantage of the large band gap. Given that less than ∼5% of the solar flux
high surface area of 0-D TiO2 and the better transport incident at the earth’s surface lies in this spectral region, one
properties of 1-D TiO2 by breaking 1-D TiO2 into 0-D TiO2 urgent task is to tune the band gap response of TiO2 into the
with larger surface area.224−228 In these reports, 1-D TiO2 visible light region to utilize more natural solar light for
nanotube arrays or nanowire arrays were first prepared by those photocatalytic or photoelectrochemical processes.200 In fact,
well-studied 1-D TiO2 synthesis methods such as electro- much research focuses on the modification of TiO2 to make it
chemical anodization of a titanium foil.206,229−231 Then they sensitive to visible light.128 More importantly, such modifica-
were treated with different chemical etching approaches. Some tions can also be used to inhibit the recombination of
anatase TiO2 nanotube arrays have been turned into anatase photogenerated charge carriers, which limits its overall
TiO2 nanoparticle−nanotube arrays, as shown in Figure 23. photocatalytic efficiency.
6.1. TiO2 NPs as Hosts for Metal and Nonmetal Dopants
Among of the many various modification strategies for visible-
light-driven titania, TiO2 NPs can be hosts for metal or
nonmetal dopants.238,239 The band gap of TiO2 can be
modified effectively by doping with transition metal ions
(e.g., V,240 Co,241 Fe49,242) at Ti sites. Introducing various bulk
irregularities via doping can provide shallow trap states for
conduction band electrons. Vanadium doping led to a red-shift
in the UV−vis absorbance spectra due to a decrease in band
gap compared to pristine titania.243 More importantly,
vanadium doping can play a key role for efficient transfer of
electrons due to its multiple oxidation states ranging from 3+ to
5+ and generation of nonstoichiometry due to the incorpo-
ration of V3+ or V5+ in place of Ti4+.240 For Co-doped anatase
TiO2 nanostructures synthesized by a simple solvothermal
method, additional absorption bands are due to the ligand field
transitions, 4T1g(4F)−4T1g(4P) of Co2+, and also due to
transitions from different trap states related to oxygen
Figure 23. TEM images of two-step anodized TiO2 nanotubes: (A) vacancies.241 Interestingly, the introduction of Co2+ into TiO2
without any post treatment, (B) with hydrothermal treatment.236 nanocolloids (<10 nm) can induce phase transformation at
TEM and HRTEM images of unetched rutile TiO2 nanorods (C) and
room temperature due to lowering the activation energy for the
7 h etched rutile TiO2 nanorods (D).237 Reproduced with permission
from refs 236 and 237. Copyright 2011 American Chemical Society. anatase to rutile phase conversion (Figure 24).132 When
Copyright 2013 Royal Society of Chemistry. comparing to P25, the cobalt doped samples, even with a low
CoCl2 content, exhibit a sharp difference in the visible light
response.132 In the case of the indium-doped TiO2 synthesized
Single-crystal rutile TiO2 nanorods have also been etched into with the precursor indium chloride, the transition of electrons
networks of rutile TiO 2 NPs. 232−237 These new TiO 2 from the valence band of TiO2 to surface-state energy levels
photoelectrodes prepared by breaking 1-D TiO2 into 0-D (the surface O−In−Clx species; x = 1 or 2) is responsive to
nanoparticle fragments have exhibited much better efficiency. visible light.239
Except for metal ion dopants, anions such as iodine or
6. CHEMICAL AND PHYSICAL MODIFICATIONS OF phosphorus can replace lattice Ti4+ to form Ti−O−I or Ti−O−
TIO2 NPS P sites.134,154 In fact, much of the recent doping has been
TiO2 has attracted significant attention as a functional material performed using nonmetal anions such as N,244 C,245 S,246 and
mainly because of its promising applications in solar energy B247 at O sites. The possible changes that might occur in the
conversion as well as in environmental remediation. However, band gap electronic structure of anatase TiO2 upon doping with

Figure 24. (A) Raman spectra for a TiO2 nanocolloids (<10 nm) prepared with various concentrations of CoII using CoCl2. The typical phonon
modes of rutile TiO2 are indicated by solid vertical lines. (B) Reflectance spectra of P25 and cobalt-doped TiO2 nanoparticles with varying CoII
content.132 Reproduced with permission from ref 132. Copyright 2010 Royal Society of Chemistry.

9297 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

Figure 25. (A) Various schemes illustrating the possible changes that might occur to the electronic band gap of anatase TiO2 upon doping with
various nonmetals: (a) band gap of pristine TiO2, (b) doped TiO2 with localized dopant levels near the valence band (VB) and the conduction band
(CB), (c) band gap narrowing resulting from broadening of the VB, (d) localized dopant levels and electronic transitions to the CB, and (e)
electronic transitions from localized levels near the VB to their corresponding excited states for Ti3+ and F+ centers.248 Diffuse reflectance spectra of
undoped and N-doped TiO2 NPs sintered at different temperatures in N2 (B) and in air (C).256 Reproduced with permission from refs 248 and 256.
Copyright 2008 and 2006 American Chemical Society.

various nonmetals are listed in Figure 25A.248 Burda and his co- produce different nitrogen concentrations and visible light
workers reviewed the chemical synthesis, physical properties, absorbances (Figure 25B,C).256 In other studies,257 boron
and applications of nitrogen-doped metal oxide nanoparticles, doping produces smaller band gap n-type semiconductors,
with special emphasis on titania-based NPs.249 TiO2−xNx NPs while nitrogen doping produces p-type or n-type semi-
can be prepared by employing the direct amination of conductors depending on whether or not nearby oxygen
nanosized titania particles (6−10 nm).250 Doping nitrogen atoms occupy surface sites. That is, for undercoordinated TiO2
into TiO2 on the nanometer scale can lead to higher nitrogen surface structures found in small clusters, nanorods and
concentrations than that in thin films and micrometer-scale nanotubes, layers, and surfaces, doped nitrogen gives rise to
TiO2 powders, resulting in enhanced catalytic and bactericidal acceptor states, while for larger clusters and bulk structures
activity.251 TiO2−xNx NPs also show switchable surface nitrogen dopants act as donor states.
wettability upon visible light illumination, which is crucial for
6.2. Codoping
the photocatalytic activity and decomposition efficiency of
organic molecules on its surface.252−254 For nonmetal light- To enhance the visible light efficiency, much research has been
element dopants, the modified optical properties are, at least undertaken to modify the electron structure of TiO2 and to
partially, due to the electronic coupling of the dopant 2p or 3p narrow its band gap by doping with either anions or cations;
orbitals to Ti 3d orbitals.246 By means of X-ray absorption, X- this body of work has been reported and reviewed.258−271
ray emission, and XPS, it can be found that main-group element However, the UV photocatalytic activities of these doped TiO2
doping of TiO2 leads to an increase of the density of states nanomaterials were found to be low due to carrier
above the TiO2 valence band edge which leads to visible light recombinations induced by the dopant. Recent studies focus
absorption.129,248,255 During the synthesis of N-doped TiO2 on the modification of TiO2 by codoping. The coupling of the
NPs, the sintering atmosphere affects the crystallization and can dopants can enhance the response to visible light more
9298 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 26. (A) Optical absorption spectra taken from the Cr/N codoped nanoclusters (red) and from an undoped TiO2 nanocluster film (gray).277
(B) Absorption spectra of pure, N-doped, Fe-doped, and N/Fe codoped TiO2.279 Reproduced with permission from refs 277 and 279. Copyright
2012 and 2010 American Chemical Society.

efficiently. The coupling of one dopant with a second or more of their optical properties to utilize the solar energy more
dopants has been proven to enable a reduction in the number efficiently.228,232,280−291
of carrier recombination centers by the proposed charge One-dimensional photonic crystals or Bragg mirrors are the
equilibrium mechanism.272 (C,N)-TiO2NPs,273 Co/(C,S)- group of multilayered structures in which the refractive index
TiO2NPs,274 (Nb,Ta)/(C,N)-TiO2 nanostructures,272 N/Pd- varies periodically along one dimension, which was first studied
TiO2(0.8% Pd) NPs,275 Cr/N-TiO2 nanoclusters,276 C−Cl- by physicist Rayleigh in the 19th century. Photonic crystals
codoped anatase TiO2 NCs,277 and Nb/N-codoped anatase attracted much research interest after Yablonovitch and John
TiO2 nanocomposites278 have been studied by theoretical and extended in 1987 the optics of varying refractive index toward
experimental methods and proved to be beneficial or synergistic 2- and 3-D periodic structures.292,293 The periodic refractive
for visible light absorption and electron−hole pair separation index modulation gives rise to optical interference effects, which
over single-doped TiO2 with comparable surface areas. In increase the absorbance of certain wavelengths near the stop
(C,N)-TiO2NPs, N atoms could incorporate into the lattice of band. This effect is illustrated in Figure 27A,B. It can be
anatase through substitution of oxygen atoms, while most C adjusted by the structural and optical parameters of the
atoms could form mixed layers of deposited active carbon and multilayer. The titania−air periodic structure demonstrates a
complex carbonate species at the surface of TiO2 nano- strong 1-D photonic crystals effect.294 A typical TiO2 inverse
particles.273 For (Nb,Ta)/(C,N)-TiO2 nanostructures, the opal is constructed by using monodisperse polystyrene
codoping with transition metals facilitates the enhancement microspheres as template to form the photonic crystal
of p-type dopants (N and C) in the host lattice.272 structure. Then the gaps are infiltrated with TiO2 NPs paste
Nanostructured Cr and N codoped TiO2 thin films obtained and finally the polystyrene template is removed to increase the
by supersonic cluster beam deposition facilitate a high refractive index contrast. Their typical morphology is shown in
concentration of dopants and furthermore a reduced band Figure 27C,D.295 Ordered photonic TiO2 NPs have been
gap (Figure 26A), which can overcome the limitation of the low successfully utilized both in DSSCs and photocatalysis296,297
solubility of dopants inside the TiO2 matrix and produce highly and provided increased photovoltaic and photocatalytic
promising photoactive nanophase materials.271 However, N/ efficiency by increasing the light absorbance near the stop
Fe-codoped TiO2 shows worse photocatalytic performance band of the photonic crystal.298
than pure TiO2 under both UV and visible light (Figure In DSSC applications, the photonic crystal structures are
26B).279 On the basis of XPS and low-temperature PL, both N usually related to mesoporous TiO2 NP films to form a bilayer
and Fe facilitate the absorption of the visible light, but it is photonic crystal nano-TiO2 photoelectrode.299,300 In these
found that only Fe increases the electron−hole recombination structures, the good physical contact between photonic crystals
rate, leading to the opposite effects of N and Fe doping on the and the nano-TiO2 photoelectrode is important to achieve a
photocatalytic performance of TiO2. Therefore, it is evident strong shift in light absorption due to the stop band of the
that doping can produce its own disadvantage: notably, the photonic crystal.299 Otherwise, the photonic crystal layer can
defects within the band gap of the semiconductor to trap only increase light absorption by scattering, which is then
photogenerated charge carriers and decreases its activity. similar to normal scattering layers in DSSCs.299 The contact
Currently, this research area still needs much improvement, problem is mainly caused by the bilayer structure. Although
so that TiO2 codoping with two or more dopants can achieve some advanced techniques have been developed to overcome
significant synergistic effects compared to their single-ion the contact issue in bilayers,300 it is better to avoid such
doped or undoped TiO2 counterparts. structure. The aim of any bilayer structures is to load more dye
onto the TiO2 NCs films. Due to the large air void, the
6.3. Photonic Crystals Constructed from TiO2 NPs photonic structure cannot absorb enough dye. In order to
Other than the above-mentioned chemical modifications to achieve better photonic crystal effects, recent hierarchical
broadening the photoactive spectrum of TiO2 and increasing its photonic crystal structures use 5−6 nm sized TiO2 NCs.301
photocatalytic and photoelectrical efficiencies, it is possible to In this structure, the TiO2 NCs layer disappears and the
physically organize the TiO2 NPs in ordered structures, photonic TiO2 backbone comprising 5−6 nm-sized TiO2 NPs
specifically, photonic crystals, which allows for manipulation can absorb enough dye onto the TiO2 photonic nanocrystal
9299 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

photonic TiO2 structures are good recyclable photocatalysts.307


Furthermore, Ozin and co-workers have demonstrated that
these photonic crystal structures of TiO2 NP films have a high
structure-disorder tolerance to sustain their high photocatalytic
activities for potential environmental cleanup and water
treatment applications.297 One important aspect for photonic
structure TiO2 NPs is that the stop band of the photonic crystal
should match with the band gap of TiO2. There is no enhanced
photocatalytic activity found in those photonic TiO2 NPs
where the stop band is in the visible region, as shown in Figure
29.

Figure 27. (A) Simplified optical band structure of a photonic crystal.


At photon energies approaching a full band gap or a stop band from
the red side, the light absorbance can be confined in the high refractive
index part of the photonic crystal. At energies above the band gap or
stop band, the standing wave is predominantly localized in the low-
index part of the photonic crystal, i.e., in the air voids. (B) Illustration
of the effect of standing wave localization on dye absorbance. In an
isotropic medium, the dye absorbs strongly in the blue but weakly in Figure 29. Process of phenol photocatalytic degradation under UV
the red (heavy line). If the stop band is tuned to the position shown by light illumination. The stop bands of these photonic TiO2 NPs are at
the arrow, the blue absorbance is diminished and the red absorbance is 280, 320, 450, and 525 nm.303 Reproduced with permission from ref
increased when the dye is confined to the high refractive index part of 303. Copyright 2010 American Chemical Society.
the photonic crystal (dotted line).298 (C) SEM image of a typical
photonic crystal template constructed by polystyrene microspheres.
(D) SEM image of the templated inverse-opal TiO2 NPs films showing
well-formed inverse-opal structures.295 Reproduced with permission 6.4. Ti3+ Self-Doped TiO2
from refs 295 and 298. Copyright 2009 Wiley-VCH Verlag GmbH &
Co. KGaA, Weinheim. Copyright 2003 American Chemical Society. In situ self-doping with Ti3+ is regarded as one of the promising
strategies of chemical modifications of TiO2 NPs. In fact, ideal
film. The comparison of the bilayer and hierarchical photonic crystals are not stable thermodynamically due to a configuration
crystal structure is shown in Figure 28. entropy term that requires the formation of point defects. Thus,
TiO2 is a nonstoichiometric compound. Accordingly, its
formula has been assumed as TiO2‑x, where x is the deficit of
oxygen. The deficit of oxygen can be interpreted as a
combination of oxygen vacancies and Ti interstitials.308 In
order to obtain the desired properties, it is necessary to increase
the understanding of the defect chemistry of TiO2 and the
relationship between defect disorder and the electrical proper-
ties.309,310
Defects in TiO2 can exist intrinsically or be formed as a result
of interactions between the TiO2 lattice and oxygen in the gas
Figure 28. Schematic illustration of (A) bilayer photonic crystal-nano- phase. Generally, TiO2‑x, which contains Ti3+ ions or oxygen
TiO2 photoelectrode300 and (B) photonic crystal formed by vacancies, can be produced by heating TiO2 under vacuum or
hierarchical titania frameworks.301 Reproduced with permission from
refs 300 and 301. Copyright 2011 and 2010 American Chemical
reducing conditions (e.g., H2).311,312 High energy particle
Society. (laser, electron, or Ar+ ion) bombardment is also used to create
Ti3+ sites in the surface of TiO2‑x.313−315 Figure 30 shows that
the Ti3+ content in the TiO2 annealed in 5% H2/N2 is higher
TiO2 materials have been reported to be designed and than that annealed in pure N2. That means that in oxygen-free
structured into hierarchical photonic crystal with different pore atmosphere and reducing conditions one can increase the Ti3+
sizes. The photonic structure with stop bands overlapping with content.312 Moreover, reduced TiO2 has been demonstrated to
the absorption of TiO2 show doubled photocatalytic hydrogen exhibit visible light absorption. Most importantly, more defects
generation activities compared to the random TiO2 NCs due to in TiO2 can help to provide more surface active sites and
the increased light absorption in photonic structure.302 The decrease the transport resistance of photogenerated electrons,
photonic crystal structure of TiO2 films shows much better which allows the particles to obtain larger charge carrier density
photocatalytic activities for the decomposition of organic dyes and higher hydrogen production rates through water
compared to the normal solid TiO2 NCs films.303−307 These splitting.312,316
9300 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 30. XPS spectra in Ti 2p region for anodized TiO2 nanotubes annealed (a) in N2 and (b) in 5% H2/N2.312 Reproduced with permission from
ref 312. Copyright 2011 American Chemical Society.

Figure 31. (A) Experimental (, measured under 75 K) and simulated (---) EPR spectra for reduced TiO2 sample. (B) UV−vis diffuse reflectance
spectra for commercial anatase TiO2 () and sample (reduced TiO2, ---). (C) Time course of evolved H2 under visible light (>400 nm) irradiation.
(a) Reaction for 4 h. (b) Evacuated and continued reaction for another 4 h. (c) Illuminated for 200 h, evacuated system, and continued reaction for 4
h:319(●) sample (reduced TiO2); (▼) anatase. Reproduced with permission from ref 319. Copyright 2010 American Chemical Society.

However, the surface Ti3+ ions are usually not stable in air as studies imply that synthesis of a stable Ti3+ self-doped TiO2 is a
they tend to adsorb atmospheric oxygen or dissolved oxygen in simple and economical method for narrowing the band gap
water and so are oxidized to Ti4+.317,318 Therefore, it is which can promote the development of a highly active
intriguing to synthesize a stably reduced TiO2 photo- photocatalyst for visible light irradiation.
catalyst.319−321 The reported methods include a facile one-
step combustion method and a one-step solvothermal method. 7. TIO2 NP HETERONANOSTRUCTURES
Among different synthetic parameters, the reductant or the 7.1. Mixed Phase TiO2
source of the reducing agent, imidazole,319 NaBH4,320 or
hydrated hydrazine,321 has significant effects on the phase In general, it is well-known that anatase TiO2 NPs show higher
composition, UV−vis absorption, surface area, and photo- photocatalytic and photoelectrochemical activities than rutile
catalytic activity.322 It is reported that the combination of TiO2 NPs.326−328 In order to achieve highly active TiO2 NPs,
microwave heating and ethylene glycol, a mild reducing agent, most reported TiO2 syntheses were designed to obtain phase-
pure anatase TiO2 NPs. However, the commercially available
can induce Ti3+ self-doping in TiO2 photonic crystals.323 The
photocatalysts, particularly Degussa P25 TiO2, are a mixed
captured light by the photonic crystal layer is absorbed by the
phase TiO2 NP sample with about 70% anatase and 30% rutile.
self-doped interbands, which lead to significant enhancement of
The mixed phase P25 was found to be more active than the
the photoelectrochemical performance. phase-pure anatase or rutile one. Thus, many mixed phase TiO2
To test for the presence of Ti3+, low-temperature EPR NPs were synthesized and exhibited enhanced photocatalytic
spectra were recorded (Figure 31A).319 Anisotropic powder activities.218,304,329−335 One of the most fundamentally
pattern g-values of gx = gy = 1.975 and gz = 1.944 are intriguing issues in TiO2 photocatalysis is the proposed
characteristics of a paramagnetic Ti3+ center in a distorted synergistic effect between anatase and rutile at appropriate
rhombic oxygen ligand field.324 It also indicates that there is no ratios on their optoelectronic and photocatalytic properties.
Ti3+ present on the surface of as-prepared TiO2 because surface The difference in the energy levels of anatase and rutile
Ti3+ shows an EPR signal at g ≈ 2.02.325 The UV−vis conduction bands acts as an effective charge separator.326
absorption spectra (Figure 31B) show the extension of light Electron and hole transfer between the two phases can induce
absorption which is proposed to originate from a miniband increased lifetimes of electrons and holes. However, the
arising closely below the conduction band minimum of regular negative effects of the presence of rutile was also discussed.
TiO2. As shown in Figure 31C, Ti3+ sites in the reduced TiO2 The presence of rutile can result in inferior performance due to
exhibited high stability in air and water with light irradiation. a decrease in the effective hole concentration at the anatase
The H2 evolution rate also is stable in repeated experiments; surface.326
that is, the reduced TiO2 can be repeatedly used without Although it is a widely accepted model of enhanced charge
degradation in activity. Similar results were found in the transfer, there is one long-standing controversy on the energetic
photodegradation of methyl orange and phenol.320,321 These alignment of the band edges of the rutile and anatase
9301 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 32. (A) Two proposed valence and conduction band alignment mechanisms for the anatase/rutile interface: a, type-II (rutile); b, type-II
(anatase). Red arrows indicate the flow of electrons (holes) in the conduction band (valence band). Blue and orange dots represent electrons and
holes, respectively. (B) Ti 2p3/2 spectra are taken from two phase composite particles with rutile to anatase ratios of 1:1 (top) and 2:1 (middle) and
1:2 (bottom). Experimental data, shown as black dots, are fitted with the peak shapes derived from phase-pure anatase (red) and rutile (blue). (C)
Schematic of the XPS alignment between rutile and anatase. ΔECL is the core level offset between the Ti 2p3/2 core levels.337 Reproduced with
permission from ref 337. Copyright 2013 Macmillan Publishers.

TiO2,337,338 shown as in Figure 32A. On the basis of titania core.354−358 When the semiconductor and metal
electrochemical impedance analysis, the flat band potential of nanoparticles are in contact, the photogenerated electrons are
anatase is ∼0.2 eV more negative than that of rutile,339 placing distributed and transferred until the semiconductor/metal
the conduction band of anatase 0.2 eV above that of rutile composite system attains equilibration. The electron accumu-
(Figure 32A-a). However, it is revealed that the conduction lation on the metal deposits shifts the Fermi level of metal to
band of anatase lies 0.2 eV below that of rutile, in line with the more negative potentials, and the resultant Fermi level of the
work function of rutile, which is 0.2 eV lower than that of composite shifts closer to the CB of the semiconductor. The
anatase on the basis of recent photoemission measurements340 negative shift in the Fermi level is an indication of better charge
(Figure 32A-b). The latest report in 2013 supports the latter by separation and more reductive power of the composite system,
combined periodic hybrid DFT calculations and XPS resulting in possible enhancement of the photocatalytic process.
experimental results (Figure 32B,C). According to the In addition to the size and shape of the deposited metal
demonstrated model, the electron affinity of anatase is higher nanoclusters, the coupling and morphology between the metal
than that of rutile, and photogenerated conduction electrons and the TiO2 NPs are important factors for metal-modified
flow from rutile to anatase. TiO2 photocatalysts because they influence the catalytic
This mechanism study points out that the charge transport properties and performance characteristics of the whole system.
processes critically depend on the interface between the TiO2 The metal particles change the Schottky barrier height between
phases and particle size. The simplest way to prepare mixed the metal and the TiO2, affecting the efficiency of charge
phase TiO2 is to directly mix different ratios of anatase and separation, while the variation in the morphology of the metal
rutile TiO2 NPs together;341−344 however, the interface contact particles influences the manner of adsorption, affecting the
between anatase and rutile is not ideal by just mechanical photocatalytic activity. When Rh, Pt, and Au metal particles
mixing. In order to obtain mixed phase TiO2 NPs with better were photodeposited on TiO2, the spectral analysis shows that
interface contact, two main strategies can be employed. One is a strong metal−support interaction takes place in the case of Rh
to convert anatase TiO2 into mixed phase TiO2, and the other and Pt, resulting in the disorganized structure, while Au metal
is to directly synthesize mixed phase TiO2. The most particles showed a well-organized and spherical structure,
commonly adopted synthesis approach for mixed phase TiO2 indicating that Au does not strongly interact with TiO2.359
is to first achieve the anatase or amorphous TiO2 NPs, These metal particles were partially covered by TiO2 (more
aggregate, and then anneal these aggregates at a high specifically, TiO2−δ), which provides electrons to the surface of
temperature to control the different anatase-to-rutile ratio in
metal particles. In another computational study, strong
the mixed phase TiO2.345−348 Another approach to directly
interactions between gold and anatase TiO2(001) have been
prepare mixed phase TiO2 through synthesis is to form anatase
predicted by a DFT approach,360 which is beneficial for
TiO2 NPs and to cap them with rutile TiO2 to get mixed phase
experimentalists to design novel catalysts.
TiO2 at different ratios.218,304,349−352 Moreover, synthesis of
To develop metal-decorated TiO2 NPs, a photocatalytic
TiO2 with different ratios of anatase or rutile to brookite can be
controlled by adjusting the concentration of Na+ and the pH of linker molecule, e.g., Keggin ions, can be utilized to bind to the
the solution55,113,114 or by changing the volume ratio of the TiO2 surface to reduce Au nanoparticles via photoirradiation.
“adjusting reagent” triethylamine to water.353 The brookite− The Au nanoparticle-decorated Keggin ion/TiO2 architecture
anatase mixture also exhibited the enhanced activity due to a led to a dramatic increase in the photocatalytic performance of
synergistic effect between TiO2 polymorphs.51,55,59 the composite system, as opposed to a TiO2 surface directly
decorated with metal nanoparticles.361 Cozzoli et al.362 have
7.2. TiO2−Metal Nanostructures reported the synthesis of organic-soluble Ag−TiO2 composites
7.2.1. Improving Charge Separation with Metal NPs. using surfactant capped TiO2 nanorods as stabilizers for the
Another effective modification method to enhance the silver particles. In order to have more sites available for Ag
functionality of TiO2 NCs by increasing the visible light clusters to grow, a facile photodeposition technique was
response and increasing the lifetime of photoexcited electrons developed to make small Ag clusters highly dispersed on
and holes is to deposit a noble metal on the surface of the individual TiO2 nanocrystal surfaces. The population of Ag
9302 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

Figure 33. TEM images of hybrid Ag/TiO2 NCs synthesized with additional OA and UV-irradiated for 40 min: (A) OA:TiO2 = 1:1, and (B)
OA:TiO2 = 0.1:1. (C) Formation of colloidal hybrid Ag−TiO2 NCs upon UV light irradiation.354 Reproduced with permission from ref 354.
Copyright 2011 American Chemical Society.

clusters can be controlled by tuning the amount of OA added crystal grow preferentially on the surface of gold nanocrystals
to the synthesis solution, and their size by irradiation time stabilizing the heterointerfaces. On the initial deposition TiO2
(Figure 33).354 The thus-obtained hybrid Ag/TiO2 nanocrystals layers produced in the “budding” process, the TiO2 crystals
exhibit high photocatalytic performance and unique optical grow into truncated wedge-shaped morphologies. The resulting
properties ranging from those of colloidal Ag nanoparticles to nanostructured core−shell Au@TiO2 photocatalyst presents
those of aggregates of Ag nanoparticles.354 A 1-D film structure high photocatalytic activity when exposed to UV or visible light
composed of TiO2 single crystals was coated with ultrafine Pt irradiation, which is largely attributed to the preferentially
nanoparticles (0.5−2 nm) via a gas phase deposition method, grown TiO2 shell structures and metal (Au)−TiO2 hetero-
which exhibited extremely high CO2 photoreduction efficiency interfaces. As for the core−shell nanocomposites of M@TiO2
with selective formation of methane.355 (M = Au, Pd, Pt) synthesized via the hydrothermal route, it is
7.2.2. Core−Shell Metal@TiO2. In the cases of section interesting that the incorporation of noble metal cores into the
7.2.1, the metal nanoparticles are anchored on the surfaces of shells of TiO2 only contributed to the enhancement of visible
TiO2 as isolated “nanoislands” to produce heterointerfaces and light photocatalytic activity of TiO2. 366 However, the
trap the photogenerated electrons. This type of noble metal− incorporation of noble metal cores also significantly lowered
titania heterostructure, M/TiO2, provides only two-dimensional the UV light absorption on the titania part, thus leading to a
interaction of the TiO2 with noble metals, which leaves the lower UV photocatalytic activity of M@TiO2 compared to bare
majority of the noble metal atoms unused. Moreover, the noble P25 NPs. 366 Photoactivity tests demonstrate that the
metal particles may be corroded or dissolved during a incorporation of metal cores into the shell of TiO2 will inhibit
photocatalytic reaction. On account of these problems, photocorrosion behavior and provide much better photo-
heterostructures with metal cores and semiconductor shells catalytic stability of M@TiO2 (M = Au, Pd, Pt) nano-
have attracted much attention.363,364 The often unstructured composites. Therefore, metal@TiO2 core−shell nanocompo-
shells in core−shell metal@TiO2 photocatalysts may expose sites are considered to be some of the strongest photocatalysts
low reactivity crystal facets or have small specific surface areas due to their desirable chemical composition, controllable
due to the inherent limitations of core−shell heterostructures. surface chemistry, photostability within the TiO2 shell, three-
Thus, tailoring the shell structures and morphologies has dimensional electronic interaction, and the resulting effective
recently become an interesting new research objective. The charge transfer between the metal core and TiO2.
availability of new chemical properties of heteronanostructures 7.2.3. Increasing UV and Visible Light Absorption
depends completely on their correct architecture. Novel core− with Plasmonics NPs. The concept of plasmonic photo-
shell Au@TiO2 NPs with truncated wedge-shaped morphology catalysis367 has been recently proposed and demonstrated as an
have been recently synthesized by a relatively simple hydro- effective approach to achieve a higher photocatalytic efficiency
thermal route (Figure 34).365 Depending on the produced F− for TiO2 NPs, similar to extending visible light activity by
ion concentration from hydrolyzing the TiF4 precursors, TiO2 chemical doping and using noble metal particle interactions. In
the plasmonic TiO2 photocatalysts, plasmonic NPs show a very
intense localized surface plasmon (LSP) band in the near-UV
region. The LSP causes a considerable enhancement of the
electric near field in the vicinity of the plasmonic NPs, and it
can boost the excitation of electron−hole pairs in TiO2, thus
increasing the efficiency of the photocatalyst.367−369 Figure 35A
illustrates how the plasmonic NPs increase the enhancement of
light absorption.369 The enhanced light absorption can then
elevate the photocatalytic activities, as shown in Figure 35B.367
Currently, the most common plasmonic nanomaterial that
has been applied to modify TiO2 NPs are Au (λPl ∼ 525 nm)
Figure 34. Schematic diagram of the proposed formation process of and Ag (λPl ∼ 420 nm). Because of the strong photo-oxidation
core−shell Au@TiO2 NPs with truncated wedge-shaped morphol- properties of the photogenerated holes on TiO2 NPs, they can
ogy.365 Reproduced with permission from ref 365. Copyright 2009 easily oxidize the Au or Ag NPs. Thus, insulation layers of SiO2
American Chemical Society. are usually coated onto the Au or Ag NPs to form a stable
9303 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

controlled; thus, SnO2/TiO2 heterostructures with SnO2 NPs


or nanorods were fabricated. On the basis of the analysis of
band gap, electron affinity, work function, and PL emission of
SnO2 and TiO2, efficient charge separation increases the
lifetime of the charge carriers and enhances the efficiency of the
interfacial charge transfer to adsorbed substrates, and therefore
accounts for the higher activity of the heterostructured SnO2/
TiO2 photocatalyst.422 SrTiO3, a well-known cubic-perovskite-
type multimetallic oxide with a band gap of 3.2 eV comparable
to TiO2, can offer favorable energetics for photocatalysis
because its conduction band edge is 200 mV more negative
Figure 35. (A) Image illustrating the energy flux (Poynting vectors)
around the plasmon NPs and the electric field intensity for an incident
than TiO2.423,424 In order to have good contact between
electromagnetic wave with an electric field in the plane of the image. SrTiO3 and TiO2 NPs, it is necessary to design and prepare the
Red/blue stands for high/low electric field intensity, respectively.369 heterostructures with favorable size, shape, and surface−
(B) Comparison of the photocatalytic decomposition rate of interface energy. Nanofibers of SrTiO3/TiO2 heterostructures
methylene blue (MB) under near-UV irradiation. (a) A TiO2 film were fabricated by an in situ hydrothermal method using TiO2
on a SiO2 substrate. (b) A TiO2 film on an Ag/SiO2 core−shell as both template and reactant.425 Notably, by simply tuning the
structure on a SiO2 substrate. (c) A TiO2 film on a Ag/SiO2 core− precursor Sr(OH)2 concentration or reaction temperature, the
shell structure.367 Reproduced with permission from refs 369 and 367. density as well as the morphology of the SrTiO3 nanostructures
Copyright 2012 Royal Society of Chemistry. Copyright 2008 can be controlled in order to improve the photocatalytic
American Chemical Society.
properties of TiO2 to a greater degree.
The above-mentioned examples shall illustrate that TiO2
nanoheterostructures of plasmonic NPs/TiO2 NPs.268,367,370 NCs have become hosts and centers for fine-tuning chemical
The thickness of the SiO2 needs to be optimized to balance the functionality. From the viewpoints of solar energy conversion
photostability of the plasmonic NPs and the field enhancement to photochemical reactions and photocatalysis, the semi-
by the plasmonic NPs. Since the LSP bands of nanostructured conductor sensitizer should have a strong oscillator strength
Ag appear around 400 nm, most plasmon TiO2 photocatalysts in the visible region. At the same time, the CB edge potential of
adopt plasmonic Ag NPs.286,367,370−395 Au NPs have also been the sensitizer should be higher than the TiO2 CB edge to
utilized to modify TiO2 NPs for visible light photocataly- facilitate electron transfer. The formation of the heterojunction
sis.368,396−400 The simplest way to prepare Ag/TiO2 nano- between a narrower band gap semiconductor and TiO2 with
composites is to prepare Ag NPs and TiO2 NPs separately and matching band potentials provides an effective way to extend
then mix them together.286 In order to attain optimized the photosensitivity of TiO2 into the visible part of the solar
plasmonics-induced light absorption, Ag/SiO2/TiO2 core−shell spectrum. Combinations of TiO2 nanostructures with various
plasmonic photocatalysts, such as Ag/SiO2/TiO2 and Au/TiO2, semiconductors, such as In2O3 (Eg = 2.8 eV),426 Bi2WO6 (Eg =
have been prepared by sol−gel synthesis.396,401,402 2.8 eV),427 and ZnFe2O4 (Eg = 1.86 eV),428 have been
Other than enhancing the light absorption of TiO2 NPs, successfully prepared and show enhanced visible light photo-
plasmons in metal/TiO2 core−shell structures can also support catalytic activity. In the case of In2O3/TiO2 heteroarchitectures,
electron transfer to the semiconductor scaffold in photo- the photocatalysts could be reclaimed by sedimentation without
electrochemical cells, in which the plasmonic metal NPs work a decrease in photocatalytic activity.426 Both the CB and VB
as a sensitizer to utilize visible light to generate photoelectrons, edges of TiO2 lie energetically below the band edges of In2O3.
very similar to dye-sensitized solar cells.369,403,404 Plasmonic Thus, the In2O3 in the heterostructure could be excited under
metal/semiconductor composites for photoelectrochemical visible light irradiation, and the electrons generated in the In2O3
water splitting have recently been reviewed elsewhere.369,404 migrated to the CB of TiO2. As confirmed by PL emission, the
7.3. TiO2−Semiconductor Nanoheterostructures heterojunction increased the lifetime of the charge carriers and
Semiconductor nanostructures have been utilized abundantly enabled efficient interfacial charge transfer to TiO2.
for light-harvesting (solar cells and photocatalysis),405,406 light The II−VI semiconductors, represented by CdS429−431 and
generation (light emitting diodes),407−409 and other newer CdSe,432−435 particularly in the form of quantum dots, are
energy conversion schemes, such as thermoelectrics.410,411 widely utilized as photosensitizers.436 CdS has a narrower band
One effective way to affect charge transfer of semiconductors gap (2.4 eV), and its conduction band is 0.5 eV more negative
and hence boost carrier collection is to couple them with other than that of TiO2. The band gap of quantum dots (QDs)
semiconductors having appropriate electronic band structures enables tunability through control of the QD size, which allows
and band edges. For example, coupling of TiO2 nanostructures matching of the absorption spectrum to the sunlight spectral
with a wide band gap semiconductor, like SnO2,412,413 distribution. In addition, semiconductor QDs possess higher
SrTiO3,414,415 ZnO,416−418 Nb2O5,419−421 to form a hetero- extinction coefficients and greater intrinsic dipole moments
junction has been widely used to enhance the photocatalytic than molecular dyes, leading to efficient light harvesting and
performance of photocatalysts due to the promotion of spatial exciton formation.437 Moreover, colloidal QDs can add
separation of electrons and holes. Hierarchical SnO2/TiO2 efficiency through their multiexciton-generation (MEG)
heterostructured photocatalysts can be fabricated by combining character when the semiconductor band gap is narrow and
the electrospinning technique with the hydrothermal meth- the excitation energy is at least twice that of the band gap.438
od.422 The SnO2 nanostructures were uniformly distributed Accordingly, among the most attractive configurations to
without aggregation on TiO2 nanofibers. By adjusting exploit these fascinating properties is the quantum-dot-
fabrication parameters, the morphology as well as coverage sensitized TiO2 electrode for solar cells or photoelectrochem-
density of secondary SnO2 nanostructures could be further ical water-splitting.434,437,439 QDs can be adsorbed or anchored
9304 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

on TiO2 through in situ growth of QDs by chemical bath monolayers. The enhanced adsorption to mixed monolayers is
deposition (CBD), deposition of presynthesized colloidal QDs attributed to decreased disulfide formation, caused by the
by direct adsorption (DA), and deposition of presynthesized dilution of thiol groups within the monolayer.434 This method
colloidal QDs by linker-assisted adsorption (LA).440 For CBD allows for precise control of the particle size. However, the
samples with the lower recombination rates by electron transfer insulating molecules intervening between QDs and TiO2 at the
to electrolytes as well as the higher light absorption, only junction could slow the interfacial electron transfer. In addition,
moderate increases of photocurrents are obtained, which the loading level of QDs is limited to below monolayer
indicates the presence of additional recombination pathways coverage, and thus, the amount of visible light absorbed by the
in QD sensitizer layers.440 To better understand the QDs becomes overall low. Optimizing the quantum dot surface
recombination mechanisms of quantum-dot-sensitized solar coverage is necessary to maximize the light harvesting and
cells (QDSCs), a physical model has been presented on the energy conversion efficiencies. In contrast with MPA-mediated
basis of the recombination through a monoenergetic TiO2 QD bonding, the direct adsorption leads to a high degree of
surface state, which takes into account the effect of the surface QD aggregation,441 as revealed by atomic force microscopy.433
coverage.440 At present, the successive ionic layer adsorption and reaction
As illustrated in Figure 36, the deposition of uniform layers (SILAR) technique is believed to be the best way to prepare
of colloidal QDs onto TiO2 NP substrate surfaces is an metal sulfide QD-loaded TiO2 films. However, it is not well-
suited for light harvesting applications in nanoparticulate
systems because several repetitions of QDs adsorptions are
necessary. Although CBD or SILAR techniques ensure high
optical densities in the visible region, the precise control of
particle size and crystallinity, afforded by typical bulk solution-
based syntheses, are sacrificed. A variety of surface treatments
were investigated in order to evaluate the efficiency of QD
adsorption and sensitization for photocurrent generation on
single crystals of low index anatase and rutile surfaces.435
Hiroaki Tada et al.430,436 have recently developed simple and
versatile low-temperature photodeposition techniques for
Figure 36. CdSe QDs adsorption on TiO 2 occurs through directly coupling metal sulfide QDs onto TiO2 by taking
submonolayer formation followed by particle aggregation onto advantage of its photocatalysis and the photoinduced surface
already-adsorbed QDs.441 Reproduced with permission from ref 441. superhydrophilicity. The features of the photodeposition
Copyright 2011 American Chemical Society. technique were summarized showing excellent reproducibility,
tunable band energies, and uniform deposition on external/
important challenge in nanofabrication.441 Deleterious inter- inner surfaces of mesoporous-TiO2 and relatively simple, low-
molecular interactions within monolayers must be minimized cost production. Apart from CdS, CdSe, PbS,439 and PbSe,443
to efficiently adsorb nanoparticles onto substrate surfaces. The many other sulfide or selenide QDs, like I−III−VI ternary near-
bifunctional molecular linker mercaptopropionic acid (MPA) is infrared CuInS2444 and AgInS2,445−447 have been shown to act
frequently used to form self-assembled monolayers of metal as heavy-metal free photosensitizers with promising efficiencies.
sulfide QDs on TiO2. CdSe QDs adsorb to MPA-functionalized The CuInS2 QDs greatly enhance the absorption of UV and
TiO2 surfaces with a thiolated surface through the carboxylic visible light due to the p−n heterojunction being constructed
acid group. Kamat and co-workers442 reported that mercapto- between p-type CuInS2 QDs, and n-type TiO2 NPs.444
propionic acid (MPA) is a better linker than thiolacetic or Compared to CdSe QDs, core/shell QDs, like CdSe/
mercaptohexadecanoic acid (MHDA). Higher CdSe QDs CdS448and CdSe/ZnS,449 can protect photoexcited electrons
surface coverages were measured on mixed monolayers of in the CdSe core from surface trapping and facilitate the
MHDA and hexadecanoic acid (HDA) than on pure MHDA transfer of the photoexcited electron from the QD to the TiO2

Figure 37. Photoactivated processes occurring in irradiated TiO2 NPs451 and the effects of particle volume and surface on photocatalytic reactivity
due to carrier recombination or trapping.452 Reproduced with permission from refs 451 and 452. Copyright 2008 and 2009 Elsevier.

9305 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

substrate, resulting in significantly improved electron injection catalytic efficiency. A higher crystallinity of TiO2 NPs is often
efficiencies. required and is a positive factor as shown in Figure 37.336,451
Bian et al.201 found that anatase TiO2 mesocrystal super-
8. TIO2 NPS AS CHARGE SEPARATION CENTERS structures with high surface area and large (001) facet exposure
could largely enhance charge separation and had remarkably
In order to further improve the photocatalytic and photo- long-lived charge separated states. The compact and dense
electrochemical performance of TiO2 NPs, a fundamental packing of TiO2 NPs forming a uniform agglomerate enables
mechanistic understanding of their surface reactions is efficient charge separation through interparticle charge transfer,
important. In photocatalytic and photoelectrochemical pro- resulting in enhanced photocatalytic activity.455 The location of
cesses involving TiO2 NPs, there are three initial steps that hole and electron traps on specific facets and their spatial
need to be controlled: charge generation, charge injection/ separation on different facets have been addressed.456−458 It has
transfer, and charge separation. Charge separation is especially been shown that the carrier recombination and inter- or
important for photocatalytic or photoelectrochemical applica- intraparticle charge migration depend strongly on the surface
tions involving TiO2 NPs matrices or photoelectrodes. structure459 and surface trap occurrence.460 To allow for
8.1. Photoinduced Charge Separation in TiO2 NPs effective charge separation in TiO2 NPs, the electron−hole
The mechanisms of the photoinduced charge generation, recombination needs to be prevented before the desired redox
separation, and surface reactions occurring in TiO2 NPs can be reaction can occur on its surface. As described above, the most
widely adopted way to promote spatial separation of electrons
represented as illustrated in Figure 37, and have in large been
and holes following photoexcitation is to modify the TiO2 NPs
adopted in this way by many researchers in the field. The first
by doping with metal or nonmetal ions, coupling with other
step (1) is the formation of photoinduced electron−hole pairs
semiconductors having appropriate band structure,171 or
by absorption of photons. When the energy of the incident light
depositing noble metal islands on the surface of titania
is larger than that of the band gap, electrons and holes are
particles.461,462 A common strategy to ensure charge separation
generated in the CB and VB, respectively. At n-type
on TiO2 NP surfaces is to bind electron-accepting or electron-
semiconductor−solution interfaces, upward band bending can
donating species.145 Using surface molecules as conductive
significantly depress the electron−hole pair recombination rate
leads, the proper ligand can result in fine-tuning of the
and enhance the migration rates of one charge carrier (electron
electronic properties of TiO2 NPs, which allows electronic
or hole) to the semiconductor surface.213,450 Recently,
coupling of the NPs into molecular circuits, providing further
Yamakata and co-worker142 proposed that the bent band drives
transport of photoinduced electrons originating from TiO2
free electrons and free holes to separate within picoseconds,
NCs.463,464 Thus, measuring and understanding the charge
while the carrier separation driven by the bent band no longer
carrier dynamics of semiconductors and TiO2 NPs is an
determines the decay kinetics at 50 ns or later when the
important aspect in evaluating the materials and their
electrons and holes are localized at traps (e.g., trapped hole
heterostructures for the development of energy conversion
centers (O−) and electron centers (Ti3+)). Charge trapping− systems.462
detrapping (by extrinsic excitation) is one of the plausible
models to describe the transport of charge carriers after 8.2. Charge Extraction for Photocatalytic Redox Reactions
photoexcitation. Undercoordinated Ti3+ and oxygen vacancies If efficient charge separation of the electron−hole pairs can lead
are often cited as likely sites for electrons to be trapped, while to migration of the charges to the particle surface, it will enable
surface hydroxyls have been implicated as sites for trapping the oxidation and reduction reactions, step 3 in Figure 37.
holes. Therefore, depending on the excitation lifetime relative Generally, within <400 fs, these primary charges separate and
to that of charge recombination, only a net fraction of the delocalize in the band tails of the semiconductor.465 Charge
photoinduced charges are available for chemistry, as illustrated separation across interfaces is also fast and occurs generally
in Figure 37. Because of such surface and volume within 1 ps.466−468
recombination processes (step 2 in Figure 37), only a small The driving force for the reduction and oxidation processes
fraction on the order of <10% of photogenerated electrons and on the surface of TiO2 is determined by the redox potentials of
holes can be utilized to initiate the redox processes at the both delocalized (conduction band electrons, ECB = −0.5 V,
interface (step 3 in Figure 37). Therefore, the charge separation and valence bands holes, EVB = +2.70 V versus NHE)238,469 and
efficiency in TiO2 NPs is central to the whole photoconversion surface-trapped charges relative to the redox potential of the
process. It is clear that this charge separation is very sensitive to molecules at the TiO2 surface. The above discussion highlights
various physicochemical properties, such as particle size and that, even when the photogenerated electrons and holes possess
shape, crystallinity, crystal phase, and so forth. sufficient thermodynamic potential for the target reactions, the
As implied in Figure 37, for small NPs, the distances that efficiency of a photocatalytic reaction is still strongly dependent
photogenerated charges have to migrate to reaction sites on the on the excitation conditions, and the dynamics of the created
NP surface become short, resulting in a decrease in the charge carriers in the photocatalytic cycle.470 Among the most
recombination probability. Even then, in the nanosize regime, significant applications of photocatalytic redox reactions on
because of the large curvature of TiO2 particle surfaces, the TiO2 NPs are the purification of water and air, solar water
atoms tend to reconstruct leading to distortions in the splitting, and CO2 reduction. They will be described in the
crystalline lattice of Ti atoms forming coordinatively unsatu- following sections along with the Grätzel-type solar cell
rated Ti cations at the surface. It has been reported that a application.
higher density (3 × 1012 cm−2) of surface states is located about 8.2.1. Charge Extraction for Photocatalytic Purifica-
0.5 V below the conduction band edge.453,454 The defects in tion of Water and Air. There are many remarkable advances
surface and volume act as trapping and recombination centers on the mineralization of toxic and nonbiodegradable organic
for photogenerated electrons and holes, affecting the photo- pollutants on titanium-oxide-based photocatalysts, both in gas
9306 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

and liquid media.451,471 A detailed mechanistic analysis of these Recent studies involving real-time single-molecule imaging
reactions has appeared in several reviews.98,472,473 In the show that hydroxyl and oxygen-containing radicals are not
photocatalytic purification of water and air on TiO2, molecular strictly retained at the photoirradiated surface but diffuse freely
oxygen (O2) usually acts as an electron acceptor and is through the gas/liquid-catalyst interface of the TiO2 photo-
converted to a superoxide anion (O2•−), which is a strong catalyst.477 For the electron transfer that occurs between the
oxidizing agent. In contrast, the hole tends to react with surface surface site and an adsorbed molecule, the driving force for the
hydroxyl groups (−OH) or adsorbed water molecules and heterogeneous electron transfer is the energy difference
produces hydroxyl radicals (•OH), another strong oxidizing between the conduction band of TiO2 and the reduction
agent. The reactions of these active oxygen radicals can potential of the acceptor redox couple.478 The photocatalytic
promote total decomposition of organic pollutants to CO2 and activity of TiO2 NPs depends on the specificity toward the
H2O under mild conditions (room temperature and atmos- targeted organic substrates. Each of the substrate’s functional
pheric pressure). However, in gas phase photocatalysis, group governs a dominant degradation mechanism. For
photocatalytic activity additionally depends on competitive example, carboxylic acids cause direct hole transfer; saccharides,
adsorption between water molecules and the target pollu- on the other hand, decompose via efficient reductive pathways,
tants.474,475 Thus, chemisorbed oxygen, either dissociated or while alcohols and aromatics lead to •OH-mediated oxidation
undissociated, may play a greater part in gas phase photo- on TiO2 surfaces.479 These surface chemistries can be improved
degradation than in aqueous phase photocatalysis. Electron and enhanced by further tracing and tuning the material’s
trapping by gas phase oxygen occurs in micro- to milli- intrinsic efficiencies for •OH generation and direct charge
seconds.476 Therefore, interfacial carrier transfer must be very transfer. Through investigating the TiO2-photoassisted one-
rapid in order to achieve photochemical efficiency. This electron redox reactions of organic compounds using various
requires the electron or hole trapping species to be preadsorbed steady-state and time-resolved spectroscopies, it was clearly
on the catalyst surface. Figure 38 shows a schematic view of the demonstrated that the adsorption dynamics of substrates and
time scale of the different charge-carrier-related phenomena. intermediates, the electronic interaction between TiO2 and
adsorbents, and the band structure and morphology of TiO2
nanomaterials are crucial factors for establishing efficient
photocatalytic systems.145,480,481
8.2.2. Charge Extraction for Solar Water Splitting. For
overall water splitting similar to electrolysis, water molecules
are reduced by electrons to form H2 and are oxidized by holes
to form O2. The thermodynamic requirement is that the band
edge potential of the conduction band has to be more negative
than the redox potential of H+/H2 (0 V vs NHE), while the
band edge potential of the valence band has to be more positive
than the redox potential of O2/H2O (1.23 V). Holes in the VB
of TiO2 are thermodynamically more favorable to react with
water than electrons in the CB based on the O2/H2O and H+/
H 2 redox potentials, respectively. In addition to the
thermodynamic aspects, kinetics plays a decisive role. The
oxidation and reduction steps in the water-splitting reactions
can be distinguished by their reaction times. Photogenerated
Figure 38. Time scales of “elemental steps” occurring in a typical holes can quickly oxidize water within 2 μs, whereas the
photocatalytic process.476 Reproduced with permission from ref 476. electrons reduce water in much longer time scales of 10−900
Copyright 2012 American Chemical Society. μs.142 The reduction step of the water splitting requires the

Figure 39. (A) TA spectra recorded 2 μs after 308 nm excimer laser pulse excitation of 23 mM TiO2 (5% acetic acid/95% ethanol) containing (a) 0,
(b) 0.02, (c) 0.04, and (d) 0.06 mM IrO2. The absorbance at 380 nm is assigned to trapped holes. (B) Normalized kinetics recorded at 380 nm
following the excitation of 23 mM TiO2 (a) without IrO2 and (b) with 0.06 mM IrO2.484 Reproduced with permission from ref 484. Copyright 2011
American Chemical Society.

9307 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

Figure 40. (A) TEM image of a single-layer graphene flake completely covered with ∼7 nm diameter TiO2 NPs. (B) Comparison of the TiO2
photocurrent afforded by the photocatalytic incorporation of RGO−TiO2 (blue) and TiO2 (magenta). Inset shows wavelength-dependent
photocurrent response of RGO−TiO2.488 Reproduced with permission from ref 488. Copyright 2011 American Chemical Society.

addition of Pt. The combination of appropriate loading, small 40). The UV-activated intrinsic charge generation (photo-
particle size of the metal islands, and the presence of surface current) of pure TiO2 was enhanced by a factor of 10 by
hydroxyl groups favors the photocatalytic activity of the incorporating RGO, which is ascribed to both the highly
materials.480 If not, the excess electrons accumulate in the conductive nature of the RGO and to improved charge
TiO2 and reduce Ti4+ cations. It is also possible that the collection facilitated by the intimate contact between RGO and
photogenerated holes may cause oxidation of the TiO2 surface, the TiO2.488
in which case TiO2 itself acts as an electron donor.482 The 8.2.3. Charge Extraction for Photocatalytic Reduction
electron-capture reaction exhibits a positive dependence on the of CO2. Currently, the science is emerging to convert the
water pressure, suggesting the promotion of physisorbed water greenhouse gas CO2 into organic molecules such as CH4,
species. Although the hole-capture reaction was shown to be CH3OH, formaldehyde, and formic acid by using solar energy
insensitive to the water pressure, the following O2 production and semiconductor nanostructures. Though currently chal-
may be the rate-determining step in water splitting.470 Thus, lenged by low yields, it represents a particularly rewarding
the photogenerated hole lifetime in TiO2 is a strong approach to the artificial photosynthesis of fuels. Under
determinant of the ability of TiO2 to split water. It has also heterogeneous gas−solid conditions, CH4 is the major product
been clarified that oxygen production requires four photons for of the photocatalytic reduction of CO2 with H2O using TiO2
each molecule of oxygen, which is reminiscent of the natural powders at room temperature.25 Special attention has been paid
photosynthetic water-splitting mechanism.470 To date, the most to the relationship between the local structure of the active Ti
effective photocatalysts for water splitting incorporate cocata- site and its photocatalytic performance.490,491 In situ spectro-
lysts including mainly NiO, RuO2, and IrO2,333,462,483 which scopic studies of the system indicate that the photocatalytic
shuttle the photogenerated holes across the photocatalyst reduction of CO2 with H2O is linked to a highly activated
interface and efficiently separate electrons from holes. With the charge-transfer excited state of the tetrahedrally coordinated
help of the spectral fingerprint of trapped holes with absorption Ti−oxide species, i.e., (Ti3+−O−)*, formed on the surface.25,492
at around 360 nm, one can monitor the transfer of trapped The reduction of CO2 to CH4 in aqueous solution is a
holes to IrO2. As shown in Figure 39, the absorbance at 380 nm multistep process, where the electrons are provided by
remains steady during the time scale of 15 μs in the absence of photoexcitation of TiO2 while water in the hydration layer at
IrO2. However, a rapid decay of the 380 nm absorbance is the surface acts as both a proton donor and an electron donor.
observed in the presence of IrO2. The lifetime of 1.5 μs Dissolved CO2 in the form of carbonates/bicarbonates can act
obtained from the first-order kinetics yields a rate constant for as hole scavengers. To detail the reaction mechanisms,
the hole transfer of 6 × 105 s−1.484 methoxyl (•OCH3) and methyl (•CH3) radicals were identified
In photoelectrochemical water-splitting systems, the interface as reaction intermediates with EPR spectroscopy.493 Dimi-
functionalities of TiO2 photoelectrodes, the flat band potential, trijevic et al.146 have indirectly identified the intermediate
the depletion layer, and the capacitance and impedance are all products formic acid, formaldehyde, and methanol as main hole
critical to achieve effective charge carrier migration.312,316,485 In scavengers by using time-resolved TA spectroscopy. In order to
addition, the modification with semiconductor QDs can enlarge analyze the experimental kinetics obtained in the presence of
both the depletion layer capacitance and the double-layer hole scavengers, the 2−2500 ps kinetic curves obtained at 520
capacitance and reduce their charge-transfer resistance and nm detection wavelength were fitted using a triexponential
electrochemical impedance. The wider depletion layers also function (Figure 41). The ultrafast dynamics of hole scavenging
produce longer lifetimes of the charge carriers.483 Very recently, was found to be an order of magnitude faster on the surface of
it was proven that photoinduced charge extraction can be TiO2 than in the corresponding homogeneous systems with
accelerated by assembly with single and multiwall carbon hydroxyl radicals. They inferred that the efficiency of the CO2
nanotubes486 and graphene oxide,487,488 which can provide a reduction is independent of the hole scavenger, and the small
convenient platform to anchor TiO2 NPs and metal nano- driving force of the electron transfer is one of the main factors
particles (e.g., Pt). When TiO2 NPs on a single graphene or contributing to the overall yield of the initial CO2 reduction.
reduced graphene oxide (RGO) sheet absorbs the light and There is currently much work in progress with regard to the
induces the oxidation reaction, the reduced graphene oxide can organic intermediates, which are oxidized by photoinduced
capture electrons and shuttle them across the 2-D carbon holes or •OH radicals and the one-electron or two/multi-
network (the π−π network) to the Pt site to facilitate hydrogen electron-reduction mechanism analysis.462,494 Such photo-
reduction.489 RGO−TiO2 composite films exhibit electron catalytic reduction of CO2 still needs to be carefully and
lifetimes up to 4 times longer than that of pure TiO2 (Figure systematically scrutinized.
9308 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318
Chemical Reviews Review

transfer from the dye to the semiconductor usually takes place


on the femto- to picosecond time scale.498 Multiexponential
kinetics of electron injection was observed for the electron
injection process in dye-sensitized TiO2 nanocrystalline
systems.499 The observed relaxation rates of dye-sensitized
solar cell electrodes are thus extremely fast. However, fast
relaxation dynamics does not necessarily correlate with a high
efficiency of electron injection. A high degree of dye
aggregation or inactive, loose dyes on the surface can also
lead to accelerated electronic relaxation kinetics while bypassing
the injection process. As studied by TRPL and femtosecond TA
spectroscopy (Figure 42), small NPs produced high electron
injection yields of ∼90%, whereas larger diameter systems gave
lower efficiencies from 35% to 70%.500 Two possible
mechanisms, dye aggregation on N3−TiO2(N3 dye: (Ru-
Figure 41. Normalized kinetic traces at 520 nm transient absorption (dcbpy)2(NCS)2)) and localized surface states, can explain the
for aqueous TiO2 dispersions in the absence and the presence of 1 M difference in the obtained electron injection yields for different
methanol, formaldehyde, and glycerol.146 Reproduced with permission size sensitizers. Other competing relaxation processes can also
from ref 146. Copyright 2012 American Chemical Society. affect the electron injection yields. The free energy change ΔG
for electron injection is one of the most important factors to
8.3. Charge Separation and Injection in Grätzel-type Solar determine the electron-transfer process,143 which in turn can be
Cells affected by different crystal facets as well as by adsorbates on
the TiO2 surface.501
In TiO2 NPs-based (Grätzel-type) DSSCs, charge separation
Charge injection dynamics are substantially influenced by the
occurs ultimately by electron injection from the photoexcited
local environment of the sensitizer−TiO2 NP interface that is
sensitizer dye into ground state TiO2 NCs. This is enabled by
governed by sensitizer binding modes and the density of states
an energy level offset between the sensitizer dye and TiO2 NCs
that partially make up the photoanode material.495 The available for injection.502 To achieve an efficient photoelectric
sensitizer, which is often a transition metal complex but conversion, the injection rate should be much higher than those
organic pigments and QDs have been successfully employed, of all competing processes from the sensitizer excited state.
absorbs a photon of visible light to produce an excited state that That is, a high photon-to-current conversion efficiency of
then injects an electron into the conduction band of a wide DSSCs requires fast-forward electron injection from the excited
band gap semiconductor film, usually composed of TiO2 NPs. state of the adsorbed sensitizer to the TiO2 NPs as well as slow
Another pathway discussed for electron injection is the back electron transfer from TiO2 NPs to the adsorbates. Figure
possibility of direct hot electron injection into the CB of the 43a illustrates how the charge recombination reaction following
TiO2 NPs, leading to instantaneous charge separation without photoinduced electron injection may be rate-limited by either
the formation of any low-lying excited states of the dye by way electron transport dynamics within the TiO2 NPs or by the
of electron relaxation.496 However, most dye molecules inject electron-transfer rate at the sensitizer/TiO2 interface.503 This
electrons following the first pathway (sensitization, excited state localized distribution model of injected electrons is based on
relaxation, and subsequent electron transfer).143 The design of the experimental observation of size-independent charge
new sensitizers which combine strong visible light absorption recombination kinetics,504 in which the injected electron is
with favorable electron-transfer dynamics is a key issue in the assumed to be randomly distributed within a small region near
development and optimization of DSSCs.497 The electron the parent adsorbate. In addition, a continuous-time random

Figure 42. (a) Fluorescence decay kinetics of different N3−TiO2 systems, N3 dye(Ru(dcbpy)2(NCS)2) solution, and N3 powder, all excited at 500
nm. (b) Electron injection yield of all the samples as a function of TiO2 diameter.500 Reproduced with permission from ref 500. Copyright 2010
American Chemical Society.

9309 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

different size TiO2 NPs have been developed, and their


properties have been extensively investigated. Currently the
nanoscience of TiO2 has matured to a degree that nanoma-
terials of targeted sizes and composition with designed
interfaces can be prepared for specific properties, unparalleled
to any other nanomaterial. Most recent publications indicate
that titania continues to be a versatile material with wide
structural and functional variations. TiO2−graphene nano-
composites with tunable TiO2 crystal facets are synthesized and
achieved excellent photocatalytic H2 evolution rate, confirming
Figure 43. Schematic diagrams illustrating electron injection, that the charge separation efficiency and band structures of
relaxation diffusion, and spatial trapping. (a) Model of charge resulting nanocomposites are greatly dependent on the crystal
recombination for the trapped electron in a sensitized semiconductor facets.509 In order to develop high surface area nanostructured
nanoparticle, (b) surface electron trapping by random flight model electrodes with rapid charge transport, the single-crystal-like 3-
(shaded area refers to the lattice), and (c) surface electron trapping by D TiO2 branched nanowire arrays were assembled and
the localized distribution model. Abbreviations follow: FL, Fermi level; exhibited 52% improvement in solar conversion efficiency for
CB, conduction band; VB, valence band; S*/S+, reduction potential of solar cells.510 Constructing 3-D branched TiO2 nanowire
the excited sensitizer; S/S+, reduction potential of the ground-state coated macroporous metal oxide electrodes can enhance light
sensitizer. κ is defined as Kt/(4πDR), where Kt is a second-order rate
harvesting and charge collection, and reduce interfacial
constant of the electron trapping, D is the diffusion constant of the
electron, and R is the radius of the particle.507 Reproduced with recombination, leading to significantly increase the photo-
permission from ref 507. Copyright 2007 American Chemical Society. current and photovoltage of DSSCs.511 Obviously, the themes
for the coming years will be increasing complexity and
formation of bulk materials based on nanoscale building blocks
walk503,505 and a random flight model143,506 of the injected with specific target properties in mind. For these reasons alone,
electron were proposed for sensitized TiO2 NCs (Figure 43). TiO2 will further cement its characteristic as the most
The random flight model described a homogeneous trapping functional metal oxide (if not overall inorganic) material
of the injected electron at the surface. Recently, a proposed being investigated. For example, recently a revolutionary low-
electron diffusion-and-trapping mechanism further supported cost high efficiency solar cell, sensitized with perovskites, has
the physical mechanism of a localized charge distribution achieved up to 15% efficiencies, as reported by the Grätzel and
model.143 In general, it is common to all descriptions of Snaith groups.512,513 High-quality nanoparticulate TiO2 films
electron transport in nanoparticle films that electron diffusion still play a key role in solar cells, similarly as they do in
processes depend on the energy of the injected electron. In DSSCs.514−516 It seems likely that TiO2 nanostructures will
order to reveal the kinetics and mechanism of electron remain among the most cost-efficient and useful materials that
injection, several types of transient absorption spectrometers the field of nanoscience has produced.
with high sensitivity have been developed to observe the time
profiles for the absorption spectra of the oxidized form of dyes AUTHOR INFORMATION
and conducting electrons over a wide time range from Corresponding Author
femtoseconds to sub-microseconds. Thus, the efficiency of *E-mail: burda@case.edu. Phone: 216-368-5918. Fax: 216-368-
electron injection can be measured and explained from the 3006.
consideration of experiment and theory. Notes
To sum up, there are many investigations on TiO2
nanoparticles as charge separation centers combined with The authors declare no competing financial interest.
their applications of photocatalytic redox reactions (including Biographies
the purification of water and air, solar water splitting, and CO2
reduction) and DSSCs. Usually, semiconductor photocatalysis
has been considered as a rather nonselective process, especially
in aqueous media. Despite the low selectivity of metal oxides in
water, the selective oxidation of alcohols and amines on TiO2
under UV irradiation has been successfully achieved. Moreover,
an increasing number of researchers have investigated the
photocatalytic organic transformations by TiO2 NPs.142
Published in 2014, Lang et al.508 have reviewed the significant
progress in this area of research. The authors also note that the
redox transformations of organic compounds by heterogeneous
photocatalysis should be investigated to understand the
photoinduced interfacial electron-transfer processes in more
detail. This may have a far-reaching impact on the development
of environmentally friendly and energy sustainable strategies for
selective organic synthesis. Professor Lixia Sang received her Ph.D. in 2004 in Industrial Catalysis
from Tianjin University in China with Professor Shunhe Zhong,
9. OUTLOOK studying photostimulated surface catalysis. Since then, she has worked
TiO2 NP-based materials are today well-defined and well- at Beijing University of Technology and Key Laboratory of Enhanced
understood. In the past decade, different approaches to prepare Heat Transfer and Energy Conservation, Ministry of Education of

9310 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

China. In 2012, she spent one year studying as a Visiting Scholar at REFERENCES
Case Western Reserve University under the directorship of Professor (1) Fujishima, A.; Honda, K. Nature 1972, 238, 37.
Burda. Her current research is focused on solar energy conversion (2) Tributsch, H. Photochem. Photobiol. 1972, 16, 261.
materials with specific interest in charge-transfer mechanism for solar (3) Varghese, O. K.; Gong, D. W.; Paulose, M.; Ong, K. G.; Grimes,
fuels. C. A. Sens. Actuators, B 2003, 93, 338.
(4) Zhu, Y. F.; Shi, J. J.; Zhang, Z. Y.; Zhang, C.; Zhang, X. R. Anal.
Chem. 2002, 74, 120.
(5) Heinlaan, M.; Ivask, A.; Blinova, I.; Dubourguier, H. C.; Kahru, A.
Chemosphere 2008, 71, 1308.
(6) Cai, R. X.; Kubota, Y.; Shuin, T.; Sakai, H.; Hashimoto, K.;
Fujishima, A. Cancer Res. 1992, 52, 2346.
(7) Song, Y. Y.; Schmidt-Stein, F.; Bauer, S.; Schmuki, P. J. Am.
Chem. Soc. 2009, 131, 4230.
(8) Shrestha, N. K.; Macak, J. M.; Schmidt-Stein, F.; Hahn, R.;
Mierke, C. T.; Fabry, B.; Schmuki, P. Angew. Chem., Int. Ed. 2009, 48,
969.
(9) Majewski, L. A.; Schroeder, R.; Grell, M. Adv. Funct. Mater. 2005,
15, 1017.
(10) Maliakal, A.; Katz, H.; Cotts, P. M.; Subramoney, S.; Mirau, P. J.
Am. Chem. Soc. 2005, 127, 14655.
(11) Sundstrom, V. Prog. Quantum Electron. 2000, 24, 187.
Professor Yixin Zhao obtained his Ph.D. from Case Western Reserve (12) Albaret, T.; Finocchi, F.; Noguera, C. Faraday Discuss. 1999,
114, 285.
University in 2010 with Professor Burda as his advisor. After receiving
(13) Hagfeldt, A.; Lunell, S.; Siegbahn, H. O. G. Int. J. Quantum
his Ph.D., he worked as postdoctoral fellow at Pennsylvania State Chem. 1994, 49, 97.
University and National Renewable Energy Laboratory on solar energy (14) Liu, X. H.; Zhang, X. G.; Li, Y.; Wang, X. Y.; Lou, N. Q. Int. J.
conversion including photovoltaics and water splitting. Currently, he Mass Spectrom. 1998, 177, L1.
holds a faculty position at Shanghai Jiao Tong University, and his (15) Velegrakis, M.; Sfounis, A. Appl. Phys. A: Mater. Sci. Process.
research interest is in solar energy and nanomaterials for environ- 2009, 97, 765.
mental remediation. (16) Tsipis, A. C.; Tsipis, C. A. Phys. Chem. Chem. Phys. 1999, 1,
4453.
(17) Matsuda, Y.; Bernstein, E. R. J. Phys. Chem. A 2005, 109, 314.
(18) Yu, W.; Freas, R. B. J. Am. Chem. Soc. 1990, 112, 7126.
(19) Shevlin, S. A.; Woodley, S. M. J. Phys. Chem. C 2010, 114,
17333.
(20) Auvinen, S.; Alatalo, M.; Haario, H.; Jalava, J.-P.; Lamminmäki,
R.-J. J. Phys. Chem.C 2011, 115, 8484.
(21) Brus, L. E. J. Chem. Phys. 1984, 80, 4403.
(22) Lundqvist, M. J.; Nilsing, M.; Persson, P.; Lunell, S. Int. J.
Quantum Chem. 2006, 106, 3214.
(23) Li, S. G.; Dixon, D. A. J. Phys. Chem. A 2008, 112, 6646.
(24) Zhang, H. Z.; Banfield, J. F. J. Phys. Chem. B 2000, 104, 3481.
(25) Mori, K.; Yamashita, H.; Anpo, M. RSC Adv. 2012, 2, 3165.
(26) Qu, Z. W.; Kroes, G. J. J. Phys. Chem. C 2007, 111, 16808.
(27) Zhang, D. J.; Sun, H.; Liu, J. Q.; Liut, C. B. J. Phys. Chem. C
2009, 113, 21.
(28) Zhai, H. J.; Wang, L. S. J. Am. Chem. Soc. 2007, 129, 3022.
Professor Clemens Burda is a director of the Center for Chemical (29) Dambournet, D.; Belharouak, I.; Amine, K. Chem. Mater. 2010,
Dynamics and Nanomaterials Research at Case Western Reserve 22, 1173.
University. His research focuses on energy and biomedical applications (30) Nosheen, S.; Galasso, F. S.; Suib, S. L. Langmuir 2009, 25, 7623.
of nanomaterials. He obtained his Ph.D. in 1997 in Physical Chemistry (31) Diebold, U. Surf. Sci. Rep. 2003, 48, 53.
from the University of Basel in Switzerland with Professor Jakob Wirz, (32) Lu, H. M.; Zhang, W. X.; Jiang, Q. Adv. Eng. Mater. 2003, 5, 787.
studying laser-induced photochemical processes. He pursued post- (33) Kumar, K. N. P.; Keizer, K.; Burggraaf, A. J.; Okubo, T.;
doctoral research with Professor El-Sayed at Georgia Institute of Nagamoto, H.; Morooka, S. Nature 1992, 358, 48.
Technology where he was exposed to nanoscience and nano- (34) Machado, N. R. C. F.; Santana, V. S. Catal. Today 2005, 107−
technology. Since 2001 Professor Burda has been a faculty member 08, 595.
(35) Koparde, V. N.; Cummings, P. T. ACS Nano 2008, 2, 1620.
at Case Western Reserve University and Full Professor since 2011. He
(36) Sanz, J.; Soria, J.; Sobrados, I.; Yurdakal, S.; Augugliaro, V. J.
is primarily interested in the photophysical and photochemical Phys. Chem. C 2012, 116, 5110.
processes of nanoscale materials. Professor Burda is on the editorial (37) Finnegan, M. P.; Zhang, H. Z.; Banfield, J. F. J. Phys. Chem. C
board for several energy- and nanotechnology-related journals and an 2007, 111, 1962.
author and coauthor of over 150 publications in the areas of (38) Machon, D.; Daniel, M.; Bouvier, P.; Daniele, S.; Le Floch, S.;
nanotechnology and nanomedicine. Melinon, P.; Pischedda, V. J. Phys. Chem. C 2011, 115, 22286.
(39) Hummer, D. R.; Kubicki, J. D.; Kent, P. R. C.; Post, J. E.;
ACKNOWLEDGMENTS Heaney, P. J. J. Phys. Chem. C 2009, 113, 4240.
(40) Fernandez-Garcia, M.; Belver, C.; Hanson, J. C.; Wang, X.;
The authors gratefully acknowledge the National Natural Rodriguez, J. A. J. Am. Chem. Soc. 2007, 129, 13604.
Science Foundation of China (Grants 51376013 and (41) Zhang, J.; Li, M. J.; Feng, Z. C.; Chen, J.; Li, C. J. Phys. Chem. B
51372151) for financial support. 2006, 110, 927.

9311 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(42) Zhang, J.; Xu, Q.; Li, M. J.; Feng, Z. C.; Li, C. J. Phys. Chem. C (79) Wang, W.; Gu, B.; Liang, L.; Hamilton, W. A.; Wesolowski, D. J.
2009, 113, 1698. J. Phys. Chem. B 2004, 108, 14789.
(43) Zou, C. W.; Yan, X. D.; Chen, R. Q.; Alyamanl, A.; Wu, Z. Y.; (80) Niu, A. P.; Li, Y.; Chen, J. F.; Xu, Q.; Cao, Y. X. Ind. Eng. Chem.
Gao, W. Cryst. Growth Des. 2011, 11, 367. Res. 2009, 48, 7103.
(44) Yuan, M. Q.; Zhang, J.; Yan, S.; Luo, G. X.; Xu, Q.; Wang, X.; Li, (81) Zheng, W. J.; Liu, X. D.; Yan, Z. Y.; Zhu, L. J. ACS Nano 2009,
C. J. Alloys Compd. 2011, 509, 6227. 3, 115.
(45) Zhao, J. S.; Wang, Y. P.; Lou, X. R.; Li, K.; Li, Z.; Huang, W. (82) Cihlar, J.; Bartonickova, E. J. Sol-Gel Sci. Technol. 2013, 65, 430.
Inorg. Chim. Acta 2013, 405, 395. (83) Durupthy, O.; Bill, J.; Aldinger, F. Cryst. Growth Des. 2007, 7,
(46) Shi, H. X.; Zhang, T. Y.; An, T. C.; Li, B.; Wang, X. J. Colloid 2696.
Interface Sci. 2012, 380, 121. (84) Puddu, V.; Slocik, J. M.; Naik, R. R.; Perry, C. C. Langmuir
(47) Xiao, J. R.; Peng, T. Y.; Li, R.; Peng, Z. H.; Yan, C. H. J. Solid 2013, 29, 9464.
State Chim. 2006, 179, 1161. (85) Kharlampieva, E.; Slocik, J. M.; Singamaneni, S.; Poulsen, N.;
(48) Yang, J.; Dai, J.; Li, J. T. Appl. Surf. Sci. 2011, 257, 8965. Kröger, N.; Naik, R. R.; Tsukruk, V. V. Adv. Funct. Mater. 2009, 19,
(49) Sang, L. X.; Gole, J. L.; Wang, J. W.; Brauer, J.; Mao, B. D.; 2303.
Prokes, S. M.; Burda, C. J. Phys. Chem.C 2013, 117, 15287. (86) Kharlampieva, E.; Tsukruk, T.; Slocik, J. M.; Ko, H.; Poulsen,
(50) Luan, Y.; Jing, L.; Meng, Q.; Nan, H.; Luan, P.; Xie, M.; Feng, Y. N.; Naik, R. R.; Kröger, N.; Tsukruk, V. V. Adv. Mater. 2008, 20, 3274.
J. Phys. Chem. C 2012, 116, 17094. (87) Kim, H.; Pippel, E.; Gösele, U.; Knez, M. Langmuir 2009, 25,
(51) Shen, X. J.; Tian, B. Z.; Zhang, J. L. Catal. Today 2013, 201, 151. 13284.
(52) Zhang, L. J.; Menendez-Flores, V. M.; Murakami, N.; Ohno, T. (88) Ding, K. L.; Miao, Z. J.; Liu, Z. M.; Zhang, Z. F.; Han, B. X.; An,
Appl. Surf. Sci. 2012, 258, 5803. G. M.; Miao, S. D.; Xie, Y. J. Am. Chem. Soc. 2007, 129, 6362.
(53) Deng, Q. X.; Wei, M. D.; Ding, X. K.; Jiang, L. L.; Ye, B. H.; (89) Garnweitner, G.; Antonietti, M.; Niederberger, M. Chem.
Wei, K. M. Chem. Commun. 2008, 3657. Commun. 2005, 397.
(54) Magne, C.; Cassaignon, S.; Lancel, G.; Pauporte, T. (90) Jensen, G. V.; Bremholm, M.; Lock, N.; Deen, G. R.; Jensen, T.
ChemPhysChem. 2011, 12, 2461. R.; Iversen, B. B.; Niederberger, M.; Pedersen, J. S.; Birkedal, H. Chem.
(55) Lin, C. H.; Chao, J. H.; Liu, C. H.; Chang, J. C.; Wang, F. C. Mater. 2010, 22, 6044.
Langmuir 2008, 24, 9907. (91) Kotsokechagia, T.; Cellesi, F.; Thomas, A.; Niederberger, M.;
(56) Perego, C.; Wang, Y. H.; Durupthy, O.; Cassaignon, S.; Revel, Tirelli, N. Langmuir 2008, 24, 6988.
R.; Jolivet, J. P. ACS Appl. Mater. Interfaces 2012, 4, 752. (92) Niederberger, M.; Garnweitner, G. Chem.Eur. J. 2006, 12,
(57) Buonsanti, R.; Grillo, V.; Carlino, E.; Giannini, C.; Kipp, T.; 7282.
(93) Wu, B.; Guo, C.; Zheng, N.; Xie, Z.; Stucky, G. D. J. Am. Chem.
Cingolani, R.; Cozzoli, P. D. J. Am. Chem. Soc. 2008, 130, 11223.
(58) Zhao, B.; Chen, F.; Huang, Q. W.; Zhang, J. L. Chem. Commun. Soc. 2008, 130, 17563.
(94) Zimmermann, M.; Garnweitner, G. CrystEngComm 2012, 14,
2009, 34, 5115.
8562.
(59) Hu, W. B.; Li, L. P.; Li, G. S.; Tang, C. L.; Sun, L. Cryst. Growth
(95) Koziej, D.; Fischer, F.; Kränzlin, N.; Caseri, W. R.; Niederberger,
Des. 2009, 9, 3676.
M. ACS Appl. Mater. Interfaces 2009, 1, 1097.
(60) Chen, X.; Mao, S. S. Chem. Rev. 2007, 107, 2891.
(96) Dinh, C. T.; Nguyen, T. D.; Kleitz, F.; Do, T. O. ACS Nano
(61) Li, G. H.; Gray, K. A. Chem. Mater. 2007, 19, 1143.
2009, 3, 3737.
(62) Serpone, N.; Lawless, D.; Khairutdinov, R. J. Phys. Chem. 1995,
(97) Ahmed, A. Y.; Kandiel, T. A.; Oekermann, T.; Bahnemann, D. J.
99, 16646.
Phys. Chem. Lett. 2011, 2, 2461.
(63) Ichijo, T.; Sato, S.; Fujita, M. J. Am. Chem. Soc. 2013, 135, 6786.
(98) Fujishima, A.; Zhang, X.; Tryk, D. Surf. Sci. Rep. 2008, 63, 515.
(64) Satoh, N.; Nakashima, T.; Kamikura, K.; Yamamoto, K. Nat.
(99) Thompson, T. L.; Yates, J. T. Chem. Rev. 2006, 106, 4428.
Nanotechnol. 2008, 3, 106. (100) Kubo, T.; Orita, H.; Nozoye, H. J. Am. Chem. Soc. 2007, 129,
(65) Burunkaya, E.; Akarsu, M.; Camurlu, H. E.; Kesmez, O.; Yesil, 10474.
Z.; Asilturk, M.; Arpac, E. Appl. Surf. Sci. 2013, 265, 317. (101) Wang, H.; Shao, W.; Gu, F.; Zhang, L.; Lu, M.; Li, C. Inorg.
(66) Athar, T.; Han, K.; Han, S.; Ko, T.; Lee, I. M. Synth. React. Chem. 2009, 48, 9732.
Inorg., Met.-Org., Nano-Met. Chem. 2009, 39, 261. (102) Yang, H. G.; Sun, C. H.; Qiao, S. Z.; Zou, J.; Liu, G.; Smith, S.
(67) Cihlar, J.; Bartonickova, E. J. Sol-Gel Sci. Technol. 2013, 65, 430. C.; Cheng, H. M.; Lu, G. Q. Nature 2008, 453, 638.
(68) Abbas, Z.; Holmberg, J. P.; Hellstrom, A. K.; Hagstrom, M.; (103) Dai, Y. Q.; Cobley, C. M.; Zeng, J.; Sun, Y. M.; Xia, Y. N. Nano
Bergenholtz, J.; Hassellov, M.; Ahlberg, E. Colloids Surf., A 2011, 384, Lett. 2009, 9, 2455.
254. (104) Nguyen, C. K.; Cha, H. G.; Kang, Y. S. Cryst. Growth Des.
(69) Chuang, H. Y.; Chen, D. H. J. Nanopart. Res. 2008, 10, 233. 2011, 11, 3947.
(70) Dhage, S. R.; Choube, V. D.; Samuel, V.; Ravi, V. Mater. Lett. (105) Gordon, T. R.; Cargnello, M.; Paik, T.; Mangolini, F.; Weber,
2004, 58, 2310. R. T.; Fornasiero, P.; Murray, C. B. J. Am. Chem. Soc. 2012, 134, 6751.
(71) Hart, J. N.; Bourgeois, L.; Cervini, R.; Cheng, Y. B.; Simon, G. (106) Chen, C.; Hu, R.; Mai, K.; Ren, Z.; Wang, H.; Qian, G.; Wang,
P.; Spiccia, L. J. Sol-Gel Sci. Technol. 2007, 42, 107. Z. Cryst. Growth Des. 2011, 11, 5221.
(72) Li, G. S.; Li, L. P.; Boerio-Goates, J.; Woodfield, B. F. J. Am. (107) Zheng, Y.; Lv, K. L.; Wang, Z. Y.; Deng, K. J.; Li, M. J. Mol.
Chem. Soc. 2005, 127, 8659. Catal. A: Chem. 2012, 356, 137.
(73) Pottier, A.; Cassaignon, S.; Chaneac, C.; Villain, F.; Tronc, E.; (108) Yang, H. G.; Liu, G.; Qiao, S. Z.; Sun, C. H.; Jin, Y. G.; Smith,
Jolivet, J.-P. J. Mater. Chem. 2003, 13, 877. S. C.; Zou, J.; Cheng, H. M.; Lu, G. Q. J. Am. Chem. Soc. 2009, 131,
(74) Li, S. F.; Ye, G. L.; Chen, G. Q. J. Phys. Chem. C 2009, 113, 4078.
4031. (109) Han, X. G.; Kuang, Q.; Jin, M. S.; Xie, Z. X.; Zheng, L. S. J. Am.
(75) Wang, P.; Xie, T. F.; Peng, L. A.; Li, H. Y.; Wu, T. S.; Pang, S.; Chem. Soc. 2009, 131, 3152.
Wang, D. J. J. Phys. Chem. C 2008, 112, 6648. (110) Kobayashi, M.; Petrykin, V. V.; Kakihana, M. Chem. Mater.
(76) Szeifert, J. M.; Feckl, J. M.; Fattakhova-Rohlfing, D.; Liu, Y. J.; 2007, 19, 5373.
Kalousek, V.; Rathousky, J.; Bein, T. J. Am. Chem. Soc. 2010, 132, (111) Gong, X. Q.; Selloni, A. Phys. Rev. B 2007, 76, 235307.
12605. (112) Zhao, B.; Chen, F.; Huang, Q. W.; Zhang, J. L. Chem. Commun.
(77) Burunkaya, E.; Akarsu, M.; Camurlu, H. E.; Kesmez, O.; Yesil, 2009, 7, 5115.
Z.; Asilturk, M.; Arpac, E. Appl. Surf. Sci. 2013, 265, 317. (113) Kandiel, T. A.; Feldhoff, A.; Robben, L.; Dillert, R.;
(78) Zhao, Y.; Ren, W.; Cui, H. J. Colloid Interface Sci. 2013, 398, 7. Bahnemann, D. W. Chem. Mater. 2010, 22, 2050.

9312 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(114) Shen, X. J.; Tian, B. Z.; Zhang, J. L. Catal. Today 2013, 201, (149) Zhao, H.; Zhang, Q. L.; Weng, Y. X. J. Phys. Chem. C 2007,
151. 111, 3762.
(115) Liu, L. C.; Gu, X. R.; Ji, Z. Y.; Zou, W. X.; Tang, C. J.; Gao, F.; (150) Guo, L. J.; Wang, Y. M.; Lu, H. P. J. Am. Chem. Soc. 2010, 132,
Dong, L. J. Phys. Chem. C 2013, 117, 18578. 1999.
(116) Fernández-García, M.; Martínez-Arias, A.; Hanson, J. C.; (151) Du, L. C.; Furube, A.; Yamamoto, K.; Hara, K.; Katoh, R.;
Rodriguez, J. A. Chem. Rev. 2004, 104, 4063. Tachiya, M. J. Phys. Chem. C 2009, 113, 6454.
(117) Li Bassi, A.; Cattaneo, D.; Russo, V.; Bottani, C. E.; Barborini, (152) Ardo, S.; Meyer, G. J. J. Am. Chem. Soc. 2010, 132, 9283.
E.; Mazza, T.; Piseri, P.; Milani, P.; Ernst, F. O.; Wegner, K.; Pratsinis, (153) Yamanaka, K.; Morikawa, T. J. Phys. Chem. C 2012, 116, 1286.
S. E. J. Appl. Phys. 2005, 98, 074305. (154) Tojo, S.; Tachikawa, T.; Fujitsuka, M.; Majima, T. J. Phys.
(118) Zhang, M. S.; Zhang, W. F.; Yin, Z. P. Soc. Photo-Opt. Instrum. Chem. C 2008, 112, 14948.
Eng. 2001, 4469, 93. (155) Tachikawa, T.; Takai, Y.; Tojo, S.; Fujitsuka, M.; Irie, H.;
(119) Xue, X. X.; Ji, W.; Mao, Z.; Mao, H. J.; Wang, Y.; Wang, X.; Hashimoto, K.; Majima, T. J. Phys. Chem. B 2006, 110, 13158.
Ruan, W. D.; Zhao, B.; Lombardi, J. R. J. Phys. Chem. C 2012, 116, (156) Tachikawa, T.; Yoshida, A.; Tojo, S.; Sugimoto, A.; Fujitsuka,
8792. M.; Majima, T. Chem.Eur. J. 2004, 10, 5345.
(120) Tompsett, G. A.; Bowmaker, G. A.; Cooney, R. P.; Metson, J. (157) Salafsky, J. S.; Lubberhuizen, W. H.; van Faassen, E.; Schropp,
B.; Rodgers, K. A.; Seakins, J. M. J. Raman. Spectrosc. 1995, 26, 57. R. E. I. J. Phys. Chem. B 1998, 102, 766.
(121) Parker, J. C.; Siegel, R. W. Appl. Phys. Lett. 1990, 57, 943. (158) Savenije, T. J.; Vermeulen, M. J. W.; de Haas, M. P.; Warman,
(122) Tang, H.; Prasad, K.; Sanjines, R.; Schmid, P. E.; Levy, F. J. J. M. Sol. Energy Mater. Sol. Cells 2000, 61, 9.
Appl. Phys. 1994, 75, 2042. (159) Kroeze, J. E.; Savenije, T. J.; Vermeulen, M. J. W.; Warman, J.
(123) Hu, Z.; Metiu, H. J. Phys. Chem. C 2011, 115, 5841. M. J. Phys. Chem. B 2003, 107, 7696.
(124) Jedidi, A.; Markovits, A.; Minot, C.; Bouzriba, S.; Abderraba, (160) Katoh, R.; Huijser, A.; Hara, K.; Savenije, T. J.; Siebbeles, L. D.
M. Langmuir 2010, 26, 16232. A. J. Phys. Chem. C 2007, 111, 10741.
(125) Sang, L. X.; Zhong, S. H.; Ma, C. F. Spectrosc. Spect. Anal. 2007, (161) Kroeze, J. E.; Savenije, T. J.; Warman, J. M. J. Am. Chem. Soc.
27, 720. 2004, 126, 7608.
(126) Zhao, Y.; Li, C. Z.; Liu, X. H.; Gu, F.; Jiang, H. B.; Shao, W.; (162) Salafsky, J. S. Phys. Rev. B 1999, 59, 10885.
Zhang, L.; He, Y. Mater. Lett. 2007, 61, 79. (163) Kowalska, E.; Remita, H.; Colbeau-Justin, C.; Hupka, J.;
(127) Zallen, R.; Moret, M. P. Solid. State. Commun. 2006, 137, 154. Belloni, J. J. Phys. Chem. C 2008, 112, 1124.
(128) Kuznetsov, V. N.; Serpone, N. J. Phys. Chem. C 2009, 113, (164) Boujday, S.; Wunsch, F.; Portes, P.; Bocquet, J. F.; Colbeau-
15110. Justin, C. Sol. Energy Mater. Sol. Cells 2004, 83, 421.
(129) Chen, X.; Glans, P. A.; Qiu, X.; Dayal, S.; Jennings, W. D.; (165) Martin, S. T.; Herrmann, H.; Choi, W. Y.; Hoffmann, M. R. J.
Smith, K. E.; Guo, J. J. Electron Spectrosc. 2008, 162, 67. Chem. Soc., Faraday Trans. 1994, 90, 3315.
(130) Chen, X. B.; Burda, C. J. Phys. Chem. B 2004, 108, 15446. (166) Martin, S. T.; Herrmann, H.; Hoffmann, M. R. J. Chem. Soc.,
(131) Gole, J. L.; Stout, J. D.; Burda, C.; Lou, Y. B.; Chen, X. B. J. Faraday Trans. 1994, 90, 3323.
Phys. Chem. B 2004, 108, 1230. (167) Mele, G.; Del Sole, R.; Vasapollo, G.; Marci, G.; Garcia-Lopez,
(132) Gole, J. L.; Prokes, S. M.; Glembocki, O. J.; Wang, J. W.; Qiu, E.; Palmisano, L.; Coronado, J. M.; Hernandez-Alonso, M. D.;
X. F.; Burda, C. Nanoscale 2010, 2, 1134. Malitesta, C.; Guascito, M. R. J. Phys. Chem. B 2005, 109, 12347.
(133) Macdonald, I. R.; Rhydderch, S.; Holt, E.; Grant, N.; Storey, J. (168) Chuang, C. H.; Burda, C. J. Phys. Chem. Lett. 2012, 3, 1921.
M. D.; Howe, R. F. Catal. Today 2012, 182, 39. (169) Varaganti, S.; Ramakrishna, G. J. Phys. Chem. C 2010, 114,
(134) Gopal, N. O.; Lo, H. H.; Ke, T. F.; Lee, C. H.; Chou, C. C.; 13917.
Wu, J. D.; Sheu, S. C.; Ke, S. C. J. Phys. Chem. C 2012, 116, 16191. (170) McNeil, I. J.; Ashford, D. L.; Luo, H. L.; Fecko, C. J. J. Phys.
(135) Berger, T.; Sterrer, M.; Diwald, O.; Knozinger, E.; Panayotov, Chem. C 2012, 116, 15888.
D.; Thompson, T. L.; Yates, J. T. J. Phys. Chem. B 2005, 109, 6061. (171) Chen, Y. C.; Pu, Y. C.; Hsu, Y. J. J. Phys. Chem. C 2012, 116,
(136) Panayotov, D. A.; Burrows, S. P.; Morris, J. R. J. Phys. Chem. C 2967.
2012, 116, 4535. (172) Murakami, Y.; Endo, K.; Ohta, I.; Nosaka, A. Y.; Nosaka, Y. J.
(137) Emeline, A. V.; Ryabchuk, V. K.; Serpone, N. J. Phys. Chem. B Phys. Chem. C 2007, 111, 11339.
2005, 109, 18515. (173) Park, S. W.; Jang, J. T.; Cheon, J.; Lee, H. H.; Lee, D. R.; Lee,
(138) Knorr, F. J.; Mercado, C. C.; McHale, J. L. J. Phys. Chem. C Y. J. Phys. Chem. C 2008, 112, 9627.
2008, 112, 12786. (174) Xiang, Q. J.; Lv, K. L.; Yu, J. G. Appl. Catal., B 2010, 96, 557.
(139) Winder, E. J.; Moore, D. E.; Neu, D. R.; Ellis, A. B.; Geisz, J. F.; (175) Yip, C. T.; Guo, M.; Huang, H. T.; Zhou, L. M.; Wang, Y.;
Kuech, T. F. J. Cryst. Growth 1995, 148, 63. Huang, C. J. Nanoscale 2012, 4, 448.
(140) Stevanovic, A.; Buttner, M.; Zhang, Z.; Yates, J. T. J. Am. Chem. (176) Sang, L. X.; Dai, H. X.; Sun, J. H.; Xu, L. X.; Wang, F.; Ma, C.
Soc. 2012, 134, 324. F. Int. J. Hydrogen Energy 2010, 35, 7098.
(141) Kamat, P. V.; Fox, M. A. J. Phys. Chem. 1983, 87, 59. (177) Burda, C.; Chen, X. B.; Narayanan, R.; El-Sayed, M. A. Chem.
(142) Yamakata, A.; Ishibashi, T.; Onishi, H. J. Mol. Catal. A: Chem. Rev. 2005, 105, 1025.
2003, 199, 85. (178) Dai, L.; Sow, C. H.; Lim, C. T.; Cheong, W. C. D.; Tan, V. B.
(143) Katoh, R.; Furube, A.; Barzykin, A. V.; Arakawa, H.; Tachiya, C. Nano Lett. 2009, 9, 576.
M. Coord. Chem. Rev. 2004, 248, 1195. (179) Yu, J.; Chen, Y.; Glushenkov, A. M. Cryst. Growth Des. 2009, 9,
(144) He, T.; Hu, Z. S.; Li, J. L.; Yang, G. W. J. Phys. Chem. C 2011, 1240.
115, 13837. (180) Kandiel, T. A.; Dillert, R.; Feldhoff, A.; Bahnemann, D. W. J.
(145) Tachikawa, T.; Fujitsuka, M.; Majima, T. J. Phys. Chem. C Phys. Chem. C 2010, 114, 4909.
2007, 111, 5259. (181) Wu, Y.; Liu, H. M.; Xu, B. Q.; Zhang, Z. L.; Su, D. S. Inorg.
(146) Dimitrijevic, N. M.; Shkrob, I. A.; Gosztola, D. J.; Rajh, T. J. Chem. 2007, 46, 5093.
Phys. Chem. C 2012, 116, 878. (182) Chuangchote, S.; Jitputti, J.; Sagawa, T.; Yoshikawa, S. ACS
(147) Tamaki, Y.; Hara, K.; Katoh, R.; Tachiya, M.; Furube, A. J. Appl. Mater. Interfaces 2009, 1, 1140.
Phys. Chem. C 2009, 113, 11741. (183) Choi, S. K.; Kim, S.; Lim, S. K.; Park, H. J. Phys. Chem. C 2010,
(148) Katoh, R.; Furube, A.; Yoshihara, T.; Hara, K.; Fujihashi, G.; 114, 16475.
Takano, S.; Murata, S.; Arakawa, H.; Tachiya, M. J. Phys. Chem. B (184) Kim, G. M.; Lee, S. M.; Michler, G. H.; Roggendorf, H.;
2004, 108, 4818. Gosele, U.; Knez, M. Chem. Mater. 2008, 20, 3085.

9313 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(185) He, T.; Pan, F. C.; Xi, Z. X.; Zhang, X. J.; Zhang, H. Y.; Wang, (222) Zhu, K.; Vinzant, T. B.; Neale, N. R.; Frank, A. J. Nano Lett.
Z. H.; Zhao, M. W.; Yan, S. S.; Xia, Y. Y. J. Phys. Chem. C 2010, 114, 2007, 7, 3739.
9234. (223) Wang, J.; Lin, Z. Q. Chem.Asian J. 2012, 7, 2754.
(186) Ribeiro, C.; Vila, C.; Stroppa, D. B.; Mastelaro, V. R.; Bettini, (224) Roy, P.; Kim, D.; Paramasivam, I.; Schmuki, P. Electrochem.
J.; Longo, E.; Leite, E. R. J. Phys. Chem. C 2007, 111, 5871. Commun. 2009, 11, 1001.
(187) Chen, Y.; Kang, K. S.; Yoo, K. H.; Jyoti, N.; Kim, J. J. Phys. (225) Alivov, Y.; Fan, Z. Y. Appl. Phys. Lett. 2009, 95, 3.
Chem. C 2009, 113, 19753. (226) Sheng, J. A.; Hu, L. H.; Xu, S. Y.; Liu, W. Q.; Mo, L. E.; Tian,
(188) Kumar, A.; Jose, R.; Fujihara, K.; Wang, J.; Ramakrishna, S. H. J.; Dai, S. Y. J. Mater. Chem. 2011, 21, 5457.
Chem. Mater. 2007, 19, 6536. (227) Zhang, Q. F.; Cao, G. Z. Nano Today 2011, 6, 91.
(189) Kubo, T.; Nakahira. J. Phys. Chem. C 2008, 112, 1658. (228) Sun, P. P.; Zhang, X. T.; Wang, C. H.; Wei, Y. A.; Wang, L. L.;
(190) Khan, M. A.; Yang, O. B. Cryst. Growth Des. 2009, 9, 1767. Liu, Y. C. J. Mater. Chem. A 2013, 1, 3309.
(191) Fahim, N. F.; Sekino, T. Chem. Mater. 2009, 21, 1967. (229) Zhang, Z. Y.; Sang, L. X.; Sun, B. A.; Zhang, X. M.; Ma, C. F.
(192) He, T.; Zhao, M. W.; Zhang, X. J.; Zhang, H. Y.; Wang, Z. H.; Acta Phys.Chim. Sin. 2010, 26, 2935.
Xi, Z. X.; Liu, X. D.; Yan, S. S.; Xia, Y. Y.; Mei, L. M. J. Phys. Chem. C (230) Zhang, H. M.; Liu, P. R.; Liu, X. L.; Zhang, S. Q.; Yao, X. D.;
2009, 113, 13610. An, T. C.; Amal, R.; Zhao, H. J. Langmuir 2010, 26, 11226.
(193) Palummo, M.; Giorgi, G.; Chiodo, L.; Rubio, A.; Yamashita, K. (231) Wang, F.; Liu, Y.; Dong, W.; Shen, M. R.; Kang, Z. H. J. Phys.
J. Phys. Chem. C 2012, 116, 18495. Chem. C 2011, 115, 14635.
(194) Alivov, Y.; Fan, Z. Y. J. Phys. Chem. C 2009, 113, 12954. (232) Colonna, D.; Colodrero, S.; Lindstrom, H.; Di Carlo, A.;
(195) Mogilevsky, G.; Chen, Q.; Kulkarni, H.; Kleinhammes, A.; Miguez, H. Energy Environ. Sci. 2012, 5, 8238.
Mullins, W. M.; Wu, Y. J. Phys. Chem. C 2008, 112, 3239. (233) Lv, M. Q.; Zheng, D. J.; Ye, M. D.; Sun, L.; Xiao, J.; Guo, W.
(196) Wang, C.; Yin, L.; Zhang, L.; Qi, Y.; Lun, N.; Liu, N. Langmuir X.; Lin, C. J. Nanoscale 2012, 4, 5872.
2010, 26, 12841. (234) Ye, M.; Zheng, D.; Lv, M.; Chen, C.; Lin, C.; Lin, Z. Adv.
(197) Ge, M.; Li, J. W.; Liu, L.; Zhou, Z. Ind. Eng. Chem. Res. 2011, Mater. 2013, 25, 3039.
50, 6681. (235) Ye, M. D.; Liu, H. Y.; Lin, C. J.; Lin, Z. Q. Small 2013, 9, 312.
(198) Masuda, Y.; Ohji, T.; Kato, K. Cryst. Growth Des. 2010, 10, 913. (236) Ye, M. D.; Xin, X. K.; Lin, C. J.; Lin, Z. Q. Nano Lett. 2011, 11,
(199) Yang, X. F.; Zhuang, J. L.; Li, X. Y.; Chen, D. H.; Ouyang, G. 3214.
F.; Mao, Z. Q.; Han, Y. X.; He, Z. H.; Liang, C. L.; Wu, M. M.; Yu, J. (237) Lv, M. Q.; Zheng, D. J.; Ye, M. D.; Xiao, J.; Guo, W. X.; Lai, Y.
C. ACS Nano 2009, 3, 1212. K.; Sun, L.; Lin, C. J.; Zuo, J. Energy Environ. Sci. 2013, 6, 1615.
(200) Soejima, T.; Jin, R.-H.; Terayama, Y.; Takahara, A.; Shiraishi, (238) Anpo, M.; Takeuchi, M. J. Catal. 2003, 216, 505.
T.; Ito, S.; Kimizuka, N. Langmuir 2012, 28, 2637. (239) Das, K.; Sharma, S. N.; Kumar, M.; De, S. K. J. Phys. Chem. C
(201) Bian, Z. F.; Tachikawa, T.; Majima, T. J. Phys. Chem. Lett.
2009, 113, 14783.
2012, 3, 1422.
(240) Kumar, S. G.; Devi, L. G. J. Phys. Chem. A 2011, 115, 13211.
(202) Flaherty, D. W.; Dohnalek, Z.; Dohnalkova, A.; Arey, B. W.;
(241) Bhattacharyya, K.; Varma, S.; Tripathi, A. K.; Bharadwaj, S. R.;
McCready, D. E.; Ponnusamy, N.; Mullins, C. B.; Kay, B. D. J. Phys.
Tyagi, A. K. J. Phys. Chem. C 2008, 112, 19102.
Chem. C 2007, 111, 4765.
(242) Rockafellow, E. M.; Haywood, J. M.; Witte, T.; Houk, R. S.;
(203) Hartmann, P.; Lee, D. K.; Smarsly, B. M.; Janek, J. ACS Nano
Jenks, W. S. Langmuir 2010, 26, 19052.
2010, 4, 3147.
(243) Sang, L. X.; Ma, C. F.; Sun, J. H.; Dai, H. X.; Wang, F.; Li, J.
(204) Ji, Y. J.; Lin, K. C.; Zheng, H. G.; Liu, C. C.; Dudik, L.; Zhu, J.
J.; Burda, C. ACS Appl. Mater. Interfaces 2010, 2, 3075. Chem. J. Chin. Univ. 2006, 27, 1608.
(205) Shao, F.; Sun, J.; Gao, L. A.; Yang, S. W.; Luo, J. Q. J. Phys. (244) Qiu, X. F.; Zhao, Y. X.; Burda, C. Adv. Mater. 2007, 19, 3995.
(245) Wu, G.; Nishikawa, T.; Ohtani, B.; Chen, A. Chem. Mater.
Chem. C 2011, 115, 1819.
(206) Hwang, Y. J.; Hahn, C.; Liu, B.; Yang, P. D. ACS Nano 2012, 6, 2007, 19, 4530.
5060. (246) Chen, X. B.; Burda, C. J. Am. Chem. Soc. 2008, 130, 5018.
(207) Wang, C. H.; Zhang, X. T.; Zhang, Y. L.; Jia, Y.; Yang, J. K.; (247) Yang, K. S.; Dai, Y.; Huang, B. B. J. Phys. Chem. C 2010, 114,
Sun, P. P.; Liu, Y. C. J. Phys. Chem. C 2011, 115, 22276. 19830.
(208) Pol, V. G.; Zaban, A. J. Phys. Chem. C 2007, 111, 14574. (248) Serpone, N. J. Phys. Chem. B 2006, 110, 24287.
(209) Joshi, R. K.; Schneider, J. J. Chem. Soc. Rev. 2012, 41, 5285. (249) Qiu, X. F.; Burda, C. Chem. Phys. 2007, 339, 1.
(210) Fang, W. Q.; Zhou, J. Z.; Liu, J.; Chen, Z. G.; Yang, C.; Sun, C. (250) Burda, C.; Lou, Y. B.; Chen, X. B.; Samia, A. C. S.; Stout, J.;
H.; Qian, G. R.; Zou, J.; Qiao, S. Z.; Yang, H. G. Chem.Eur. J. 2011, Gole, J. L. Nano Lett. 2003, 3, 1049.
17, 1423. (251) Liu, Y.; Li, J.; Qiu, X. F.; Burda, C. J. Photochem. Photobiol., A
(211) Liu, S.; Yu, J.; Jaroniec, M. Chem. Mater. 2011, 23, 4085. 2007, 190, 94.
(212) Hosono, E.; Matsuda, H.; Honma, I.; Ichihara, M.; Zhou, H. (252) Clouser, S.; Samia, A. C. S.; Navok, E.; Alred, J.; Burda, C. Top.
Langmuir 2007, 23, 7447. Catal. 2008, 47, 42.
(213) Hagfeldt, A.; Grätzel, M. Chem. Rev. 1995, 95, 49. (253) Liu, Y.; Li, J.; Qiu, X.; Burda, C. Water Sci. Technol. 2006, 54,
(214) Oregan, B.; Grätzel, M. Nature 1991, 353, 737. 47.
(215) Frank, A. J.; Kopidakis, N.; van de Lagemaat, J. Coord. Chem. (254) Wang, J. W.; Mao, B. D.; Gole, J. L.; Burda, C. Nanoscale 2010,
Rev. 2004, 248, 1165. 2, 2257.
(216) Nakade, S.; Matsuda, M.; Kambe, S.; Saito, Y.; Kitamura, T.; (255) Chen, X. B.; Burda, C.; Guo, J. H.; Smith, K. E.; Glans, P. A.;
Sakata, T.; Wada, Y.; Mori, H.; Yanagida, S. J. Phys. Chem. B 2002, 106, Learmonth, T. Arabian J. Sci. Eng. 2010, 35, 65.
10004. (256) Zhao, Y. X.; Qiu, X. F.; Burda, C. Chem. Mater. 2008, 20, 2629.
(217) Nakade, S.; Saito, Y.; Kubo, W.; Kitamura, T.; Wada, Y.; (257) Mowbray, D. J.; Martinez, J. I.; Lastra, J. M. G.; Thygesen, K.
Yanagida, S. J. Phys. Chem. B 2003, 107, 8607. S.; Jacobsen, K. W. J. Phys. Chem. C 2009, 113, 12301.
(218) Ito, S.; Murakami, T. N.; Comte, P.; Liska, P.; Grätzel, C.; (258) Asahi, R.; Morikawa, T. Chem. Phys. 2007, 339, 57.
Nazeeruddin, M. K.; Grätzel, M. Thin Solid Films 2008, 516, 4613. (259) Asong, N.; Dukes, F.; Wang, C.-y.; Shultz, M. J. Chem. Phys.
(219) Feng, X. J.; Zhu, K.; Frank, A. J.; Grimes, C. A.; Mallouk, T. E. 2007, 339, 86.
Angew. Chem.-Int. Ed. 2012, 51, 2727. (260) Batzill, M.; Morales, E. H.; Diebold, U. Chem. Phys. 2007, 339,
(220) Zhu, K.; Frank, A. J. MRS Bull. 2011, 36, 446. 36.
(221) Zhu, K.; Neale, N. R.; Miedaner, A.; Frank, A. J. Nano Lett. (261) Bellardita, M.; Addamo, M.; Di Paola, A.; Palmisano, L. Chem.
2007, 7, 69. Phys. 2007, 339, 94.

9314 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(262) Beranek, R.; Neumann, B.; Sakthivel, S.; Janczarek, M.; (298) Nishimura, S.; Abrams, N.; Lewis, B. A.; Halaoui, L. I.;
Dittrich, T.; Tributsch, H.; Kisch, H. Chem. Phys. 2007, 339, 11. Mallouk, T. E.; Benkstein, K. D.; van de Lagemaat, J.; Frank, A. J. J.
(263) Chambers, S. A.; Cheung, S. H.; Shutthanandan, V.; Am. Chem. Soc. 2003, 125, 6306.
Thevuthasan, S.; Bowman, M. K.; Joly, A. G. Chem. Phys. 2007, 339, (299) Lee, S. H. A.; Abrams, N. M.; Hoertz, P. G.; Barber, G. D.;
27. Halaoui, L. I.; Mallouk, T. E. J. Phys. Chem. B 2008, 112, 14415.
(264) Di Valentin, C.; Finazzi, E.; Pacchioni, G.; Selloni, A.; Livraghi, (300) Guldin, S.; Hütttner, S.; Kolle, M.; Welland, M. E.; Müller-
S.; Paganini, M. C.; Giamello, E. Chem. Phys. 2007, 339, 44. Buschbaum, P.; Friend, R. H.; Steiner, U.; Tétreault, N. Nano Lett.
(265) Gombac, V.; De Rogatis, L.; Gasparotto, A.; Vicario, G.; 2010, 10, 2303.
Montini, T.; Barreca, D.; Balducci, G.; Fornasiero, P.; Tondello, E.; (301) Mandlmeier, B.; Szeifert, J. M.; Fattakhova-Rohlfing, D.;
Graziani, M. Chem. Phys. 2007, 339, 111. Amenitsch, H.; Bein, T. J. Am. Chem. Soc. 2011, 133, 17274.
(266) Kudo, A.; Niishiro, R.; Iwase, A.; Kato, H. Chem. Phys. 2007, (302) Liu, J.; Liu, G. L.; Li, M. Z.; Shen, W. Z.; Liu, Z. Y.; Wang, J. X.;
339, 104. Zhao, J. C.; Jiang, L.; Song, Y. L. Energy Environ. Sci. 2010, 3, 1503.
(267) Nakano, Y.; Morikawa, T.; Ohwaki, T.; Taga, Y. Chem. Phys. (303) Chen, H. A.; Chen, S.; Quan, X.; Zhang, Y. B. Environ. Sci.
2007, 339, 20. Technol. 2010, 44, 451.
(268) Nishijima, K.; Ohtani, B.; Yan, X.; Kamai, T.-a.; Chiyoya, T.; (304) Liao, G. Z.; Chen, S.; Quan, X.; Chen, H.; Zhang, Y. B. Environ.
Tsubota, T.; Murakami, N.; Ohno, T. Chem. Phys. 2007, 339, 64. Sci. Technol. 2011, 44, 3481.
(269) Obata, K.; Irie, H.; Hashimoto, K. Chem. Phys. 2007, 339, 124. (305) Liu, J.; Li, M. Z.; Wang, J. X.; Song, Y. L.; Jiang, L.; Murakami,
(270) Qiu, X.; Burda, C. Chem. Phys. 2007, 339, 1. T.; Fujishima, A. Environ. Sci. Technol. 2009, 43, 9425.
(271) Shaban, Y. A.; Khan, S. U. M. Chem. Phys. 2007, 339, 73. (306) Li, Y. Z.; Kunitake, T.; Fujikawa, S. J. Phys. Chem. B 2006, 110,
(272) Ma, X. G.; Wu, Y.; Lu, Y. H.; Xu, J.; Wang, Y. J.; Zhu, Y. F. J. 13000.
Phys. Chem. C 2011, 115, 16963. (307) Li, J.; Qin, Y.; Jin, C.; Li, Y.; Shi, D. L.; Schmidt-Mende, L.;
(273) Chen, D. M.; Jiang, Z. Y.; Geng, J. Q.; Wang, Q.; Yang, D. Ind. Gan, L. H.; Yang, J. H. Nanoscale 2013, 5, 5009.
Eng. Chem. Res. 2007, 46, 2741. (308) Nowotny, M. K.; Bak, T.; Nowotny, J.; Sorrell, C. C. Phys.
(274) Hamal, D. B.; Klabunde, K. J. J. Phys. Chem. C 2011, 115, Status Solidi 2005, 242, 88.
17359. (309) Nowotny, J.; Bak, T.; Nowotny, M. K.; Sheppard, L. R. Int. J.
(275) Kuvarega, A. T.; Krause, R. W. M.; Mamba, B. B. J. Phys. Chem. Hydrogen Energy 2007, 32, 2630.
C 2011, 115, 22110. (310) Nowotny, J.; Bak, T.; Nowotny, M. K.; Sheppard, L. R. Int. J.
(276) Chiodi, M.; Cheney, C. P.; Vilmercati, P.; Cavaliere, E.; Hydrogen Energy 2007, 32, 2609.
Mannella, N.; Weitering, H. H.; Gavioli, L. J. Phys. Chem. C 2012, 116, (311) Samiee, M.; Luo, J. Mater. Lett. 2013, 98, 205.
(312) Sang, L. X.; Zhang, Z. Y.; Ma, C. F. Int. J. Hydrogen Energy
311.
2011, 36, 4732.
(277) Xu, H.; Zhang, L. Z. J. Phys. Chem. C 2010, 114, 940.
(313) Kimmel, G. A.; Petrik, N. G. Phys. Rev. Lett. 2008, 100, 196102.
(278) Breault, T. M.; Bartlett, B. M. J. Phys. Chem. C 2012, 116, 5986.
(314) Ohtsu, N.; Kodama, K.; Kitagawa, K.; Wagatsuma, K. Appl.
(279) Yang, M. J.; Hume, C.; Lee, S.; Son, Y. H.; Lee, J. K. J. Phys.
Surf. Sci. 2010, 256, 4522.
Chem. C 2010, 114, 15292.
(315) Hashimoto, S.; Tanaka, A. Surf. Interface Anal. 2002, 34, 262.
(280) Han, S. H.; Lee, S.; Shin, H.; Jung, H. S. Adv. Energy Mater.
(316) Sang, L. X.; Zhang, Z. Y.; Bai, G. M.; Du, C. X.; Ma, C. F. Int. J.
2011, 1, 546. Hydrogen Energy 2012, 37, 854.
(281) Calvo, M. E.; Colodrero, S.; Rojas, T. C.; Anta, J. A.; Ocana, (317) Teleki, A.; Pratsinis, S. E. Phys. Chem. Chem. Phys. 2009, 11,
M.; Miguez, H. Adv. Funct. Mater. 2008, 18, 2708. 3742.
(282) Colodrero, S.; Forneli, A.; Lopez-Lopez, C.; Pelleja, L.; (318) Komaguchi, K.; Maruoka, T.; Nakano, H.; Imae, I.; Ooyama,
Miguez, H.; Palomares, E. Adv. Funct. Mater. 2012, 22, 1303. Y.; Harima, Y. J. Phys. Chem. C 2010, 114, 124.
(283) Seo, Y. G.; Woo, K.; Kim, J.; Lee, H.; Lee, W. Adv. Funct. (319) Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P.
Mater. 2011, 21, 3094. Y. J. Am. Chem. Soc. 2010, 132, 11856.
(284) Calvo, M. E.; Colodrero, S.; Hidalgo, N.; Lozano, G.; Lopez- (320) Xing, M. Y.; Fang, W. Z.; Nasir, M.; Ma, Y. F.; Zhang, J. L.;
Lopez, C.; Sanchez-Sobrado, O.; Miguez, H. Energy Environ. Sci. 2011, Anpo, M. J. Catal. 2013, 297, 236.
4, 4800. (321) Zhou, X. S.; Yang, F.; Jin, B.; Huang, Y. S.; Wu, Z. J. Mater.
(285) Tetreault, N.; Grätzel, M. Energy Environ. Sci. 2013, 5, 8506. Lett. 2013, 112, 145.
(286) Sun, Z. Q.; Kim, J. H.; Zhao, Y.; Bijarbooneh, F.; Malgras, V.; (322) Zuo, F.; Wang, L.; Feng, P. Y. Int. J. Hydrogen Energy 2014, 39,
Dou, S. X. J. Mater. Chem. 2012, 22, 11711. 711.
(287) Jin, M.; Kim, S. S.; Yoon, M.; Li, Z.; Lee, Y. Y.; Kim, J. M. J. (323) Zhang, Z. H.; Yang, X. L.; Hedhili, M. N.; Ahmed, E.; Shi, L.;
Nanosci. Nanotechnol. 2012, 12, 815. Wang, P. ACS Appl. Mater. Interfaces. 2011, 6, 691.
(288) Chen, J. I. L.; Loso, E.; Ebrahim, N.; Ozin, G. A. J. Am. Chem. (324) Khomenko, V. M.; Langer, K.; Rager, H.; Fett, A. Phys. Chem.
Soc. 2008, 130, 5420. Miner. 1998, 25, 338.
(289) Orilall, M. C.; Abrams, N. M.; Lee, J.; DiSalvo, F. J.; Wiesner, (325) Anpo, M.; Che, M.; Fubini, B.; Garrone, E.; Giamello, E.;
U. J. Am. Chem. Soc. 2008, 130, 8882. Paganini, M. C. Top. Catal. 1999, 8, 189.
(290) Cheng, C. W.; Karuturi, S. K.; Liu, L. J.; Liu, J. P.; Li, H. X.; Su, (326) Bickley, R. I.; Gonzalez-Carreno, T.; Lees, J. S.; Palmisano, L.;
L. T.; Tok, A. I. Y.; Fan, H. J. Small 2012, 8, 37. Tilley, R. J. D. J. Solid State Chem. 1991, 92, 178.
(291) Ruani, G.; Ancora, C.; Corticelli, F.; Dionigi, C.; Rossi, C. Sol. (327) Hurum, D. C.; Agrios, A. G.; Gray, K. A.; Rajh, T.; Thurnauer,
Energy Mater. Sol. Cells 2008, 92, 537. M. C. J. Phys. Chem. B 2003, 107, 4545.
(292) Yablonovitch, E. Phys. Rev. Lett. 1987, 58, 2059. (328) Hurum, D. C.; Agrios, A. G.; Crist, S. E.; Gray, K. A.; Rajh, T.;
(293) John, S. Phys. Rev. Lett. 1987, 58, 2486. Thurnauer, M. C. J. Electron Spectrosc. 2006, 150, 155.
(294) Wijnhoven, J.; Vos, W. L. Science 1998, 281, 802. (329) Yan, M. C.; Chen, F.; Zhang, J. L.; Anpo, M. J. Phys. Chem. B
(295) Kwak, E. S.; Lee, W.; Park, N. G.; Kim, J.; Lee, H. Adv. Funct. 2005, 109, 8673.
Mater. 2009, 19, 1093. (330) Cong, S.; Xu, Y. M. J. Phys. Chem. C 2011, 115, 21161.
(296) Halaoui, L. I.; Abrams, N. M.; Mallouk, T. E. J. Phys. Chem. B (331) Nair, R. G.; Paul, S.; Samdarshi, S. K. Sol. Energy Mater. Sol.
2005, 109, 6334. Cells 2011, 95, 1901.
(297) Chen, J. I. L.; von Freymann, G.; Kitaev, V.; Ozin, G. A. J. Am. (332) He, H.; Liu, C.; Dubois, K. D.; Jin, T.; Louis, M. E.; Li, G. H.
Chem. Soc. 2007, 129, 1196. Ind. Eng. Chem. Res. 2012, 51, 11841.

9315 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(333) Antoniadou, M.; Panagiotopoulou, P.; Kondarides, D. I.; (366) Zhang, N.; Liu, S. Q.; Fu, X. Z.; Xu, Y. J. J. Phys. Chem. C 2011,
Lianos, P. J. Appl. Electrochem. 2012, 42, 737. 115, 9136.
(334) Liu, S.; Yan, Y. H.; Guan, W. W.; Li, M. Z.; Jiang, R. Y. Asian J. (367) Awazu, K.; Fujimaki, M.; Rockstuhl, C.; Tominaga, J.;
Chem. 2013, 25, 1307. Murakami, H.; Ohki, Y.; Yoshida, N.; Watanabe, T. J. Am. Chem.
(335) Pan, L.; Huang, H.; Lim, C. K.; Hong, Q. Y.; Tse, M. S.; Tan, Soc. 2008, 130, 1676.
O. K. RSC Adv. 2013, 3, 3566. (368) Ingram, D. B.; Christopher, P.; Bauer, J. L.; Linic, S. ACS Catal.
(336) Carneiro, J. T.; Savenije, T. J.; Moulijn, J. A.; Mul, G. J. Phys. 2011, 1, 1441.
Chem. C 2011, 115, 2211. (369) Warren, S. C.; Thimsen, E. Energy Environ. Sci. 2012, 5, 5133.
(337) Scanlon, D. O.; Dunnill, C. W.; Buckeridge, J.; Shevlin, S. A.; (370) Sauthier, G.; del Pino, A. P.; Figueras, A.; Gyorgy, E. J. Am.
Logsdail, A. J.; Woodley, S. M.; Catlow, C. R. A.; Powell, M.. J.; Ceram. Soc. 2011, 94, 3780.
Palgrave, R. G.; Parkin, I. P.; Watson, G. W.; Keal, T. W.; Sherwood, (371) Zhou, J. B.; Cheng, Y.; Yu, J. G. J. Photochem. Photobiol., A
P.; Walsh, A.; Sokol, A. A. Nat. Mater. 2013, 12, 798. 2011, 223, 82.
(338) Pfeifer, V.; Erhart, P.; Li, S.; Rachut, K.; Morasch, J.; Brotz, J.; (372) Nakata, K.; Udagawa, K.; Tryk, D. A.; Ochiai, T.; Nishimoto,
Reckers, P.; Mayer, T.; Ruhle, S.; Zaban, A.; Sero, I. M.; Bisquert, J.; S.; Sakai, H.; Murakami, T.; Abe, M.; Fujishima, A. Mater. Lett. 2009,
Jaegermann, W.; Klein, A. J. Phys. Chem. Lett. 2013, 4, 4182. 63, 1628.
(339) Kavan, L.; Grätzel, M.; Gilbert, S. E.; Klemenz, C.; Scheel, H. J. (373) Christopher, P.; Ingram, D. B.; Linic, S. J. Phys. Chem. C 2010,
Am. Chem. Soc. 1996, 118, 6716. 114, 9173.
(340) Xiong, G.; Shao, R.; Droubay, T. C.; Joly, A. G.; Beck, K. M.; (374) Es-Souni, M.; Habouti, S.; Pfeiffer, N.; Lahmar, A.; Dietze, M.;
Chambers, S. A.; Hess, W. P. Adv. Funct. Mater. 2007, 17, 2133. Solterbeck, C. H. Adv. Funct. Mater. 2010, 20, 377.
(341) Bojinova, A.; Kralchevska, R.; Poulios, I.; Dushkin, C. Mater. (375) Logar, M.; Jancar, B.; Sturm, S.; Suvorov, D. Langmuir 2010,
Chem. Phys. 2007, 106, 187. 26, 12215.
(342) Bernardini, C.; Cappelletti, G.; Dozzi, M. V.; Selli, E. J. (376) Wang, P.; Huang, B. B.; Zhang, Q. Q.; Zhang, X. Y.; Qin, X. Y.;
Photochem. Photobiol., A 2010, 211, 185. Dai, Y.; Zhan, J.; Yu, J. X.; Liu, H. X.; Lou, Z. Z. Chem.Eur. J. 2010,
(343) Su, R.; Bechstein, R.; So, L.; Vang, R. T.; Sillassen, M.; 16, 10042.
Esbjornsson, B.; Palmqvist, A.; Besenbacher, F. J. Phys. Chem. C 2011, (377) Wu, T. S.; Wang, K. X.; Li, G. D.; Sun, S. Y.; Sun, J.; Chen, J. S.
115, 24287. ACS Appl. Mater. Interfaces 2010, 2, 544.
(344) Guo, Y. W.; Cheng, C. P.; Wang, J.; Wang, Z. Q.; Jin, X. D.; Li, (378) Xiang, Q. J.; Yu, J. G.; Cheng, B.; Ong, H. C. Chem.Asian J.
2010, 5, 1466.
K.; Kang, P. L.; Gao, J. Q. J. Hazard. Mater. 2011, 192, 786.
(379) Liu, Z. W.; Hou, W. B.; Pavaskar, P.; Aykol, M.; Cronin, S. B.
(345) Jiang, D. L.; Zhang, S. Q.; Zhao, H. J. Environ. Sci. Technol.
Nano Lett. 2011, 11, 1111.
2007, 41, 303.
(380) Wen, Y. Y.; Ding, H. M.; Shan, Y. K. Nanoscale 2011, 3, 4411.
(346) Zhang, J.; Xu, Q.; Feng, Z.; Li, M.; Li, C. Angew. Chem., Int. Ed.
(381) Abdulla-Al-Mamun, M.; Kusumoto, Y.; Islam, M. S. J. Mater.
2008, 47, 1766.
Chem. 2012, 22, 5460.
(347) Xu, Q. A.; Ma, Y.; Zhang, J.; Wang, X. L.; Feng, Z. C.; Li, C. J.
(382) Guo, J. F.; Ma, B. W.; Yin, A. Y.; Fan, K. N.; Dai, W. L. J.
Catal. 2011, 278, 329. Hazard. Mater. 2012, 211, 77.
(348) Ma, Y.; Xu, Q.; Chong, R. F.; Li, C. J. Mater. Res. 2012, 28, 394. (383) Jia, Z. X.; Ben Amar, M.; Brinza, O.; Astafiev, A.; Nadtochenko,
(349) vander Meulen, T.; Mattson, A.; Osterlund, L. J. Catal. 2007,
V.; Evlyukhin, A. B.; Chichkov, B. N.; Duten, X.; Kanaev, A. J. Phys.
251, 131. Chem. C 2012, 116, 17239.
(350) Hsu, Y. C.; Lin, H. C.; Chen, C. H.; Liao, Y. T.; Yang, C. M. J. (384) Jiang, J.; Li, H.; Zhang, L. Z. Chem.Eur. J. 2012, 18, 6360.
Solid State Chem. 2010, 183, 1917. (385) Jiang, L. M.; Zhou, G.; Mi, J.; Wu, Z. Y. Catal. Commun. 2012,
(351) Kho, Y. K.; Iwase, A.; Teoh, W. Y.; Madler, L.; Kudo, A.; Amal, 24, 48.
R. J. Phys. Chem. C 2010, 114, 2821. (386) Liu, J.; Chen, F. Y. Int. J. Electrochem. Sci. 2012, 7, 9560.
(352) Li, G.; Chen, L.; Graham, M. E.; Gray, K. A. J. Mol. Catal. A: (387) Lu, Y.; Yu, H. T.; Chen, S.; Quan, X.; Zhao, H. M. Environ. Sci.
Chem. 2007, 275, 30. Technol. 2012, 46, 1724.
(353) Xu, H.; Zhang, L. Z. J. Phys. Chem. C 2009, 113, 1785. (388) Nishijima, Y.; Ueno, K.; Kotake, Y.; Murakoshi, K.; Inoue, H.;
(354) Dinh, C. T.; Nguyen, T. D.; Kleitz, F.; Do, T. O. ACS Appl. Misawa, H. J. Phys. Chem. Lett. 2012, 3, 1248.
Mater.Interfaces 2011, 3, 2228. (389) Seh, Z. W.; Liu, S. H.; Low, M.; Zhang, S. Y.; Liu, Z. L.;
(355) Wang, W.-N.; An, W.-J.; Ramalingam, B.; Mukherjee, S.; Mlayah, A.; Han, M. Y. Adv. Mater. 2012, 24, 2310.
Niedzwiedzki, D. M.; Gangopadhyay, S.; Biswas, P. J. Am. Chem. Soc. (390) Subrahmanyam, A.; Biju, K. P.; Rajesh, P.; Kumar, K. J.; Kiran,
2012, 134, 11276. M. R. Sol. Energy Mater. Sol. Cells 2012, 101, 241.
(356) Cesano, F.; Bertarione, S.; Uddin, M. J.; Agostini, G.; Scarano, (391) Xu, J. X.; Xiao, X. H.; Ren, F.; Wu, W.; Dai, Z. G.; Cai, G. X.;
D.; Zecchina, A. J. Phys. Chem. C 2010, 114, 169. Zhang, S. F.; Zhou, J.; Mei, F.; Jiang, C. Z. Nanoscale Res. Lett. 2012, 7,
(357) Li, H. X.; Bian, Z. F.; Zhu, J.; Huo, Y. N.; Li, H.; Lu, Y. F. J. 6.
Am. Chem. Soc. 2007, 129, 4538. (392) Yang, Y. Q.; Zhang, G. K.; Xu, W. J. Colloid Interface Sci. 2012,
(358) Mohapatra, S. K.; Kondamudi, N.; Banerjee, S.; Misra, M. 376, 217.
Langmuir 2008, 24, 11276. (393) Rajeshwar, K.; Janaky, C.; Lin, W. Y.; Roberts, D. A.; Wampler,
(359) Ohyama, J.; Yamamoto, A.; Teramura, K.; Shishido, T.; W. J. Phys. Chem. Lett. 2013, 4, 3468.
Tanaka, T. ACS Catal. 2011, 1, 187. (394) Grabowska, E.; Zaleska, A.; Sorgues, S.; Kunst, M.; Etcheberry,
(360) Sun, C. H.; Smith, S. C. J. Phys. Chem. C 2012, 116, 3524. A.; Colbeau-Justin, C.; Remita, H. J. Phys. Chem. C 2013, 117, 1955.
(361) Pearson, A.; Jani, H.; Kalantar-Zadeh, K.; Bhargava, S. K.; (395) Shi, X.; Ueno, K.; Takabayashi, N.; Misawa, H. J. Phys. Chem. C
Bansal, V. Langmuir 2011, 27, 6661. 2013, 117, 2494.
(362) Cozzoli, P. D.; Comparelli, R.; Fanizza, E.; Curri, M. L.; (396) Chen, J. J.; Wu, J. C. S.; Wu, P. C.; Tsai, D. P. J. Phys. Chem. C
Agostiano, A.; Laub, D. J. Am. Chem. Soc. 2004, 126, 3868. 2012, 116, 26535.
(363) Zhu, W.; Wang, G. Z.; Hong, X.; Shen, X. S. J. Phys. Chem. C (397) Tseng, Y. H.; Chang, I. G.; Tai, Y. A.; Wu, K. W. J. Nanosci.
2009, 113, 5450. Nanotechnol. 2012, 12, 416.
(364) Borras, A.; Barranco, A.; Gonzalez-Elipe, A. R. Langmuir 2008, (398) Wang, H.; You, T. T.; Shi, W. W.; Li, J. H.; Guo, L. J. Phys.
24, 8021. Chem. C 2012, 116, 6490.
(365) Wu, X. F.; Song, H. Y.; Yoon, J. M.; Yu, Y. T.; Chen, Y. F. (399) Zhang, Z. H.; Zhang, L. B.; Hedhili, M. N.; Zhang, H. N.;
Langmuir 2009, 25, 6438. Wang, P. Nano Lett. 2012, 13, 14.

9316 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(400) Alessandri, I.; Ferroni, M.; Depero, L. E. ChemPhysChem 2009, (435) Sambur, J. B.; Riha, S. C.; Choi, D.; Parkinson, B. A. Langmuir
10, 1017. 2010, 26, 4839.
(401) Zhang, X. Q.; Zhu, Y. H.; Yang, X. L.; Wang, S. W.; Shen, J. H.; (436) Tada, H.; Fujishima, M.; Kobayashi, H. Chem. Soc. Rev. 2011,
Lin, B. B.; Li, C. Z. Nanoscale 2013, 5, 3359. 40, 4232.
(402) Song, H. Y.; Yu, Y. T.; Norby, P. J. Nanosci. Nanotechnol. 2009, (437) Wu, G. S.; Tian, M.; Chen, A. C. J. Photochem. Photobiol., A
9, 5891. 2012, 233, 65.
(403) Christopher, P.; Xin, H.; Marimuthu, A.; Linic, S. Nat. Mater. (438) Beard, M. C. J. Phys. Chem. Lett. 2011, 2, 1282.
2012, 11, 1044. (439) Bubenhofer, S. B.; Schumacher, C. M.; Koehler, F. M.;
(404) Linic, S.; Christopher, P.; Ingram, D. B. Nat. Mater. 2011, 10, Luechinger, N. A.; Grass, R. N.; Stark, W. J. J. Phys. Chem. C 2012,
911. 116, 16264.
(405) Zhou, Z. Y.; Tian, N.; Li, J. T.; Broadwell, I.; Sun, S. G. Chem. (440) Mora-Sero, I.; Gimenez, S.; Fabregat-Santiago, F.; Gomez, R.;
Soc. Rev. 2011, 40, 4167. Shen, Q.; Toyoda, T.; Bisquert, J. Acc. Chem. Res. 2009, 42, 1848.
(406) Youngblood, W. J.; Lee, S. H. A.; Maeda, K.; Mallouk, T. E. (441) Pernik, D. R.; Tvrdy, K.; Radich, J. G.; Kamat, P. V. J. Phys.
Acc. Chem. Res. 2009, 42, 1966. Chem. C 2011, 115, 13511.
(407) Zhang, Z. F.; Deng, Z. B.; Liang, C. J.; Zhang, M. X.; Xu, D. H. (442) Robel, I.; Subramanian, V.; Kuno, M.; Kamat, P. V. J. Am.
Displays 2003, 24, 231. Chem. Soc. 2006, 128, 2385.
(408) Mont, F. W.; Schubert, E. F. J. Appl. Phys. 2010, 107, (443) Acharya, K. P.; Alabi, T. R.; Schmall, N.; Hewa-Kasakarage, N.
083120(1−5). N.; Kirsanova, M.; Nemchinov, A.; Khon, E.; Zamkov, M. J. Phys.
(409) Liu, X. Y.; Zhou, W. J.; Yin, Z. M.; Hao, X. P.; Wu, Y. Z.; Xu, X. Chem. C 2009, 113, 19531.
G. J. Mater. Chem. 2012, 22, 3916. (444) Shen, F. Y.; Que, W. X.; Liao, Y. L.; Yin, X. T. Ind. Eng. Chem.
(410) Li, Z.; Sun, Q.; Yao, X. D.; Zhu, Z. H.; Lu, G. Q. J. Mater. Res. 2011, 50, 9131.
Chem. 2012, 22, 22821. (445) Cheng, K. C.; Law, W. C.; Yong, K. T.; Nevins, J. S.; Watson,
(411) Ozdogan, K.; Kahaly, M. U.; Kumar, S. R. S.; Alshareef, H. N.; D. F.; Ho, H. P.; Prasad, P. N. Chem. Phys. Lett. 2011, 515, 254.
Schwingenschlogl, U. J. Appl. Phys. 2012, 111, 5. (446) Mao, B. D.; Chuang, C. H.; Lu, F.; Sang, L. X.; Zhu, J. J.;
(412) Pan, J.; Huhne, S. M.; Shen, H.; Xiao, L. S.; Born, P.; Mader, Burda, C. J. Phys. Chem. C 2013, 117, 648.
W.; Mathur, S. J. Phys. Chem. C 2011, 115, 17265. (447) Mao, B. D.; Chuang, C. H.; Wang, J. W.; Burda, C. J. Phys.
(413) Liu, S. Q.; Li, Y. X.; Xie, M. J.; Guo, X. F.; Ji, W. J.; Ding, W. P. Chem. C 2011, 115, 8945.
Mater. Lett. 2010, 64, 402. (448) Ning, Z. J.; Tian, H. N.; Qin, H. Y.; Zhang, Q. O.; Agren, H.;
(414) Bai, H. W.; Juay, J.; Liu, Z. Y.; Song, X. X.; Lee, S. S.; Sun, D. Sun, L. C.; Fu, Y. J. Phys. Chem. C 2010, 114, 15184.
D. Appl. Catal., B 2012, 125, 367. (449) Jin, S. Y.; Lian, T. Q. Nano Lett. 2009, 9, 2448.
(415) Zhang, J.; Bang, J. H.; Tang, C. C.; Kamat, P. V. ACS Nano (450) Zhang, Z.; Yates, J. T. Chem. Rev. 2012, 112, 5520.
(451) Shiraishi, Y.; Hirai, T. J. Photochem. Photobiol., C 2008, 9, 157.
2010, 4, 387.
(452) Kudo, A.; Miseki, Y. Chem. Soc. Rev. 2009, 38, 253.
(416) Liu, R.; Yang, W. D.; Qiang, L. S.; Liu, H. Y. J. Power Sources
(453) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 2228.
2012, 220, 153.
(454) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 7860.
(417) Garcia-Ramirez, E.; Mondragon-Chaparro, M.; Zelaya-Angel,
(455) Lakshminarasimhan, N.; Bae, E.; Choi, W. J. Phys. Chem. C
O. Appl. Phys. A: Mater. Sci. Process. 2012, 108, 291.
2007, 111, 15244.
(418) Zhu, L.; Liu, G. C.; Duan, X. C.; Zhang, Z. J. J. Mater. Res.
(456) Mercado, C. C.; Knorr, F. J.; McHale, J. L.; Usmani, S. M.;
2010, 25, 1278. Ichimura, A. S.; Saraf, L. V. J. Phys. Chem. C 2012, 116, 10796.
(419) Cui, H.; Dwight, K.; Soled, S.; Wold, A. J. Solid. State. Chem. (457) Chretien, S.; Metiu, H. J. Phys. Chem. C 2011, 115, 4696.
1995, 115, 187. (458) Maitani, M. M.; Tanaka, K.; Mochizuki, D.; Wada, Y. J. Phys.
(420) Joskowska, D.; Pomoni, K.; Vomvas, A.; Koscielska, B.; Chem. Lett. 2011, 2, 2655.
Anastassopoulos, D. L. J. Non-Cryst. Solids 2010, 356, 2042. (459) Tao, J. G.; Batzill, M. J. Phys. Chem. Lett. 2010, 1, 3200.
(421) Furukawa, S.; Shishido, T.; Teramura, K.; Tanaka, T. ACS (460) Sevinc, P. C.; Wang, X.; Wang, Y. M.; Zhang, D.; Meixner, A.
Catal. 2012, 2, 175. J.; Lu, H. P. Nano Lett. 2011, 11, 1490.
(422) Wang, C. H.; Shao, C. L.; Zhang, X. T.; Liu, Y. C. Inorg. Chem. (461) Takai, A.; Kamat, P. V. ACS Nano 2011, 5, 7369.
2009, 48, 7261. (462) Kamat, P. V. J. Phys. Chem. Lett. 2012, 3, 663.
(423) Seo, H.; Posadas, A. B.; Mitra, C.; Kvit, A. V.; Ramdani, J.; (463) Jankovic, I. A.; Saponjic, Z. V.; Comor, M. I.; Nedeljlkovic, J.
Demkov, A. A. Phys. Rev. B 2012, 86, 075301. M. J. Phys. Chem. C 2009, 113, 12645.
(424) D’Amico, N. R.; Cantele, G.; Ninno, D. Appl. Phys. Lett. 2012, (464) Siedl, N.; Elser, M. J.; Bernardi, J.; Diwald, O. J. Phys. Chem. C
101, 141606. 2009, 113, 15792.
(425) Cao, T. P.; Li, Y. J.; Wang, C. H.; Shao, C. L.; Liu, Y. C. (465) Tamaki, Y.; Furube, A.; Murai, M.; Hara, K.; Katoh, R.;
Langmuir 2011, 27, 2946. Tachiya, M. Phys. Chem. Chem. Phys. 2007, 9, 1453.
(426) Mu, J. B.; Chen, B.; Zhang, M. Y.; Guo, Z. C.; Zhang, P.; (466) Chuang, C. H.; Lo, S. S.; Scholes, G. D.; Burda, C. J. Phys.
Zhang, Z. Y.; Sun, Y. Y.; Shao, C. L.; Liu, Y. C. ACS Appl. Mater. Chem. Lett. 2010, 1, 2530.
Interfaces 2012, 4, 424. (467) Chuang, C. H.; Doane, T. L.; Lo, S. S.; Scholes, G. D.; Burda,
(427) Xu, Q. C.; Wellia, D. V.; Ng, Y. H.; Amal, R.; Tan, T. T. Y. J. C. ACS Nano 2011, 5, 6016.
Phys. Chem. C 2011, 115, 7419. (468) Lo, S. S.; Mirkovic, T.; Chuang, C. H.; Burda, C.; Scholes, G.
(428) Li, X.; Hou, Y.; Zhao, Q.; Chen, G. Langmuir 2011, 27, 3113. D. Adv. Mater. 2011, 23, 180.
(429) Xie, Y.; Ali, G.; Yoo, S. H.; Cho, S. O. ACS Appl. Mater. (469) Ho, W. K.; Yu, J. C.; Lin, J.; Yu, J. G.; Li, P. S. Langmuir 2004,
Interfaces 2010, 2, 2910. 20, 5865.
(430) Fujii, M.; Nagasuna, K.; Fujishima, M.; Akita, T.; Tada, H. J. (470) Tang, J. W.; Durrant, J. R.; Klug, D. R. J. Am. Chem. Soc. 2008,
Phys. Chem. C 2009, 113, 16711. 130, 13885.
(431) Yin, Y. X.; Jin, Z. G.; Hou, F. Nanotechnology 2007, 18, 495608. (471) Linsebigler, A. L.; Lu, G. Q.; Yates, J. T. Chem. Rev. 1995, 95,
(432) Lee, J. C.; Kim, T. G.; Choi, H. J.; Sung, Y. M. Cryst. Growth 735.
Des. 2007, 7, 2588. (472) Fox, M. A.; Dulay, M. T. Chem. Rev. 1993, 93, 341.
(433) Guijarro, N.; Lana-Villarreal, T.; Mora-Sero, I.; Bisquert, J.; (473) Kamat, P. V. Chem. Rev. 1993, 93, 267.
Gomez, R. J. Phys. Chem. C 2009, 113, 4208. (474) Zhang, P. Y.; Liu, J. J. Photochem. Photobiol., A 2004, 167, 87.
(434) Mann, J. R.; Watson, D. F. Langmuir 2007, 23, 10924. (475) Lyu, J.; Zhu, L.; Burda, C. ChemCatChem. 2013, 5, 3114.

9317 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318


Chemical Reviews Review

(476) Kubacka, A.; Fernandez-Garcia, M.; Colon, G. Chem. Rev. (513) Liu, M. Z.; Johnston, M. B.; Snaith, H. J. Nature 2013, 501,
2012, 112, 1555. 395.
(477) Naito, K.; Tachikawa, T.; Fujitsuka, M.; Majima, T. J. Phys. (514) Kim, H. S.; Lee, C. R.; Im, J. H.; Lee, K. B.; Moehl, T.;
Chem. C 2008, 112, 1048. Marchioro, A.; Moon, S. J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.
(478) Lewis, N. S. Annu. Rev. Phys. Chem. 1991, 42, 543. E.; Grätzel, M.; Park, N.-G. Sci. Rep. 2012, 2, 1.
(479) Teoh, W. Y.; Denny, F.; Amal, R.; Friedmann, D.; Madler, L.; (515) Zhao, Y.; Zhu, K. J. Phys. Chem. Lett. 2013, 4, 2880.
Pratsinis, S. E. Top. Catal. 2007, 44, 489. (516) Heo, J. H.; Im, S. H.; Noh, J. H.; Mandal, T. N.; Lim, C.-S.;
(480) Silva, C. G.; Juarez, R.; Marino, T.; Molinari, R.; Garcia, H. J. Chang, J. A.; Lee, Y. H.; Kim, H. J.; Sarkar, A.; Nazeeruddin, Md. K.;
Am. Chem. Soc. 2011, 133, 595. Grätzel, M.; Seok, S. I. Nat. Photonics 2013, 7, 486.
(481) Lyu, J.; Zhu, L.; Burda, C. Catal. Today 2014, 225, 24.
(482) Abe, T.; Suzuki, E.; Nagoshi, K.; Miyashita, K.; Kaneko, M. J.
Phys. Chem. B 1999, 103, 1119.
(483) Lin, F.; Zhang, Y. N.; Wang, L.; Zhang, Y. L.; Wang, D. G.;
Yang, M.; Yang, J. H.; Zhang, B. Y.; Jiang, Z. X.; Li, C. Appl. Catal., B
2012, 127, 363.
(484) Meekins, B. H.; Kamat, P. V. J. Phys. Chem. Lett. 2011, 2, 2304.
(485) Sang, L. X.; Tan, H. Y.; Zhang, X. M.; Wu, Y. T.; Ma, C. F.;
Burda, C. J. Phys. Chem. C 2012, 116, 18633.
(486) Hesabi, Z. R.; Allam, N. K.; Dahmen, K.; Garmestani, H.; El-
Sayed, M. A. ACS Appl. Mater. Interfaces 2011, 3, 952.
(487) Lambert, T. N.; Chavez, C. A.; Hernandez-Sanchez, B.; Lu, P.;
Bell, N. S.; Ambrosini, A.; Friedman, T.; Boyle, T. J.; Wheeler, D. R.;
Huber, D. L. J. Phys. Chem. C 2009, 113, 19812.
(488) Bell, N. J.; Yun, H. N.; Du, A. J.; Coster, H.; Smith, S. C.; Amal,
R. J. Phys. Chem. C 2011, 115, 6004.
(489) Lightcap, I. V.; Kosel, T. H.; Kamat, P. V. Nano Lett. 2010, 10,
577.
(490) Indrakanti, V. P.; Schobert, H. H.; Kubicki, J. D. Energy Fuels
2009, 23, 5247.
(491) Liu, L. J.; Zhao, H. L.; Andino, J. M.; Li, Y. ACS Catal. 2012, 2,
1817.
(492) Anpo, M.; Yamashita, H.; Ichihashi, Y.; Ehara, S. J. Electroanal.
Chem. 1995, 396, 21.
(493) Dimitrijevic, N. M.; Vijayan, B. K.; Poluektov, O. G.; Rajh, T.;
Gray, K. A.; He, H. Y.; Zapol, P. J. Am. Chem. Soc. 2011, 133, 3964.
(494) Yang, C. C.; Yu, Y. H.; van der Linden, B.; Wu, J. C. S.; Mul, G.
J. Am. Chem. Soc. 2010, 132, 8398.
(495) Labat, F.; Ciofini, I.; Hratchian, H. P.; Frisch, M. J.;
Raghavachari, K.; Adamo, C. J. Phys. Chem. C 2011, 115, 4297.
(496) Asbury, J. B.; Wang, Y. Q.; Hao, E. C.; Ghosh, H. N.; Lian, T.
Q. Res. Chem. Intermed. 2001, 27, 393.
(497) Meng, S.; Ren, J.; Kaxiras, E. Nano Lett. 2008, 8, 3266.
(498) Grätzel, M. Nature 2001, 414, 338.
(499) Benko, G.; Kallioinen, J.; Korppi-Tommola, J. E. I.; Yartsev, A.
P.; Sundstrom, V. J. Am. Chem. Soc. 2002, 124, 489.
(500) Du, L. C.; Furube, A.; Hara, K.; Katoh, R.; Tachiya, M. J. Phys.
Chem. C 2010, 114, 8135.
(501) Martsinovich, N.; Jones, D. R.; Troisi, A. J. Phys. Chem. C 2010,
114, 22659.
(502) Tiwana, P.; Docampo, P.; Johnston, M. B.; Snaith, H. J.; Herz,
L. M. ACS Nano 2011, 5, 5158.
(503) Clifford, J. N.; Palomares, E.; Nazeeruddin, M. K.; Grätzel, M.;
Nelson, J.; Li, X.; Long, N. J.; Durrant, J. R. J. Am. Chem. Soc. 2004,
126, 5225.
(504) Weng, Y. X.; Wang, Y. Q.; Asbury, J. B.; Ghosh, H. N.; Lian, T.
Q. J. Phys. Chem. B 2000, 104, 93.
(505) Nelson, J.; Chandler, R. E. Coord. Chem. Rev. 2004, 248, 1181.
(506) Barzykin, A. V.; Tachiya, M. J. Phys. Chem. B 2002, 106, 4356.
(507) Du, L. C.; Weng, Y. X. J. Phys. Chem. C 2007, 111, 4567.
(508) Lang, X.; Chen, X.; Zhao, J. Chem. Soc. Rev. 2014, 43, 473.
(509) Liu, L.; Liu, Z.; Liu, A.; Gu, X.; Ge, C.; Gao, F.; Dong, L.
ChemSusChem 2014, 7, 618.
(510) Sheng, X.; He, D.; Yang, J.; Zhu, K.; Feng, X. Nano Lett. 2014,
14, 1848.
(511) Wu, W. Q.; Xu, Y. F.; Rao, H. S.; Feng, H. L.; Su, C. Y.; Kuang,
D. B. Angew. Chem., Int. Ed. 2014, 53, 1.
(512) Burschka, J.; Pellet, N.; Moon, S. J.; Humphry-Baker, R.; Gao,
P.; Nazeeruddin, M. K.; Grätzel, M. Nature 2013, 499, 316.

9318 dx.doi.org/10.1021/cr400629p | Chem. Rev. 2014, 114, 9283−9318

You might also like