You are on page 1of 69

Chem. Rev.

2007, 107, 2891−2959 2891

Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and


Applications
Xiaobo Chen* and Samuel S. Mao†
Lawrence Berkeley National Laboratory, and University of California, Berkeley, California 94720

Received March 27, 2006

Contents 5.2. Photovoltaic Applications 2932


5.2.1. The TiO2 Nanocrystalline Electrode in 2932
1. Introduction 2891 DSSCs
2. Synthetic Methods for TiO2 Nanostructures 2892 5.2.2. Metal/Semiconductor Junction Schottky 2938
2.1. Sol−Gel Method 2892 Diode Solar Cell
2.2. Micelle and Inverse Micelle Methods 2895 5.2.3. Doped TiO2 Nanomaterials-Based Solar 2938
2.3. Sol Method 2896 Cell
2.4. Hydrothermal Method 2898 5.3. Photocatalytic Water Splitting 2939
2.5. Solvothermal Method 2901 5.3.1. Fundamentals of Photocatalytic Water 2939
2.6. Direct Oxidation Method 2902 Splitting
2.7. Chemical Vapor Deposition 2903 5.3.2. Use of Reversible Redox Mediators 2939
2.8. Physical Vapor Deposition 2904 5.3.3. Use of TiO2 Nanotubes 2940
2.9. Electrodeposition 2904 5.3.4. Water Splitting under Visible Light 2941
2.10. Sonochemical Method 2904 5.3.5. Coupled/Composite Water-Splitting 2942
2.11. Microwave Method 2904 System
2.12. TiO2 Mesoporous/Nanoporous Materials 2905 5.4. Electrochromic Devices 2942
2.13. TiO2 Aerogels 2906 5.4.1. Fundamentals of Electrochromic Devices 2943
2.14. TiO2 Opal and Photonic Materials 2907 5.4.2. Electrochromophore for an Electrochromic 2943
Device
2.15. Preparation of TiO2 Nanosheets 2908
5.4.3. Counterelectrode for an Electrochromic 2944
3. Properties of TiO2 Nanomaterials 2909 Device
3.1. Structural Properties of TiO2 Nanomaterials 2909 5.4.4. Photoelectrochromic Devices 2945
3.2. Thermodynamic Properties of TiO2 2911 5.5. Hydrogen Storage 2945
Nanomaterials
5.6. Sensing Applications 2947
3.3. X-ray Diffraction Properties of TiO2 2912
Nanomaterials 6. Summary 2948
3.4. Raman Vibration Properties of TiO2 2912 7. Acknowledgment 2949
Nanomaterials 8. References 2949
3.5. Electronic Properties of TiO2 Nanomaterials 2913
3.6. Optical Properties of TiO2 Nanomaterials 2915 1. Introduction
3.7. Photon-Induced Electron and Hole Properties 2918
of TiO2 Nanomaterials Since its commercial production in the early twentieth
4. Modifications of TiO2 Nanomaterials 2920 century, titanium dioxide (TiO2) has been widely used as a
pigment1 and in sunscreens,2,3 paints,4 ointments, toothpaste,5
4.1. Bulk Chemical Modification: Doping 2921
etc. In 1972, Fujishima and Honda discovered the phenom-
4.1.1. Synthesis of Doped TiO2 Nanomaterials 2921 enon of photocatalytic splitting of water on a TiO2 electrode
4.1.2. Properties of Doped TiO2 Nanomaterials 2921 under ultraviolet (UV) light.6-8 Since then, enormous efforts
4.2. Surface Chemical Modifications 2926 have been devoted to the research of TiO2 material, which
4.2.1. Inorganic Sensitization 2926 has led to many promising applications in areas ranging from
5. Applications of TiO2 Nanomaterials 2929 photovoltaics and photocatalysis to photo-/electrochromics
5.1. Photocatalytic Applications 2929 and sensors.9-12 These applications can be roughly divided
5.1.1. Pure TiO2 Nanomaterials: First 2930 into “energy” and “environmental” categories, many of which
Generation depend not only on the properties of the TiO2 material itself
5.1.2. Metal-Doped TiO2 Nanomaterials: 2930 but also on the modifications of the TiO2 material host (e.g.,
Second Generation with inorganic and organic dyes) and on the interactions of
5.1.3. Nonmetal-Doped TiO2 Nanomaterials: 2931 TiO2 materials with the environment.
Third Generation An exponential growth of research activities has been seen
in nanoscience and nanotechnology in the past decades.13-17
* Corresponding author. E-mail: XChen3@lbl.gov. New physical and chemical properties emerge when the size
† E-mail: SSMao@lbl.gov. of the material becomes smaller and smaller, and down to
10.1021/cr0500535 CCC: $65.00 © 2007 American Chemical Society
Published on Web 06/23/2007
2892 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

many serious environmental and pollution challenges. TiO2


also bears tremendous hope in helping ease the energy crisis
through effective utilization of solar energy based on
photovoltaic and water-splitting devices.9,31,32 As continued
breakthroughs have been made in the preparation, modifica-
tion, and applications of TiO2 nanomaterials in recent years,
especially after a series of great reviews of the subject in
the 1990s.7,8,10-12,33,45 we believe that a new and compre-
hensive review of TiO2 nanomaterials would further promote
TiO2-based research and development efforts to tackle the
environmental and energy challenges we are currently facing.
Here, we focus on recent progress in the synthesis, properties,
modifications, and applications of TiO2 nanomaterials. The
syntheses of TiO2 nanomaterials, including nanoparticles,
nanorods, nanowires, and nanotubes are primarily categorized
Dr. Xiaobo Chen is a research engineer at The University of California at with the preparation method. The preparations of mesopo-
Berkeley and a Lawrence Berkeley National Laboratory scientist. He
obtained his Ph.D. Degree in Chemistry from Case Western Reserve
rous/nanoporous TiO2, TiO2 aerogels, opals, and photonic
University. His research interests include photocatalysis, photovoltaics, materials are summarized separately. In reviewing nanoma-
hydrogen storage, fuel cells, environmental pollution control, and the related terial synthesis, we present a typical procedure and repre-
materials and devices development. sentative transmission or scanning electron microscopy
images to give a direct impression of how these nanomate-
rials are obtained and how they normally appear. For detailed
instructions on each synthesis, the readers are referred to
the corresponding literature.
The structural, thermal, electronic, and optical properties
of TiO2 nanomaterials are reviewed in the second section.
As the size, shape, and crystal structure of TiO2 nanomate-
rials vary, not only does surface stability change but also
the transitions between different phases of TiO2 under
pressure or heat become size dependent. The dependence of
X-ray diffraction patterns and Raman vibrational spectra on
the size of TiO2 nanomaterials is also summarized, as they
could help to determine the size to some extent, although
correlation of the spectra with the size of TiO2 nanomaterials
is not straightforward. The review of modifications of TiO2
Dr. Samuel S. Mao is a career staff scientist at Lawrence Berkeley National nanomaterials is mainly limited to the research related to
Laboratory and an adjunct faculty at The University of California at the modifications of the optical properties of TiO2 nanoma-
Berkeley. He obtained his Ph.D. degree in Engineering from The University terials, since many applications of TiO2 nanomaterials are
of California at Berkeley in 2000. His current research involves the closely related to their optical properties. TiO2 nanomaterials
development of nanostructured materials and devices, as well as ultrafast normally are transparent in the visible light region. By doping
laser technologies. Dr. Mao is the team leader of a high throughput or sensitization, it is possible to improve the optical sensitiv-
materials processing program supported by the U.S. Department of Ener-
gy. ity and activity of TiO2 nanomaterials in the visible light
region. Environmental (photocatalysis and sensing) and
energy (photovoltaics, water splitting, photo-/electrochromics,
and hydrogen storage) applications are reviewed with an
the nanometer scale. Properties also vary as the shapes of emphasis on clean and sustainable energy, since the increas-
the shrinking nanomaterials change. Many excellent reviews ing energy demand and environmental pollution create a
and reports on the preparation and properties of nanomaterials pressing need for clean and sustainable energy solutions. The
have been published recently.6-44 Among the unique proper- fundamentals and working principles of the TiO2 nanoma-
ties of nanomaterials, the movement of electrons and holes terials-based devices are discussed to facilitate the under-
in semiconductor nanomaterials is primarily governed by the standing and further improvement of current and practical
well-known quantum confinement, and the transport proper- TiO2 nanotechnology.
ties related to phonons and photons are largely affected by
the size and geometry of the materials.13-16 The specific
surface area and surface-to-volume ratio increase dramati-
2. Synthetic Methods for TiO2 Nanostructures
cally as the size of a material decreases.13,21 The high surface
area brought about by small particle size is beneficial to many
2.1. Sol−Gel Method
TiO2-based devices, as it facilitates reaction/interaction The sol-gel method is a versatile process used in making
between the devices and the interacting media, which mainly various ceramic materials.46-50 In a typical sol-gel process,
occurs on the surface or at the interface and strongly depends a colloidal suspension, or a sol, is formed from the hydrolysis
on the surface area of the material. Thus, the performance and polymerization reactions of the precursors, which are
of TiO2-based devices is largely influenced by the sizes of usually inorganic metal salts or metal organic compounds
the TiO2 building units, apparently at the nanometer scale. such as metal alkoxides. Complete polymerization and loss
As the most promising photocatalyst,7,11,12,33 TiO2 mate- of solvent leads to the transition from the liquid sol into a
rials are expected to play an important role in helping solve solid gel phase. Thin films can be produced on a piece of
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2893

substrate by spin-coating or dip-coating. A wet gel will form changes from cuboidal to ellipsoidal at pH above 11 with
when the sol is cast into a mold, and the wet gel is converted TEOA. The TiO2 nanoparticle shape evolves into ellipsoidal
into a dense ceramic with further drying and heat treatment. above pH 9.5 with diethylenetriamine with a higher aspect
A highly porous and extremely low-density material called ratio than that with TEOA. Figure 2 shows representative
an aerogel is obtained if the solvent in a wet gel is removed TEM images of the TiO2 nanoparticles under different initial
under a supercritical condition. Ceramic fibers can be drawn pH conditions with the shape control of TEOA at [TEOA]/
from the sol when the viscosity of a sol is adjusted into a [TIPO] ) 2.0. Secondary amines, such as diethylamine, and
proper viscosity range. Ultrafine and uniform ceramic tertiary amines, such as trimethylamine and triethylamine,
powders are formed by precipitation, spray pyrolysis, or act as complexing agents of Ti(IV) ions to promote the
emulsion techniques. Under proper conditions, nanomaterials growth of ellipsoidal particles with lower aspect ratios. The
can be obtained. shape of the TiO2 nanoparticle can also be tuned from round-
TiO2 nanomaterials have been synthesized with the sol- cornered cubes to sharp-edged cubes with sodium oleate and
gel method from hydrolysis of a titanium precusor.51-78 This sodium stearate.70 The shape control is attributed to the tuning
process normally proceeds via an acid-catalyzed hydrolysis of the growth rate of the different crystal planes of TiO2
step of titanium(IV) alkoxide followed by condensa- nanoparticles by the specific adsorption of shape controllers
tion.51,63,66,79-91 The development of Ti-O-Ti chains is to these planes under different pH conditions.70
favored with low content of water, low hydrolysis rates, and A prolonged heating time below 100 °C for the as-prepared
excess titanium alkoxide in the reaction mixture. Three- gel can be used to avoid the agglomeration of the TiO2 nano-
dimensional polymeric skeletons with close packing result particles during the crystallization process.58,72 By heating
from the development of Ti-O-Ti chains. The formation amorphous TiO2 in air, large quantities of single-phase ana-
of Ti(OH)4 is favored with high hydrolysis rates for a tase TiO2 nanoparticles with average particle sizes between
medium amount of water. The presence of a large quantity 7 and 50 nm can be obtained, as reported by Zhang and
of Ti-OH and insufficient development of three-dimensional Banfield.73-77 Much effort has been exerted to achieve highly
polymeric skeletons lead to loosely packed first-order crystallized and narrowly dispersed TiO2 nanoparticles using
particles. Polymeric Ti-O-Ti chains are developed in the the sol-gel method with other modifications, such as a
presence of a large excess of water. Closely packed first- semicontinuous reaction method by Znaidi et al.78 and a two-
order particles are yielded via a three-dimensionally devel- stage mixed method and a continuous reaction method by
oped gel skeleton.51,63,66,79-91 From the study on the growth Kim et al.53,54
kinetics of TiO2 nanoparticles in aqueous solution using By a combination of the sol-gel method and an anodic
titanium tetraisopropoxide (TTIP) as precursor, it is found alumina membrane (AAM) template, TiO2 nanorods have
that the rate constant for coarsening increases with temper- been successfully synthesized by dipping porous AAMs
ature due to the temperature dependence of the viscosity of into a boiled TiO2 sol followed by drying and heating
the solution and the equilibrium solubility of TiO2.63 Second- processes.92,93 In a typical experiment, a TiO2 sol solution is
ary particles are formed by epitaxial self-assembly of primary prepared by mixing TTIP dissolved in ethanol with a solution
particles at longer times and higher temperatures, and the containing water, acetyl acetone, and ethanol. An AAM is
number of primary particles per secondary particle increases immersed into the sol solution for 10 min after being boiled
with time. The average TiO2 nanoparticle radius increases in ethanol; then it is dried in air and calcined at 400 °C for
linearly with time, in agreement with the Lifshitz-Slyozov- 10 h. The AAM template is removed in a 10 wt % H3PO4
Wagner model for coarsening.63 aqueous solution. The calcination temperature can be used
Highly crystalline anatase TiO2 nanoparticles with different to control the crystal phase of the TiO2 nanorods. At low
sizes and shapes could be obtained with the polycondensation temperature, anatase nanorods can be obtained, while at
of titanium alkoxide in the presence of tetramethylammonium high temperature rutile nanorods can be obtained. The pore
hydroxide.52,62 In a typical procedure, titanium alkoxide is size of the AAM template can be used to control the size of
added to the base at 2 °C in alcoholic solvents in a three- these TiO2 nanorods, which typically range from 100 to 300
neck flask and is heated at 50-60 °C for 13 days or at 90- nm in diameter and several micrometers in length. Appar-
100 °C for 6 h. A secondary treatment involving autoclave ently, the size distribution of the final TiO2 nanorods is
heating at 175 and 200 °C is performed to improve the largely controlled by the size distribution of the pores of
crystallinity of the TiO2 nanoparticles. Representative TEM the AAM template. In order to obtain smaller and mono-
images are shown in Figure 1 from the study of Chemseddine sized TiO2 nanorods, it is necessary to fabricate high-quality
et al.52 AAM templates. Figure 3 shows a typical TEM for TiO2
A series of thorough studies have been conducted by nanorods fabricated with this method. Normally, the TiO2
Sugimoto et al. using the sol-gel method on the formation nanorods are composed of small TiO2 nanoparticles or
of TiO2 nanoparticles of different sizes and shapes by tuning nanograins.
the reaction parameters.67-71 Typically, a stock solution of By electrophoretic deposition of TiO2 colloidal suspensions
a 0.50 M Ti source is prepared by mixing TTIP with into the pores of an AAM, ordered TiO2 nanowire arrays
triethanolamine (TEOA) ([TTIP]/[TEOA] ) 1:2), followed can be obtained.94 In a typical procedure, TTIP is dissolved
by addition of water. The stock solution is diluted with a in ethanol at room temperature, and glacial acetic acid mixed
shape controller solution and then aged at 100 °C for 1 day with deionized water and ethanol is added under pH ) 2-3
and at 140 °C for 3 days. The pH of the solution can be with nitric acid. Platinum is used as the anode, and an AAM
tuned by adding HClO4 or NaOH solution. Amines are used with an Au substrate attached to Cu foil is used as the
as the shape controllers of the TiO2 nanomaterials and act cathode. A TiO2 sol is deposited into the pores of the AMM
as surfactants. These amines include TEOA, diethylenetri- under a voltage of 2-5 V and annealed at 500 °C for 24 h.
amine, ethylenediamine, trimethylenediamine, and triethyl- After dissolving the AAM template in a 5 wt % NaOH
enetetramine. The morphology of the TiO2 nanoparticles solution, isolated TiO2 nanowires are obtained. In order to
2894 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 1. TEM images of TiO2 nanoparticles prepared by hydrolysis of Ti(OR)4 in the presence of tetramethylammonium hydroxide.
Reprinted with permission from Chemseddine, A.; Moritz, T. Eur. J. Inorg. Chem. 1999, 235. Copyright 1999 Wiley-VCH.

Figure 2. TEM images of uniform anatase TiO2 nanoparticles. Reprinted from Sugimoto, T.; Zhou, X.; Muramatsu, A. J. Colloid Interface
Sci. 2003, 259, 53, Copyright 2003, with permission from Elsevier.

fabricate TiO2 nanowires instead of nanorods, an AAM with the AAM is first prepared by sucking TiO2 sol into the pores
long pores is a must. of the AAM and removing it under vacuum; TiO2 nanowires
TiO2 nanotubes can also be obtained using the sol-gel are obtained after the sol is fully developed and the AAM is
method by templating with an AAM95-98 and other organic removed. In the procedure by Lee and co-workers,96 a TTIP
compounds.99,100 For example, when an AAM is used as the solution was prepared by mixing TTIP with 2-propanol and
template, a thin layer of TiO2 sol on the wall of the pores of 2,4-pentanedione. After the AAM was dipped into this
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2895

Figure 5. SEM of a TiO2 nanotube array; the inset shows the ZnO
nanorod array template. Reprinted with permission from Qiu, J. J.;
Yu, W. D.; Gao, X. D.; Li, X. M. Nanotechnology 2006, 17, 4695.
Copyright 2006 IOP Publishing Ltd.

deposited on a ZnO nanorod template by dip-coating with a


Figure 3. TEM image of anatase nanorods and a single nanorod slow withdrawing speed, then dried at 100 °C for 10 min,
composed of small TiO2 nanoparticles or nanograins (inset). and heated at 550 °C for 1 h in air to obtain ZnO/TiO2
Reprinted from Miao, L.; Tanemura, S.; Toh, S.; Kaneko, K.;
Tanemura, M. J. Cryst. Growth 2004, 264, 246, Copyright 2004, nanorod arrays. The ZnO nanorod template is etched-up by
with permission from Elsevier. immersing the ZnO/TiO2 nanorod arrays in a dilute hydro-
chloric acid aqueous solution to obtain TiO2 nanotube arrays.
Figure 5 shows a typical SEM image of the TiO2 nanotube
array with the ZnO nanorod array template. The TiO2
nanotubes inherit the uniform hexagonal cross-sectional
shape and the length of 1.5 µm and inner diameter of 100-
120 nm of the ZnO nanorod template. As the concentration
of the TiO2 sol is constant, well-aligned TiO2 nanotube arrays
can only be obtained from an optimal dip-coating cycle
number in the range of 2-3 cycles. A dense porous TiO2
thick film with holes is obtained instead if the dip-coating
number further increases. The heating rate is critical to the
formation of TiO2 nanotube arrays. When the heating rate
is extra rapid, e.g., above 6 °C min-1, the TiO2 coat will
easily crack and flake off from the ZnO nanorods due to
great tensile stress between the TiO2 coat and the ZnO
template, and a TiO2 film with loose, porous nanostructure
is obtained.
Figure 4. SEM image of TiO2 nanotubes prepared from the AAO 2.2. Micelle and Inverse Micelle Methods
template. Reprinted with permission from Liu, S. M.; Gan, L. M.;
Liu, L. H.; Zhang, W. D.; Zeng, H. C. Chem. Mater. 2002, 14, Aggregates of surfactant molecules dispersed in a liquid
1391. Copyright 2002 American Chemical Society. colloid are called micelles when the surfactant concentration
exceeds the critical micelle concentration (CMC). The CMC
solution, it was removed from the solution and placed under is the concentration of surfactants in free solution in
vacuum until the entire volume of the solution was pulled equilibrium with surfactants in aggregated form. In micelles,
through the AAM. The AAM was hydrolyzed by water vapor the hydrophobic hydrocarbon chains of the surfactants are
over a HCl solution for 24 h, air-dried at room temperature, oriented toward the interior of the micelle, and the hydro-
and then calcined in a furnace at 673 K for 2 h and cooled philic groups of the surfactants are oriented toward the
to room temperature with a temperature ramp of 2 °C/h. Pure surrounding aqueous medium. The concentration of the lipid
TiO2 nanotubes were obtained after the AAM was dissolved present in solution determines the self-organization of the
in a 6 M NaOH solution for several minutes.96 Alternatively, molecules of surfactants and lipids. The lipids form a single
TiO2 nanotubes could be obtained by coating the AAM layer on the liquid surface and are dispersed in solution below
membranes at 60 °C for a certain period of time (12-48 h) the CMC. The lipids organize in spherical micelles at the
with dilute TiF4 under pH ) 2.1 and removing the AAM first CMC (CMC-I), into elongated pipes at the second CMC
after TiO2 nanotubes were fully developed.97 Figure 4 shows (CMC-II), and into stacked lamellae of pipes at the lamellar
a typical SEM image of the TiO2 nanotube array from the point (LM or CMC-III). The CMC depends on the chemical
AAM template.97 composition, mainly on the ratio of the head area and the
In another scheme, a ZnO nanorod array on a glass tail length. Reverse micelles are formed in nonaqueous
substrate can be used as a template to fabricate TiO2 media, and the hydrophilic headgroups are directed toward
nanotubes with the sol-gel method.101 Briefly, TiO2 sol is the core of the micelles while the hydrophobic groups are
2896 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

directed outward toward the nonaqueous media. There is no


obvious CMC for reverse micelles, because the number of
aggregates is usually small and they are not sensitive to the
surfactant concentration. Micelles are often globular and
roughly spherical in shape, but ellipsoids, cylinders, and
bilayers are also possible. The shape of a micelle is a function
of the molecular geometry of its surfactant molecules and
solution conditions such as surfactant concentration, tem-
perature, pH, and ionic strength.
Micelles and inverse micelles are commonly employed to
synthesize TiO2 nanomaterials.102-110 A statistical experi-
mental design method was conducted by Kim et al. to
optimize experimental conditions for the preparation of TiO2
nanoparticles.103 The values of H2O/surfactant, H2O/titanium
precursor, ammonia concentration, feed rate, and reaction
temperature were significant parameters in controlling TiO2
nanoparticle size and size distribution. Amorphous TiO2 Figure 6. TEM images of the shuttle-like and round-shaped (inset)
TiO2 nanoparticles. From: Zhang, D., Qi, L., Ma, J., Cheng, H. J.
nanoparticles with diameters of 10-20 nm were synthesized Mater. Chem. 2002, 12, 3677 (http://dx.doi.org/10.1039/b206996b).
and converted to the anatase phase at 600 °C and to the more s Reproduced by permission of The Royal Society of Chemistry.
thermodynamically stable rutile phase at 900 °C. Li et al.
developed TiO2 nanoparticles with the chemical reactions
between TiCl4 solution and ammonia in a reversed micro-
emulsion system consisting of cyclohexane, poly(oxyethyl-
ene)5 nonyle phenol ether, and poly(oxyethylene)9 nonyle
phenol ether.104 The produced amorphous TiO2 nanoparticles
transformed into anatase when heated at temperatures from
200 to 750 °C and into rutile at temperatures higher than
750 °C. Agglomeration and growth also occurred at elevated
temperatures.
Shuttle-like crystalline TiO2 nanoparticles were synthesized
by Zhang et al. with hydrolysis of titanium tetrabutoxide in
the presence of acids (hydrochloric acid, nitric acid, sulfuric
acid, and phosphoric acid) in NP-5 (Igepal CO-520)-
cyclohexane reverse micelles at room temperature.110 The
crystal structure, morphology, and particle size of the TiO2
nanoparticles were largely controlled by the reaction condi-
tions, and the key factors affecting the formation of rutile at
room temperature included the acidity, the type of acid used,
and the microenvironment of the reverse micelles. Ag-
glomeration of the particles occurred with prolonged reaction
times and increasing the [H2O]/[NP-5] and [H2O]/[Ti-
(OC4H9)4] ratios. When suitable acid was applied, round TiO2
nanoparticles could also be obtained. Representative TEM
images of the shuttle-like and round-shaped TiO2 nanopar-
ticles are shown in Figure 6. In the study carried out by Lim Figure 7. HRTEM images of a TiO2 nanoparticle after annealing.
et al., TiO2 nanoparticles were prepared by the controlled Reprinted with permission from Lin, J.; Lin, Y.; Liu, P.; Meziani,
hydrolysis of TTIP in reverse micelles formed in CO2 with M. J.; Allard, L. F.; Sun, Y. P. J. Am. Chem. Soc. 2002, 124, 11514.
the surfactants ammonium carboxylate perfluoropolyether Copyright 2002 American Chemical Society.
(PFPECOO-NH4+) (MW 587) and poly(dimethyl amino
ethyl methacrylate-block-1H,1H,2H,2H-perfluorooctyl meth- treatment in the solid state.108 This procedure could produce
acrylate) (PDMAEMA-b-PFOMA).106 It was found that the crystalline TiO2 nanoparticles with unchanged physical
crystallite size prepared in the presence of reverse micelles dimensions and minimal agglomeration and allows the
increased as either the molar ratio of water to surfactant or preparation of highly crystalline TiO2 nanoparticles, as shown
the precursor to surfactant ratio increased. in Figure 7, from the study of Lin et al.108
The TiO2 nanomaterials prepared with the above micelle
and reverse micelle methods normally have amorphous
2.3. Sol Method
structure, and calcination is usually necessary in order to The sol method here refers to the nonhydrolytic sol-gel
induce high crystallinity. However, this process usually leads processes and usually involves the reaction of titanium
to the growth and agglomeration of TiO2 nanoparticles. The chloride with a variety of different oxygen donor molecules,
crystallinity of TiO2 nanoparticles initially (synthesized by e.g., a metal alkoxide or an organic ether.111-119
controlled hydrolysis of titanium alkoxide in reverse micelles
in a hydrocarbon solvent) could be improved by annealing TiX4 + Ti(OR)4 f 2TiO2 + 4RX (1)
in the presence of the micelles at temperatures considerably
lower than those required for the traditional calcination TiX4 + 2ROR f TiO2 + 4RX (2)
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2897

Figure 8. TEM image of TiO2 nanoparticles derived from reaction


of TiCl4 and TTIP in TOPO/heptadecane at 300 °C. The inset shows
a HRTEM image of a single particle. Reprinted with permission
from Trentler, T. J.; Denler, T. E.; Bertone, J. F.; Agrawal, A.;
Colvin, V. L. J. Am. Chem. Soc. 1999, 121, 1613. Copyright 1999
American Chemical Society.

The condensation between Ti-Cl and Ti-OR leads to the Figure 9. TEM of TiO2 nanorods. The inset shows a HRTEM of
formation of Ti-O-Ti bridges. The alkoxide groups can a TiO2 nanorod. Reprinted with permission from Cozzoli, P. D.;
be provided by titanium alkoxides or can be formed in situ Kornowski, A.; Weller, H. J. Am. Chem. Soc. 2003, 125, 14539.
by reaction of the titanium chloride with alcohols or ethers. Copyright 2003 American Chemical Society.
In the method by Trentler and Colvin,119 a metal alkoxide
was rapidly injected into the hot solution of titanium halide can help synthesize monodispersed TiO2 nanoparticles.120,121
mixed with trioctylphosphine oxide (TOPO) in heptadecane For example, Scolan and Sanchez found that monodisperse
at 300 °C under dry inert gas protection, and reactions were nonaggregated TiO2 nanoparticles in the 1-5 nm range were
completed within 5 min. For a series of alkyl substituents obtained through hydrolysis of titanium butoxide in the
including methyl, ethyl, isopropyl, and tert-butyl, the reaction presence of acetylacetone and p-toluenesulfonic acid at 60
rate dramatically increased with greater branching of R, while °C.120 The resulting nanoparticle xerosols could be dispersed
average particle sizes were relatively unaffected. Variation in water-alcohol or alcohol solutions at concentrations
of X yielded a clear trend in average particle size, but without higher than 1 M without aggregation, which is attributed to
a discernible trend in reaction rate. Increased nucleophilicity the complexation of the surface by acetylacetonato ligands
(or size) of the halide resulted in smaller anatase nanocrystals. and through an adsorbed hybrid organic-inorganic layer
Average sizes ranged from 9.2 nm for TiF4 to 3.8 nm for made with acetylacetone, p-toluenesulfonic acid, and wa-
TiI4. The amount of passivating agent (TOPO) influenced ter.120
the chemistry. Reaction in pure TOPO was slower and With the aid of surfactants, different sized and shaped TiO2
resulted in smaller particles, while reactions without TOPO nanorods can be synthesized.122-130 For example, the growth
were much quicker and yielded mixtures of brookite, rutile, of high-aspect-ratio anatase TiO2 nanorods has been reported
and anatase with average particle sizes greater than 10 nm. by Cozzoli and co-workers by controlling the hydrolysis
Figure 8 shows typical TEM images of TiO2 nanocrystals process of TTIP in oleic acid (OA).122-126,130 Typically, TTIP
developed by Trentler et al.119 was added into dried OA at 80-100 °C under inert gas
In the method used by Niederberger and Stucky,111 TiCl4 protection (nitrogen flow) and stirred for 5 min. A 0.1-2 M
was slowly added to anhydrous benzyl alcohol under aqueous base solution was then rapidly injected and kept at
vigorous stirring at room temperature and was kept at 40- 80-100 °C for 6-12 h with stirring. The bases employed
150 °C for 1-21 days in the reaction vessel. The precipitate included organic amines, such as trimethylamino-N-oxide,
was calcinated at 450 °C for 5 h after thoroughly washing. trimethylamine, tetramethylammonium hydroxide, tetrabut-
The reaction between TiCl4 and benzyl alcohol was found ylammonium hydroxyde, triethylamine, and tributylamine.
suitable for the synthesis of highly crystalline anatase phase In this reaction, by chemical modification of the titanium
TiO2 nanoparticles with nearly uniform size and shape at precursor with the carboxylic acid, the hydrolysis rate of
very low temperatures, such as 40 °C. The particle size could titanium alkoxide was controlled. Fast (in 4-6 h) crystal-
be selectively adjusted in the range of 4-8 nm with the lization in mild conditions was promoted with the use of
appropriate thermal conditions and a proper choice of the suitable catalysts (tertiary amines or quaternary ammonium
relative amounts of benzyl alcohol and titanium tetrachloride. hydroxides). A kinetically overdriven growth mechanism led
The particle growth depended strongly on temperature, and to the growth of TiO2 nanorods instead of nanoparticles.123
lowering the titanium tetrachloride concentration led to a Typical TEM images of the TiO2 nanorods are shown in
considerable decrease of particle size.111 Figure 9.123
Surfactants have been widely used in the preparation of a Recently, Joo et al.127 and Zhang et al.129 reported similar
variety of nanoparticles with good size distribution and procedures in obtaining TiO2 nanorods without the use of
dispersity.15,16 Adding different surfactants as capping agents, catalyst. Briefly, a mixture of TTIP and OA was used to
such as acetic acid and acetylacetone, into the reaction matrix generate OA complexes of titanium at 80 °C in 1-octadecene.
2898 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 10. TEM images of TiO2 nanorods with lengths of (A) 12 nm, (B) 30 nm, and (C) 16 nm. (D) 2.3 nm TiO2 nanoparticles. Inset
in parts C and D: HR-TEM image of a single TiO2 nanorod and nanoparticle. Reprinted with permission from Zhang, Z.; Zhong, X.; Liu,
S.; Li, D.; Han, M. Angew. Chem., Int. Ed. 2005, 44, 3466. Copyright 2005 Wiley-VCH.
The injection of a predetermined amount of oleylamine at liners under controlled temperature and/or pressure with the
260 °C led to various sized TiO2 nanorods.129 Figure 10 reaction in aqueous solutions. The temperature can be
shows TEM images of TiO2 nanorods with various lengths, elevated above the boiling point of water, reaching the
and 2.3 nm TiO2 nanoparticles prepared with this method.129 pressure of vapor saturation. The temperature and the amount
In the surfactant-mediated shape evolution of TiO2 nano- of solution added to the autoclave largely determine the
crystals in nonaqueous media conducted by Jun et al.,128 it internal pressure produced. It is a method that is widely used
was found that the shape of TiO2 nanocrystals could be for the production of small particles in the ceramics industry.
modified by changing the surfactant concentration. The Many groups have used the hydrothermal method to prepare
synthesis was accomplished by an alkyl halide elimination TiO2 nanoparticles.131-140 For example, TiO2 nanoparticles
reaction between titanium chloride and titanium isopro- can be obtained by hydrothermal treatment of peptized
poxide. Briefly, a dioctyl ether solution containing TOPO precipitates of a titanium precursor with water.134 The
and lauric acid was heated to 300 °C followed by addition precipitates were prepared by adding a 0.5 M isopropanol
of titanium chloride under vigorous stirring. The reaction solution of titanium butoxide into deionized water ([H2O]/
was initiated by the rapid injection of TTIP and quenched [Ti] ) 150), and then they were peptized at 70 °C for 1 h in
with cold toluene. At low lauric acid concentrations, bullet- the presence of tetraalkylammonium hydroxides (peptizer).
and diamond-shaped nanocrystals were obtained; at higher After filtration and treatment at 240 °C for 2 h, the
concentrations, rod-shaped nanocrystals or a mixture of as-obtained powders were washed with deionized water and
nanorods and branched nanorods was observed. The bullet- absolute ethanol and then dried at 60 °C. Under the same
and diamond-shaped nanocrystals and nanorods were elon- concentration of peptizer, the particle size decreased with
gated along the [001] directions. The TiO2 nanorods were increasing alkyl chain length. The peptizers and their
found to simultaneously convert to small nanoparticles as a concentrations influenced the morphology of the particles.
function of the growth time, as shown in Figure 11, due to Typical TEM images of TiO2 nanoparticles made with the
the minimization of the overall surface energy via dissolution hydrothermal method are shown in Figure 12.134
and regrowth of monomers during an Ostwald ripening.
In another example, TiO2 nanoparticles were prepared by
hydrothermal reaction of titanium alkoxide in an acidic
2.4. Hydrothermal Method ethanol-water solution.132 Briefly, TTIP was added dropwise
Hydrothermal synthesis is normally conducted in steel to a mixed ethanol and water solution at pH 0.7 with nitric
pressure vessels called autoclaves with or without Teflon acid, and reacted at 240 °C for 4 h. The TiO2 nanoparticles
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2899

Figure 12. TEM images of TiO2 nanoparticles prepared by the


hydrothermal method. Reprinted from Yang, J.; Mei, S.; Ferreira,
J. M. F. Mater. Sci. Eng. C 2001, 15, 183, Copyright 2001, with
permission from Elsevier.

Figure 13. TEM image of TiO2 nanorods prepared with the


hydrothermal method. Reprinted with permission from Zhang, Q.;
Gao, L. Langmuir 2003, 19, 967. Copyright 2003 American
Chemical Society.
Besides TiO2 nanoparticles, TiO2 nanorods have also been
synthesized with the hydrothermal method.141-146 Zhang et
al. obtained TiO2 nanorods by treating a dilute TiCl4 solution
at 333-423 K for 12 h in the presence of acid or inorganic
salts.141,143-146 Figure 13 shows a typical TEM image of the
TiO2 nanorods prepared with the hydrothermal method.141
The morphology of the resulting nanorods can be tuned with
different surfactants146 or by changing the solvent composi-
tions.145 A film of assembled TiO2 nanorods deposited on a
glass wafer was reported by Feng et al.142 These TiO2
nanorods were prepared at 160 °C for 2 h by hydrothermal
Figure 11. Time dependent shape evolution of TiO2 nanorods: treatment of a titanium trichloride aqueous solution super-
(a) 0.25 h; (b) 24 h; (c) 48 h. Scale bar ) 50 nm. Reprinted with saturated with NaCl.
permission from Jun, Y. W.; Casula, M. F.; Sim, J. H.; Kim, S.
Y.; Cheon, J.; Alivisatos, A. P. J. Am. Chem. Soc. 2003, 125, 15981. TiO2 nanowires have also been successfully obtained with
Copyright 2003 American Chemical Society. the hydrothermal method by various groups.147-151 Typically,
TiO2 nanowires are obtained by treating TiO2 white powders
in a 10-15 M NaOH aqueous solution at 150-200 °C for
synthesized under this acidic ethanol-water environment 24-72 h without stirring within an autoclave. Figure 14
were mainly primary structure in the anatase phase without shows the SEM images of TiO2 nanowires and a TEM image
secondary structure. The sizes of the particles were controlled of a single nanowire prepared by Zhang and co-workers.150
to the range of 7-25 nm by adjusting the concentration of TiO2 nanowires can also be prepared from layered titanate
Ti precursor and the composition of the solvent system. particles using the hydrothermal method as reported by Wei
2900 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 15. TEM images of TiO2 nanowires made from the layered
Na2Ti3O7 particles, with the HRTEM image shown in the inset.
Reprinted from Wei, M.; Konishi, Y.; Zhou, H.; Sugihara, H.;
Figure 14. SEM images of TiO2 nanowires with the inset showing Arakawa, H. Chem. Phys. Lett. 2004, 400, 231, Copyright 2004,
a TEM image of a single TiO2 nanowire with a [010] selected area with permission from Elsevier.
electron diffraction (SAED) recorded perpendicular to the long axis
of the wire. Reprinted from Zhang, Y. X.; Li, G. H.; Jin, Y. X.;
Zhang, Y.; Zhang, J.; Zhang, L. D. Chem. Phys. Lett. 2002, 365,
300, Copyright 2002, with permission from Elsevier.

et al.152 In their experiment, layer-structured Na2Ti3O7 was


dispersed into a 0.05-0.1 M HCl solution and kept at 140-
170 °C for 3-7 days in an autoclave. TiO2 nanowires were
obtained after the product was washed with H2O and finally
dried. In the formation of a TiO2 nanowire from layered
H2Ti3O7, there are three steps: (i) the exfoliation of layered
Na2Ti3O7; (ii) the nanosheets formation; and (iii) the nanow-
ires formation.152 In Na2Ti3O7, [TiO6] octahedral layers are
held by the strong static interaction between the Na+ cations
between the [TiO6] octahedral layers and the [TiO6] unit.
When the larger H3+O cations replace the Na+ cations in
the interlayer space of [TiO6] sheets, this static interaction
is weakened because the interlayer distance is enlarged. As
a result, the layered compounds Na2Ti3O7 are gradually
exfoliated. When Na+ is exchanged by H+ in the dilute HCl
solution, numerous H2Ti3O7 sheet-shaped products are
formed. Since the nanosheet does not have inversion sym-
metry, an intrinsic tension exists. The nanosheets split to form
nanowires in order to release the strong stress and lower the
total energy.152 A representative TEM image of TiO2
nanowires from Na2Ti3O7 is shown in Figure 15.152
The hydrothermal method has been widely used to prepare Figure 16. TEM image of TiO2 nanotubes. Reprinted with
TiO2 nanotubes since it was introduced by Kasuga et al. in permission from Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino,
1998.153-175 Briefly, TiO2 powders are put into a 2.5-20 M T.; Niihara, K. Langmuir 1998, 14, 3160. Copyright 1998 American
NaOH aqueous solution and held at 20-110 °C for 20 h in Chemical Society.
an autoclave. TiO2 nanotubes are obtained after the products connection between the ends of the sheets, resulting in the
are washed with a dilute HCl aqueous solution and distilled formation of a tube structure. In this mechanism, the TiO2
water. They proposed the following formation process of nanotubes were formed in the stage of the acid treatment
TiO2 nanotubes.154 When the raw TiO2 material was treated following the alkali treatment. Figure 16 shows typical TEM
with NaOH aqueous solution, some of the Ti-O-Ti bonds images of TiO2 nanotubes made by Kasuga et al.153 However,
were broken and Ti-O-Na and Ti-OH bonds were formed. Du and co-workers found that the nanotubes were formed
New Ti-O-Ti bonds were formed after the Ti-O-Na and during the treatment of TiO2 in NaOH aqueous solution.161
Ti-OH bonds reacted with acid and water when the material A 3D f 2D f 1D formation mechanism of the TiO2
was treated with an aqueous HCl solution and distilled water. nanotubes was proposed by Wang and co-workers.171 It stated
The Ti-OH bond could form a sheet. Through the dehydra- that the raw TiO2 was first transformed into lamellar
tion of Ti-OH bonds by HCl aqueous solution, Ti-O-Ti structures and then bent and rolled to form the nanotubes.
bonds or Ti-O-H-O-Ti hydrogen bonds were generated. For the formation of the TiO2 nanotubes, the two-dimensional
The bond distance from one Ti to the next Ti on the surface lamellar TiO2 was essential. Yao and co-workers further
decreased. This resulted in the folding of the sheets and the suggested, based on their HRTEM study as shown in Figure
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2901

Figure 18. TEM micrographs of TiO2 nanoparticles prepared with


the solvothermal method. Reprinted with permission from Li, X.
L.; Peng, Q.; Yi, J. X.; Wang, X.; Li, Y. D. Chem.sEur. J. 2006,
12, 2383. Copyright 2006 Wiley-VCH.

synthesis of a variety of nanoparticles with narrow size


distribution and dispersity.177-179 The solvothermal method
has been employed to synthesize TiO2 nanoparticles and
nanorods with/without the aid of surfactants.177-185 For
Figure 17. (a) HRTEM images of TiO2 nanotubes. (b) Cross- example, in a typical procedure by Kim and co-workers,184
sectional view of TiO2 nanotubes. Reused with permission from TTIP was mixed with toluene at the weight ratio of 1-3:10
B. D. Yao, Y. F. Chan, X. Y. Zhang, W. F. Zhang, Z. Y. Yang, N. and kept at 250 °C for 3 h. The average particle size of TiO2
Wang, Applied Physics Letters 82, 281 (2003). Copyright 2003, powders tended to increase as the composition of TTIP in
American Institute of Physics.
the solution increased in the range of weight ratio of 1-3:
10, while the pale crystalline phase of TiO2 was not produced
17, that TiO2 nanotubes were formed by rolling up the single- at 1:20 and 2:5 weight ratios.184 By controlling the hydro-
layer TiO2 sheets with a rolling-up vector of [001] and lyzation reaction of Ti(OC4H9)4 and linoleic acid, redispers-
attracting other sheets to surround the tubes.172 Bavykin and ible TiO2 nanoparticles and nanorods could be synthesized,
co-workers suggested that the mechanism of nanotube as found by Li et al. recently.177 The decomposition of NH4-
formation involved the wrapping of multilayered nanosheets HCO3 could provide H2O for the hydrolyzation reaction, and
rather than scrolling or wrapping of single layer nanosheets linoleic acid could act as the solvent/reagent and coordination
followed by crystallization of successive layers.156 In the surfactant in the synthesis of nanoparticles. Triethylamine
mechanism proposed by Wang et al., the formation of TiO2 could act as a catalyst for the polycondensation of the Ti-
nanotubes involved several steps.176 During the reaction with O-Ti inorganic network to achieve a crystalline product and
NaOH, the Ti-O-Ti bonding between the basic building had little influence on the products’ morphology. The chain
blocks of the anatase phase, the octahedra, was broken and lengths of the carboxylic acids had a great influence on the
a zigzag structure was formed when the free octahedras formation of TiO2, and long-chain organic acids were
shared edges between the Ti ions with the formation of important and necessary in the formation of TiO2.177 Figure
hydroxy bridges, leading to the growth along the [100] 18 shows a representative TEM image of TiO2 nanoparticles
direction of the anatase phase. Two-dimensional crystalline from their study.177
sheets formed from the lateral growth of the formation of
oxo bridges between the Ti centers (Ti-O-Ti bonds) in the TiO2 nanorods with narrow size distributions can also be
[001] direction and rolled up in order to saturate these developed with the solvothermal method.177,183 For example,
dangling bonds from the surface and lower the total energy, in a typical synthesis from Kim et al., TTIP was dissolved
resulting in the formation of TiO2 nanotubes.176 in anhydrous toluene with OA as a surfactant and kept at
250 °C for 20 h in an autoclave without stirring.183 Long
dumbbell-shaped nanorods were formed when a sufficient
2.5. Solvothermal Method amount of TTIP or surfactant was added to the solution, due
The solvothermal method is almost identical to the to the oriented growth of particles along the [001] axis. At
hydrothermal method except that the solvent used here is a fixed precursor to surfactant weight ratio of 1:3, the
nonaqueous. However, the temperature can be elevated much concentration of rods in the nanoparticle assembly increased
higher than that in hydrothermal method, since a variety of as the concentration of the titanium precursor in the solution
organic solvents with high boiling points can be chosen. The increased. The average particle size was smaller and the size
solvothermal method normally has better control than hy- distribution was narrower than is the case for particles
drothermal methods of the size and shape distributions and synthesized without surfactant. The crystalline phase, diam-
the crystallinity of the TiO2 nanoparticles. The solvothermal eter, and length of these nanorods are largely influenced by
method has been found to be a versatile method for the the precursor/surfactant/solvent weight ratio. Anatase nano-
2902 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 20. TEM images of TiO2 nanowires synthesized by the


solvothermal method. From: Wen, B.; Liu, C.; Liu, Y. New J.
Chem. 2005, 29, 969 (http://dx.doi.org/10.1039/b502604k) s
Reproduced by permission of The Royal Society of Chemistry
(RSC) on behalf of the Centre National de la Recherche Scientifique
(CNRS).

Figure 19. TEM micrographs and electron diffraction patterns of


products prepared from solutions at the weight ratio of precursor/
solvent/surfactant ) 1:5:3. Reprinted from Kim, C. S.; Moon, B.
K.; Park, J. H.; Choi, B. C.; Seo, H. J. J. Cryst. Growth 2003, 257,
309, Copyright 2003, with permission from Elsevier.

rods were obtained from the solution with a precursor/


surfactant weight ratio of more than 1:3 for a precursor/
solvent weight ratio of 1:10 or from the solution with a
precursor/solvent weight ratio of more than 1:5 for a
precursor/surfactant weight ratio of 1:3. The diameter and
length of these nanorods were in the ranges of 3-5 nm and
18-25 nm, respectively. Figure 19 shows a typical TEM
image of TiO2 nanorods prepared from the solutions with
the weight ratio of precursor/solvent/surfactant ) 1:5:3.183
Similar to the hydrothermal method, the solvothermal
method has also been used for the preparation of TiO2 Figure 21. SEM morphology of TiO2 nanorods by directly
nanowires.180-182 Typically, a TiO2 powder suspension in an oxidizing a Ti plate with a H2O2 solution. Reprinted from Wu, J.
5 M NaOH water-ethanol solution is kept in an autoclave M. J. Cryst. Growth 2004, 269, 347, Copyright 2004, with
permission from Elesevier.
at 170-200 °C for 24 h and then cooled to room temperature
naturally. TiO2 nanowires are obtained after the obtained
phase only existed in heterojunctions between single-crystal
sample is washed with a dilute HCl aqueous solution and
TiO2 nanowires.180
dried at 60 °C for 12 h in air.181 The solvent plays an
important role in determining the crystal morphology.
Solvents with different physical and chemical properties can
2.6. Direct Oxidation Method
influence the solubility, reactivity, and diffusion behavior TiO2 nanomaterials can be obtained by oxidation of
of the reactants; in particular, the polarity and coordinating titanium metal using oxidants or under anodization. Crystal-
ability of the solvent can influence the morphology and the line TiO2 nanorods have been obtained by direct oxidation
crystallization behavior of the final products. The presence of a titanium metal plate with hydrogen peroxide.186-191
of ethanol at a high concentration not only can cause the Typically, TiO2 nanorods on a Ti plate are obtained when a
polarity of the solvent to change but also strongly affects cleaned Ti plate is put in 50 mL of a 30 wt % H2O2 solution
the ζ potential values of the reactant particles and the at 353 K for 72 h. The formation of crystalline TiO2 occurs
increases solution viscosity. For example, in the absence of through a dissolution precipitation mechanism. By the
ethanol, short and wide flakelike structures of TiO2 were addition of inorganic salts of NaX (X ) F-, Cl-, and SO42-),
obtained instead of nanowires. When chloroform is used, the crystalline phase of TiO2 nanorods can be controlled.
TiO2 nanorods were obtained.181 Figure 20 shows representa- The addition of F- and SO42- helps the formation of pure
tive TEM images of the TiO2 nanowires prepared from the anatase, while the addition of Cl- favors the formation of
solvothermal method.181 Alternatively, bamboo-shaped Ag- rutile.189 Figure 21 shows a typical SEM image of TiO2
doped TiO2 nanowires were developed with titanium butox- nanorods prepared with this method.186
ide as precursor and AgNO3 as catalyst.180 Through the At high temperature, acetone can be used as a good oxygen
electron diffraction (ED) pattern and HRTEM study, the Ag source and for the preparation of TiO2 nanorods by oxidizing
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2903

Figure 23. SEM images of TiO2 nanotubes prepared with anodic


Figure 22. SEM images of large-scale nanorod arrays prepared oxidation. Reprinted with permission from Varghese, O. K.; Gong,
by oxidizing a titanium with acetone at 850 °C for 90 min. From: D.; Paulose, M.; Ong, K. G.; Dickey, E. C.; Grimes, C. A. AdV.
Peng, X.; Chen, A. J. Mater. Chem. 2004, 14, 2542 (http:// Mater. 2003, 15, 624. Copyright 2003 Wiley-VCH.
dx.doi.org/10.1039/b404750h) s Reproduced by permission of The
Royal Society of Chemistry.

a Ti plate with acetone as reported by Peng and Chen.192


The oxygen source was found to play an important role.
Highly dense and well-aligned TiO2 nanorod arrays were
formed when acetone was used as the oxygen source, and
only crystal grain films or grains with random nanofibers
growing from the edges were obtained with pure oxygen or
argon mixed with oxygen. The competition of the oxygen
and titanium diffusion involved in the titanium oxidation
process largely controlled the morphology of the TiO2. With
pure oxygen, the oxidation occurred at the Ti metal and the
TiO2 interface, since oxygen diffusion predominated because
of the high oxygen concentration. When acetone was used
as the oxygen source, Ti cations diffused to the oxide surface
and reacted with the adsorbed acetone species. Figure 22
shows aligned TiO2 nanorod arrays obtained by oxidizing a
titanium substrate with acetone at 850 °C for 90 min.192
As extensively studied, TiO2 nanotubes can be obtained Figure 24. SEM images of TiO2 nanorods grown at 560 °C.
Reprinted with permission from Wu, J. J.; Yu, C. C. J. Phys. Chem.
by anodic oxidation of titanium foil.193-228 In a typical B 2004, 108, 3377. Copyright 2004 American Chemical Society.
experiment, a clean Ti plate is anodized in a 0.5% HF
solution under 10-20 V for 10-30 min. Platinum is used otherwise, it is called chemical vapor deposition (CVD). In
as counterelectrode. Crystallized TiO2 nanotubes are obtained CVD processes, thermal energy heats the gases in the coating
after the anodized Ti plate is annealed at 500 °C for 6 h in chamber and drives the deposition reaction.
oxygen.210 The length and diameter of the TiO2 nanotubes
Thick crystalline TiO2 films with grain sizes below 30 nm
could be controlled over a wide range (diameter, 15-120
as well as TiO2 nanoparticles with sizes below 10 nm can
nm; length, 20 nm to 10 µm) with the applied potential
be prepared by pyrolysis of TTIP in a mixed helium/oxygen
between 1 and 25 V in optimized phosphate/HF electro-
atmosphere, using liquid precursor delivery.230 When depos-
lytes.229 Figure 23 shows SEM images of TiO2 nanotubes
ited on the cold areas of the reactor at temperatures below
created with this method.208
90 °C with plasma enhanced CVD, amorphous TiO2 nano-
particles can be obtained and crystallize with a relatively
2.7. Chemical Vapor Deposition high surface area after being annealed at high temperatures.231
Vapor deposition refers to any process in which materials TiO2 nanorod arrays with a diameter of about 50-100 nm
in a vapor state are condensed to form a solid-phase material. and a length of 0.5-2 µm can be synthesized by metal
These processes are normally used to form coatings to alter organic CVD (MOCVD) on a WC-Co substrate using TTIP
the mechanical, electrical, thermal, optical, corrosion resis- as the precursor.232
tance, and wear resistance properties of various substrates. Figure 24 shows the TiO2 nanorods grown on fused silica
They are also used to form free-standing bodies, films, and substrates with a template- and catalyst-free MOCVD
fibers and to infiltrate fabric to form composite materials. method.233 In a typical procedure, titanium acetylacetonate
Recently, they have been widely explored to fabricate various (Ti(C10H14O5)) vaporizing in the low-temperature zone of a
nanomaterials. Vapor deposition processes usually take place furnace at 200-230 °C is carried by a N2/O2 flow into the
within a vacuum chamber. If no chemical reaction occurs, high-temperature zone of 500-700 °C, and TiO2 nanostruc-
this process is called physical vapor deposition (PVD); tures are grown directly on the substrates. The phase and
2904 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 25. SEM images of the TiO2 nanowire arrays prepared by


the PVD method. Reprinted from Wu, J. M.; Shih, H. C.; Wu, W.
T. Chem. Phys. Lett. 2005, 413, 490, Copyright 2005, with
permission from Elsevier. Figure 26. Cross-sectional SEM image of TiO2 nanowires elec-
trodeposited in AAM pores. Reprinted from Liu, S.; Huang, K.
Sol. Energy Mater. Sol. Cells 2004, 85, 125, Copyright 2004, with
morphology of the TiO2 nanostructures can be tuned with permission from Elsevier.
the reaction conditions. For example, at 630 and 560 °C
under a pressure of 5 Torr, single-crystalline rutile and pH ) 2 with a pulsed electrodeposition approach, and
anatase TiO2 nanorods were formed respectively, while, at titanium and/or its compound are deposited into the pores
535 °C under 3.6 Torr, anatase TiO2 nanowalls composed of the AAM. By heating the above deposited template at
of well-aligned nanorods were formed.233 500 °C for 4 h and removing the template, pure anatase TiO2
In addition to the above CVD approaches in preparing nanowires can be obtained. Figure 26 shows a representative
TiO2 nanomaterials, other CVD approaches are also used, SEM image of TiO2 nanowires.256
such as electrostatic spray hydrolysis,234 diffusion flame
pyrolysis,235-239 thermal plasma pyrolysis,240-246 ultrasonic 2.10. Sonochemical Method
spray pyrolysis,247 laser-induced pyrolysis,248,249 and ultronsic-
assisted hydrolysis,250,251 among others. Ultrasound has been very useful in the synthesis of a wide
range of nanostructured materials, including high-surface-
2.8. Physical Vapor Deposition area transition metals, alloys, carbides, oxides, and colloids.
The chemical effects of ultrasound do not come from a direct
In PVD, materials are first evaporated and then condensed interaction with molecular species. Instead, sonochemistry
to form a solid material. The primary PVD methods include arises from acoustic cavitation: the formation, growth, and
thermal deposition, ion plating, ion implantation, sputtering, implosive collapse of bubbles in a liquid. Cavitational
laser vaporization, and laser surface alloying. TiO2 nanowire collapse produces intense local heating (∼5000 K), high pres-
arrays have been fabricated by a simple PVD method or sures (∼1000 atm), and enormous heating and cooling rates
thermal deposition.252-254 Typically, pure Ti metal powder (>109 K/s). The sonochemical method has been applied to
is on a quartz boat in a tube furnace about 0.5 mm away prepare various TiO2 nanomaterials by different groups.257-269
from the substrate. Then the furnace chamber is pumped Yu et al. applied the sonochemical method in preparing
down to ∼300 Torr and the temperature is increased to 850 highly photoactive TiO2 nanoparticle photocatalysts with
°C under an argon gas flow with a rate of 100 sccm and anatase and brookite phases using the hydrolysis of titanium
held for 3 h. After the reaction, a layer of TiO2 nanowires tetraisoproproxide in pure water or in a 1:1 EtOH-H2O
can be obtained.254 A layer of Ti nanopowders can be solution under ultrasonic radiation.109 Huang et al. found that
deposited on the substrate before the growth of TiO2 anatase and rutile TiO2 nanoparticles as well as their mixtures
nanowires,252,253 and Au can be employed as catalyst.252 A could be selectively synthesized with various precursors
typical SEM image of TiO2 nanowires made with the PVD using ultrasound irradiation, depending on the reaction
method is shown in Figure 25.252 temperature and the precursor used.259 Zhu et al. developed
titania whiskers and nanotubes with the assistance of
2.9. Electrodeposition sonication as shown in Figure 27.269 They found that arrays
of TiO2 nanowhiskers with a diameter of 5 nm and nanotubes
Electrodeposition is commonly employed to produce a with a diameter of ∼5 nm and a length of 200-300 nm could
coating, usually metallic, on a surface by the action of be obtained by sonicating TiO2 particles in NaOH aqueous
reduction at the cathode. The substrate to be coated is used solution followed by washing with deionized water and a
as cathode and immersed into a solution which contains a dilute HNO3 aqueous solution.
salt of the metal to be deposited. The metallic ions are
attracted to the cathode and reduced to metallic form. With
the use of the template of an AAM, TiO2 nanowires can be
2.11. Microwave Method
obtained by electrodeposition.255,256 In a typical process, the A dielectric material can be processed with energy in the
electrodeposition is carried out in 0.2 M TiCl3 solution with form of high-frequency electromagnetic waves. The principal
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2905

Figure 28. SEM image of the mesoporous TiO2 film synthesized


from the acetic acid-modified precursor and autoclaved at 230 °C.
Reprinted with permission from Barbe, C. J.; Arendse, F.; Comte,
P.; Jirousek, M.; Lenzmann, F.; Shklover, V.; Grätzel, M. J. Am.
Ceram. Soc. 1997, 80, 3157. Copyright 1997 Blackwell Publishing.
Figure 27. TEM images of TiO2 nanotubes (A) and nanowhiskers
(B) prepared with the sonochemical method. From: Zhu, Y.; Li,
H.; Koltypin, Y.; Hacohen, Y. R.; Gedanken, A. Chem. Commun. surfactant templates.28,80,264,265,277-312 Barbe et al. reported the
2001, 2616 (http://dx.doi.org/10.1039/b108968b) s Reproduced by preparation of a mesoporous TiO2 film by the hydrothermal
permission of The Royal Society of Chemistry. method as shown Figure 28.80 In a typical experiment, TTIP
was added dropwise to a 0.1 M nitric acid solution under
frequencies of microwave heating are between 900 and 2450
vigorous stirring and at room temperature. A white precipitate
MHz. At lower microwave frequencies, conductive currents
formed instantaneously. Immediately after the hydrolysis, the
flowing within the material due to the movement of ionic con-
solution was heated to 80 °C and stirred vigorously for 8 h
stituents can transfer energy from the microwave field to the
for peptization. The solution was then filtered on a glass frit
material. At higher frequencies, the energy absorption is pri-
to remove agglomerates. Water was added to the filtrate to
marily due to molecules with a permanent dipole which tend
adjust the final solids concentration to ∼5 wt %. The solution
to reorientate under the influence of a microwave electric
was put in a titanium autoclave for 12 h at 200-250 °C.
field. This reorientation loss mechanism originates from the
After sonication, the colloidal suspension was put in a rotary
inability of the polarization to follow extremely rapid rever-
evaporator and evaporated to a final TiO2 concentration of
sals of the electric field, so the polarization phasor lags the
11 wt %. The precipitation pH, hydrolysis rate, autoclaving
applied electric field. This ensures that the resulting current
pH, and precursor chemistry were found to influence the
density has a component in phase with the field, and therefore
morphology of the final TiO2 nanoparticles.
power is dissipated in the dielectric material. The major
Alternative procedures without the use of hydrothermal
advantages of using microwaves for industrial processing are
processes have been reported by Liu et al.292 and Zhang et
rapid heat transfer, and volumetric and selective heating.
al.311 In the report by Liu et al., 24.0 g of titanium(IV)
Microwave radiation is applied to prepare various TiO2
n-butoxide ethanol solution (weight ratio of 1:7) was
nanomaterials.270-276 Corradi et al. found that colloidal titania
prehydrolyzed in the presence of 0.32 mL of a 0.28 M HNO3
nanoparticle suspensions could be prepared within 5 min to
aqueous solution (TBT/HNO3 ∼ 100:1) at room temperature
1 h with microwave radiation, while 1 to 32 h was needed
for 3 h. 0.32 mL of deionized water was added to the
for the conventional synthesis method of forced hydrolysis
prehydrolyzed solution under vigorous stirring and stirred
at 195 °C.270 Ma et al. developed high-quality rutile TiO2 nano-
for an additional 2 h. The sol solution in a closed vessel
rods with a microwave hydrothermal method and found that
was kept at room temperature without stirring to gel and
they aggregated radially into spherical secondary nanopartic-
age. After aging for 14 days, the gel was dried at room
les.272 Wu et al. synthesized TiO2 nanotubes by microwave
temperature, ground into a fine powder, washed thoroughly
radiation via the reaction of TiO2 crystals of anatase, rutile,
with water and ethanol, and dried to produce porous TiO2.
or mixed phase and NaOH aqueous solution under a certain
Upon calcination at 450 °C for 4 h under air, crystallized
microwave power.275 Normally, the TiO2 nanotubes had the
mesoporous TiO2 material was obtained.292
central hollow, open-ended, and multiwall structure with
Yu et al. prepared three-dimensional and thermally stable
diameters of 8-12 nm and lengths up to 200-1000 nm.275
mesoporous TiO2 without the use of any surfactants.265
Briefly, monodispersed TiO2 nanoparticles were formed
2.12. TiO2 Mesoporous/Nanoporous Materials initially by ultrasound-assisted hydrolysis of acetic acid-
In the past decade, mesoporous/nanoporous TiO2 materials modified titanium isopropoxide. Mesoporous spherical or
have been well studied with or without the use of organic globular particles were then produced by controlled conden-
2906 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

sation and agglomeration of these sol nanoparticles under


high-intensity ultrasound radiation. The mesoporous TiO2 had
a wormhole-like structure consisting of TiO2 nanoparticles
and a lack of long-range order.265
In the template method used by the Stucky
group278-280,287,295,302,306-307,313 and other groups,264,293,297,303,309
structure-directing agents were used for organizing network-
forming metal oxide species in nonaqueous solutions. These
structure-directing agents were also called organic templates.
The most commonly used organic templates were amphi-
philic poly(alkylene oxide) block copolymers, such as HO-
(CH2CH2O)20(CH2CH(CH3)O)70(CH2CH2O)20H (designated
EO20PO70EO20, called Pluronic P-123) and HO(CH2CH2O)106-
(CH2CH(CH3)O)70(CH2CH2O)106H (designated EO106PO70-
EO106, called Pluronic F-127). In a typical synthesis, poly-
(alkylene oxide) block copolymer was dissolved in ethanol.
Then TiCl4 precursor was added with vigorous stirring. The
resulting sol solution was gelled in an open Petri dish at 40
°C in air for 1-7 days. Mesoporous TiO2 was obtained after
removing the surfactant species by calcining the as-made
sample at 400 °C for 5 h in air.306 Figure 29 shows typical
TEM images of the mesoporous TiO2. Besides triblock co-
polymers as structure-directing agents, diblock polymers were
also used such as [CnH2n-1(OCH2CH2)yOH, Brij 56 (B56, n/y
) 16/10) or Brij 58 (B58, n/y ) 16/20)] by Sanchez et al.285
Other surfactants employed to direct the formation of
mesoporous TiO2 include tetradecyl phosphate (a 14-carbon
chain) by Antonelli and Ying277 and commercially available
dodecyl phosphate by Putnam and co-workers,298 cetyltri-
methylammonium bromide (CTAB) (a cationic surfac-
tant),281,283,296 the recent Gemini surfactant,294 and dodecyl-
amine (a neutral surfactant).304 Carbon nanotubes310 and
mesoporous SBA-15286 have also been used as the skeleton
for mesoporous TiO2.

2.13. TiO2 Aerogels


The study of TiO2 aerogels is worthy of special men-
tion.314-326 The combination of sol-gel processing with
supercritical drying offers the synthesis of TiO2 aerogels with
morphological and chemical properties that are not easily
achieved by other preparation methods, i.e., with high surface
area. Campbell et al. prepared TiO2 aerogels by sol-gel
synthesis from titanium n-butoxide in methanol with the
subsequent removal of solvent by supercritical CO2.315 For
a typical synthesis process, titanium n-butoxide was added
to 40 mL of methanol in a dry glovebox. This solution was
combined with another solution containing 10 mL of
methanol, nitric acid, and deionized water. The concentration
of the titanium n-butoxide was kept at 0.625 M, and the
molar ratio of water/HNO3/titanium n-butoxide was 4:0.1:
1. The gel was allowed to age for 2 h and then extracted in
a standard autoclave with supercritical CO2 at a flow rate of Figure 29. TEM micrographs of two-dimensional hexagonal
24.6 L/h, at 343 K under 2.07 × 107 Pa for 2-3 h, resulting mesoporous TiO2 recorded along the (a) [110] and (b) [001] zone
in complete removal of solvent. After extraction, the sample axes, respectively. The inset in part a is selected-area electron
diffraction patterns obtained on the image area. (c) TEM image of
was heated in a vacuum oven at 3.4 kPa and 383 K for 3 h cubic mesoporous TiO2 accompanied by the corresponding (inset)
to remove the residual solvent and at 3.4 kPa and 483 K for EDX spectrum. Reprinted with permission from Yang, P.; Zhao,
3 h to remove any residual organics. The pretreated sample D.; Margolese, D. I.; Chmelka, B. F.; Stucky, G. D. Chem. Mater.
had a brown color and turned white after calcination at 773 1999, 11, 2813. Copyright 1999 American Chemical Society.
K or above. The resulting TiO2 aerogel, after calcination at
773 K for 2 h, had a BET surface area of >200 m2/g, 600 m2/g, as compared to a surface area of 50 m2/g for TiO2
contained mesopores in the range 2-10 nm, and was of the P25.316,317 Figure 30 shows a typical SEM image of a TiO2
pure anatase form. Dagan et al. found the TiO2 aerogels aerogel with a surface area of 447 m2/g and an interpore
obtanied by using a Ti/ethanol/H2O/nitric acid ratio of 1:20: structure constructed by near uniform grains of elliptical
3:0.08 could have a porosity of 90% and surface areas of shapes with 30 nm × 50 nm axes.326
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2907

Figure 30. SEM image of a TiO2 aerogel. Reprinted with


permission from Zhu, Z.; Tsung, L. Y.; Tomkiewicz, M. J. Phys.
Chem. 1995, 99, 15945. Copyright 1995 American Chemical
Society.

2.14. TiO2 Opal and Photonic Materials


The syntheses of TiO2 opal and photonic materials have
been well studied by various groups.327-358 Holland et al.
reported the preparation of TiO2 inverse opal from the
corresponding metal alkoxides, using latex spheres as
templates.334,335 Millimeter-thick layers of latex spheres were
deposited on filter paper in a Buchner funnel under vacuum
and soaked with ethanol. Titanium ethoxide was added
dropwise to cover the latex spheres completely while suction
was applied. Typical mass ratios of alkoxide to latex were
between 1.4 and 3. After drying the composite in a vacuum
desiccator for 3 to 24 h, the latex spheres were removed by
calcination in flowing air at 575 °C for 7 to 12 h, leaving
hard and brittle powder particles with 320- to 360-nm voids.
The carbon content of the calcined samples varied from 0.4
to 1.0 wt %, indicating that most of the latex templates had
been removed from the 3D host. Figure 31 shows an
illustration of the simple synthesis of TiO2 inverse opal and
an SEM image of TiO2 inverse opals. Similar studies have
also been carried out by other researchers.327,356
Dong and Marlow prepared TiO2 inversed opals with a
skeleton-like structure of TiO2 rods by a template-directed
method using monodispersed polystyrene particles of size
270 nm.328-330,345 Infiltration of a titania precursor (Ti(i-OPr)4 Figure 31. (A) Schematic illustration of the synthesis of a TiO2
inversed opal. (B) SEM image of the TiO2 inversed opal. Reprinted
in EtOH) was followed by a drying and calcination proce- with permission from Holland, B. T.; Blanford, C.; Stein, A. Science
dure. The precursor concentration was varied from 30% to 1998, 281, 538 (http://www.sciencemag.org). Copyright 1998
100%, and the calcination temperature was tuned from 300 AAAS.
to 700 °C. A SEM picture of the TiO2 inversed opal is shown
in Figure 32.329 The skeleton structure consists of rhombo- a sapphire substrate, which was then covered with a layer
hedral windows and TiO2 cylinders forming a highly regular of gold. After removing the PS spheres with toluene, ZnO
network. The cylinders connect the centers of the former nanorods were grown using a vapor-liquid-solid process.
octahedral and tetrahedral voids of the opal. These voids form Finally, a TiO2 layer was deposited on the ZnO nanorods
a CaF2 lattice which is filled with cylindrical bonds con- by introducing TiCl4 and water vapors into the atomic layer
necting the Ca and F sites. deposition chamber at 100 °C. Figure 34 shows SEM images
Wang et al. reported their study on the large-scale of a ZnO nanorod array and the TiO2-coated ZnO nanorod
fabrication of ordered TiO2 nanobowl arrays.354 The process array.
starts with a self-assembled monolayer of polystyrene (PS) Li et al. reported the preparation of ordered arrays of TiO2
spheres, which is used as a template for atomic layer opals using opal gel templates under uniaxial compression
deposition of a TiO2 layer. After ion-milling, toluene-etching, at ambient temperature during the TiO2 sol/gel process.337
and annealing of the TiO2-coated spheres, ordered arrays of The aspect ratio was controllable by the compression degree,
nanostructured TiO2 nanobowls can be fabricated as shown R. Polystyrene inverse opal was template synthesized using
in Figure 33. silica opals as template. The silica was removed with 40 wt
Wang et al. fabricated a 2D photonic crystal by coating % aqueous hydrofluoric acid. Monomer solutions consisting
patterned and aligned ZnO nanorod arrays with TiO2.355 PS of dimethylacrylamide, acrylic acid, and methylenebisacryl-
spheres were self-assembled to make a monolayer mask on amide in 1:1:0.02 weight ratios were dissolved in a water/
2908 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 32. SEM picture of a TiO2 skeleton with a cylinder radius


of about 0.06a. a is the lattice constant of the cubic unit cell.
Reprinted from Dong, W.; Marlow, F. Physica E 2003, 17, 431,
Copyright 2003, with permission from Elsevier.

Figure 34. (A) SEM images of short and densely aligned ZnO
nanorod array on a sapphire substrate. Inset: An optical image of
the aligned ZnO nanorods over a large area. (B) SEM image of
the TiO2-coated ZnO nanorod array. Reprinted with permission from
Wang, X.; Neff, C.; Graugnard, E.; Ding, Y.; King, J. S.; Pranger,
L. A.; Tannenbaum, R.; Wang, Z. L.; Summers, C. J. AdV. Mater.
2005, 17, 2103. Copyright 2005 Wiley-VCH.

gels with correspondingly different properties can be pro-


duced. Water was completely removed from the opal
hydrogel by repeatedly rinsing it with a large amount of
ethanol. Afterward, the opal gel was put into a large amount
of tetrabutyl titanate (TBT) at ambient temperature for 24
h. The TBT-swollen opal gel was then immersed in a water/
ethanol (1:1 wt/wt) mixture for 5 h to let the TiO2 sol/gel
process proceed. Figure 35A shows the opal structure of the
gel/titania composite spheres formed. After calcination, TiO2
opal with distinctive spherical contours could be found. The
compression degree, R, was adjusted by the spacer height
when the substrates were compressed. When the substrates
were slightly compressed against each other to the extent of
producing a 20% reduction in the thickness of the composi-
tion opal, the deformation of the template-synthesized titania
spheres was not substantial (Figure 35B). When the com-
pression degree was increased to the point of reaching 35%
deformation in the opal gel, noticeably deformed titania opals
could be obtained (Figure 35C and D).

Figure 33. (A) Experimental procedure for fabricating TiO2 2.15. Preparation of TiO2 Nanosheets
nanobowl arrays. (B) Low- and high- (inset) magnification SEM
image of TiO2 nanobowl arrays. Reprinted with permission from The preparation of TiO2 nanosheets has also been explored
Wang, X. D.; Graugnard, E.; King, J. S.; Wang, Z. L.; Summers, recently.359-368 Typically, TiO2 nanosheets were synthesized
C. J. Nano Lett. 2004, 4, 2223. Copyright 2004 American Chemical by delaminating layered protonic titanate into colloidal single
Society. layers. A stoichiometric mixture of Cs2CO3 and TiO2 was
ethanol mixture (4:7 wt/wt) with total monomer content 30 calcined at 800 °C for 20 h to produce a precursor, cesium
wt %. Ethanol was used to facilitate diffusion of the titanate, Cs0.7Ti1.82500.175O4 (0: vacancy), about 70 g of
monomer solution into the inverse opal polystyrene. After which was treated with 2 L of a 1 M HCl solution at room
the inverse opal was infiltrated by the monomer solution temperature. This acid leaching was repeated three times by
containing 1 wt % of the initiator AIBN and a subsequent renewing the acid solution every 24 h. The resulting acid-
free radical polymerization at 60 °C for 3 h, a solid composite exchanged product was filtered, washed with water, and air-
resulted. The initial inverse opal polystyrene template was dried. The obtained protonic titanate, H0.7Ti1.82500.175O4‚H2O,
then removed with chloroform in a Soxhlet extractor for 12 was shaken vigorously with a 0.017 M tetrabutylammonium
h, whereupon the opal gel was formed. By using different hydroxide solution at ambient temperature for 10 days. The
compositions of the monomer solution, hole sizes, and solution-to-solid ratio was adjusted to 250 cm3 g-1. This
stacking structures of the starting inverse opal templates, opal procedure yielded a stable colloidal suspension with an
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2909

Figure 35. SEM of the TiO2 opals. (A) A gel/titania composite opal fabricated without compressing the opal gel template during the
sol/gel process. (Inset) Image of the sample after calcination at 450 °C for 3 h. (B-D) (Main panel) Oblate titania opal materials after
calcination at 450 °C for 3 h, subject to compression degree R of (B) 20%, (C) 35%, and (D) 50%. The images were taken for the fractured
surfaces containing the direction of applied compression. (Inset) Image of the same sample, but with the fracture surface perpendicular to
the direction of applied compression. From: Ji, L.; Rong, J.; Yang, Z. Chem. Commun. 2003, 1080 (http://dx.doi.org/10.1039/b300825h)
s Reproduced by permission of The Royal Society of Chemistry.

opalescent appearance. Figure 36 shows TEM and AFM octahedron shows a slight orthorhombic distortion; in anatase,
images of TiO2 nanosheets with thicknesses of 1.2-1.3 nm, the octahedron is significantly distorted so that its symmetry
which is the height of the TiO2 nanosheet with a monolayer is lower than orthorhombic. The Ti-Ti distances in anatase
of water molecules on both sides (0.70 + 0.25 × 2) thick.366 are larger, whereas the Ti-O distances are shorter than those
in rutile. In the rutile structure, each octahedron is in contact
3. Properties of TiO2 Nanomaterials with 10 neighbor octahedrons (two sharing edge oxygen pairs
and eight sharing corner oxygen atoms), while, in the anatase
3.1. Structural Properties of TiO2 Nanomaterials structure, each octahedron is in contact with eight neighbors
Figure 37 shows the unit cell structures of the rutile and (four sharing an edge and four sharing a corner). These
anatase TiO2.11 These two structures can be described in differences in lattice structures cause different mass densities
terms of chains of TiO6 octahedra, where each Ti4+ ion is and electronic band structures between the two forms of
surrounded by an octahedron of six O2- ions. The two crystal TiO2.
structures differ in the distortion of each octahedron and by Hamad et al. performed a theoretical calculation on TinO2n
the assembly pattern of the octahedra chains. In rutile, the clusters (n ) 1-15) with a combination of simulated
2910 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 36. (A) TEM of Ti1-δO24δ- nanosheets. (B and C) AFM image and height scan of the TiO2 nanosheets deposited on a Si wafer.
(D) Structural model for a hydrated TiO2 nanosheet. Closed, open, and shaded circles represent Ti atom, O atom, and H2O molecules,
respectively. All the water sites are assumed to be half occupied. Reprinted with permission from Sasaki, T.; Ebina, Y.; Kitami, Y.; Watanabe,
M.; Oikawa, T. J. Phys. Chem. B 2001, 105, 6116. Copyright 2001 American Chemical Society.

amorphous transition regime at the smallest crystallite sizes,


an anatase-baddeleyite transition regime at intermediate
crystallite sizes, and an anatase-R-PbO2 transition regime
comprising large nanocrystals to macroscopic single crystals.
Barnard et al. performed a series of theoretical studies on
the phase stability of TiO2 nanoparticles in different environ-
ments by a thermodynamic model.371-375 They found that
surface passivation had an important impact on nanocrystal
morphology and phase stability. The results showed that
surface hydrogenation induced significant changes in the
shape of rutile nanocrystals, but not in anatase, and that the
size at which the phase transition might be expected increased
Figure 37. Lattice structure of rutile and anatase TiO2. Reprinted dramatically when the undercoordinated surface titanium
with permission from Linsebigler, A. L.; Lu, G.; Yates, J. T., Jr. atoms were H-terminated. For spherical particles, the cross-
Chem. ReV. 1995, 95, 735. Copyright 1995 American Chemical over point was about 2.6 nm. For a clean and faceted surface,
Society. at low temperatures (a phase transition pointed at an average
diameter of approximately 9.3-9.4 nm for anatase nano-
annealing, Monte Carlo basin hopping simulation, and crystals), the transition size decreased slightly to 8.9 nm when
genetic algorithms methods.369 They found that the calculated the surface bridging oxygens were H-terminated, and the size
global minima consisted of compact structures, with titanium increased significantly to 23.1 nm when both the bridging
atoms reaching high coordination rapidly as n increased. For oxygens and the undercoordinated titanium atoms of the
n g 11, the particles had at least a central octahedron surface trilayer were H-terminated. Below the cross point,
surrounded by a shell of surface tetrahedra, trigonal bipyra- the anatase phase was more stable than the rutile phase.371
mids, and square base pyramids. In their study on TiO2 nanoparticles in vacuum or water
Swamy et al. found the metastability of anatase as a environments, they found that the phase transition size in
function of pressure was size dependent, with smaller water (15.1 nm) was larger than that under vacuum (9.6
crystallites preserving the structure to higher pressures.370 nm).373 In their predictions on the transition enthalpy of
Three size regimes were recognized for the pressure-induced nanocrystalline anatase and rutile, they found that thermo-
phase transition of anatase at room temperature: an anatase- chemical results could differ for various faceted or spherical
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2911

Figure 38. Morphology predicted for anatase (top), with (a) Figure 39. Enthalpy of nanocrystalline TiO2. Reprinted with
hydrogenated surfaces, (b) hydrogen-rich surface adsorbates, (c) permission from Ranade, M. R.; Navrotsky, A.; Zhang, H. Z.; Ban-
hydrated surfaces, (d) hydrogen-poor adsorbates, and (e) oxygenated field, J. F.; Elder, S. H.; Zaban, A.; Borse, P. H.; Kulkarni, S. K.;
surfaces, and for rutile (bottom), with (f) hydrogenated surfaces, Doran, G. S.; Whitfield, H. J. Proc. Natl. Acad. Sci. U.S.A. 2002,
(g) hydrogen-rich surface adsorbates, (h) hydrated surfaces, (i) 99, 6476. Copyright 2002 National Academy of Sciences, U.S.A.
hydrogen-poor adsorbates, and (j) oxygenated surfaces. Reprinted
with permission from Barnard, A. S.; Curtiss, L. A. Nano Lett. and synthetic samples. On heating concomitant with coarsen-
2005, 5, 1261. Copyright 2005 American Chemical Society. ing, the following transformations are all seen: anatase to
brookite to rutile, brookite to anatase to rutile, anatase to
nanoparticles as a function of shape, size, and degree of rutile, and brookite to rutile. These transformation sequences
surface passivation.372 Their study on anatase and rutile imply very closely balanced energetics as a function of
titanium dioxide polymorphs passivated with complete particle size. The surface enthalpies of the three polymorphs
monolayers of adsorbates by varying the hydrogen to oxygen are sufficiently different that crossover in thermodynamic
ratio with respect to a neutral, water-terminated surface stability can occur under conditions that preclude coarsening,
showed that termination with water consistently resulted in with anatase and/or brookite stable at small particle size.73,74
the lowest values of surface free energy when hydrated or However, abnormal behaviors and inconsistent results are
with a higher fraction of H on the surface on both anatase occasionally observed.
and rutile surfaces, but conversely, the surfaces generally Hwu et al. found the crystal structure of TiO2 nanoparticles
had a higher surface free energy when they had an equal depended largely on the preparation method.379 For small
ratio of H and O in the adsorbates or were O-terminated.375 TiO2 nanoparticles (<50 nm), anatase seemed more stable
They demonstrated that, under different pH conditions from and transformed to rutile at >973 K. Banfield et al. found
acid to basic, the phase transition size of a TiO2 nanoparticle that the prepared TiO2 nanoparticles had anatase and/or
varied from 6.9 to 22.7 nm, accompanied with shape changes brookite structures, which transformed to rutile after reaching
of the TiO2 nanoparticles as shown in Figure 38.374 a certain particle size.73,380 Once rutile was formed, it grew
Enyashin and Seifert conducted a theoretical study on the much faster than anatase. They found that rutile became more
structural stability of TiO2 layer modifications (anatase and stable than anatase for particle size > 14 nm.
lepidocrocite) using the density-functional-based tight bind- Ye et al. observed a slow brookite to anatase phase
ing method (DFTB).376 They found that anatase nanotubes transition below 1053 K along with grain growth, rapid
were the most stable modifications in a comparison of single- brookite to anatase and anatase to rutile transformations
walled nanotubes, nanostrips, and nanorolls. Their stability between 1053 K and 1123 K, and rapid grain growth of rutile
increased as their radii grew. The energies for all TiO2 above 1123 K as the dominant phase.381 They concluded that
nanostructures relative to the infinite monolayer followed a brookite could not transform directly to rutile but had to
1/R2 curve. transform to anatase first. However, direct transformation
Chen et al. found that severe distortions existed in Ti site of brookite nanocrystals to rutile was observed above 973
environments in the structures of 1.9 nm TiO2 nanoparticles K by Kominami et al.382
compared to those octahedral Ti sites in bulk anatase Ti using
In a later study, Zhang and Banfield found that the
K-edge XANES.377 The distorted Ti sites were likely to adopt
transformation sequence and thermodynamic phase stability
a pentacoordinate square pyramidal geometry due to the
depended on the initial particle sizes of anatase and brookite
truncation of the lattice. The distortions in the TiO2 lattice
in their study on the phase transformation behavior of
were mainly located on the surface of the nanoparticles and
nanocrystalline aggregates during their growth for isothermal
were responsible for binding with other small molecules.
and isochronal reactions.74 They concluded that, for equally
Qian et al. found that the density of the surface states on
sized nanoparticles, anatase was thermodynamically stable
TiO2 nanoparticles was likely dependent upon the details of
for sizes < 11 nm, brookite was stable for sizes between 11
the preparation methods.378 The TiO2 nanoparticles prepared
and 35 nm, and rutile was stable for sizes > 35 nm.
from basic sol were found to have more surface states than
those prepared from acidic sol based on a surface photo- Ranade et al. investigated the energetics of the TiO2
voltage spectroscopy study. polymorphs (rutile, anatase, and brookite) by high-temper-
ature oxide melt drop solution calorimetry, and they found
the energetic stability crossed over between the three phases
3.2. Thermodynamic Properties of TiO2 as shown in Figure 39.383 The dark solid line represents the
Nanomaterials phases of lowest enthalpy as a function of surface area. Rutile
Rutile is the stable phase at high temperatures, but anatase was energetically stable for surface area < 592 m2/mol (7
and brookite are common in fine grained (nanoscale) natural m2/g or >200 nm), brookite was energetically stable from
2912 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

592 to 3174 m2/mol (7-40 m2/g or 200-40 nm), and anatase


was energetically stable for greater surface areas or smaller
sizes (<40 nm). The anatase and rutile energetics cross at
1452 m2/mol (18 m2/g or 66 nm). Assuming spherical
particles, the calculated average diameters of rutile and
brookite for a 7 m2/g surface area were 201 and 206 nm,
and those of brookite and anatase for a 40 m2/g surface area
are 36 and 39 nm. These differences in particle size at the
same surface area existed because of the differences in
density. If the phase transformation took place without further
coarsening, the particle size should be smaller after the
transformation. Phase stability in a thermodynamic sense is
governed by the Gibbs free energy (∆G ) ∆H - T∆S) rather
than the enthalpy. Rutile and anatase have the same entropy.
Thus, the T∆S will not significantly perturb the sequence of
stability seen from the enthalpies. For nanocrystalline TiO2,
if the initially formed brookite had surface area > 40 m2/g,
it was metastable with respect to both anatase and rutile,
and the sequence brookite to anatase to rutile during
coarsening was energetically downhill. If anatase formed
initially, it could coarsen and transform first to brookite (at
40 m2/g) and then to rutile. The energetic driving force for
the latter reaction (brookite to rutile) was very small,
explaining the natural persistence of coarse brookite. In
contrast, the absence of coarse-grained anatase was consistent
with the much larger driving force for its transformation to
rutile.383
Li et al. found that only anatase to rutile phase transforma- Figure 40. (A) Changes in particle sizes of anatase and rutile
tion occurred in the temperature range of 973-1073 K.384 phases as a function of the annealing temperatures. (B) Arrenhius
Both anatase and rutile particle sizes increased with the plot of ln(AR/A0) vs 1/T for activation energy calculations as a
increase of temperature, but the growth rate was different, function of the size of the TiO2 nanoparticles. AR and A0 are the
as shown in Figure 40. Rutile had a much higher growth integrated diffraction peak intensity from rutile (110), and the total
rate than anatase. The growth rate of anatase leveled off at integrated anatase (101) and rutile (110) peak intensity, respectively.
Reused with permission from W. Li, C. Ni, H. Lin, C. P. Huang,
800 °C. Rutile particles, after nucleation, grew rapidly, and S. Ismat Shah, Journal of Applied Physics, 96, 6663 (2004).
whereas anatase particle size remained practically unchanged. Copyright 2004, American Institute of Physics.
With the decrease of initial particle size, the onset transition
temperature was decreased. An increased lattice compression domains reinforces the diffraction of the X-ray beam,
of anatase with increasing temperature was observed. Larger resulting in a tall narrow peak. If the crystals are randomly
distortions existed in samples with smaller particle size. The arranged or have low degrees of periodicity, the result is a
values for the activation energies obtained were 299, 236, broader peak. This is normally the case for nanomaterial
and 180 kJ/mol for 23, 17, and 12 nm TiO2 nanoparticles, assemblies. Thus, it is apparent that the fwhm of the
respectively. The decreased thermal stability in finer nano- diffraction peak is related to the size of the nanomaterials.
particles was primarily due to the reduced activation energy Figure 41 shows the XRD patterns for TiO2 nanoparticles
as the size-related surface enthalpy and stress energy of different sizes111 and for TiO2 nanorods of different
increased. lengths.129 As the nanoparticle size increased, the diffraction
peaks became narrower. In the anatase nanoparticle and
3.3. X-ray Diffraction Properties of TiO2 nanorods developed by Zhang et al., the diameters of the
Nanomaterials TiO2 nanoparticles and nanorods were both around 2.3 nm.
XRD is essential in the determination of the crystal The nanorods were elongated along the [001] direction with
structure and the crystallinity, and in the estimate of the preferred anisotropic growth along the c-axis of the anatase
crystal grain size according to the Scherrer equation lattice, which was indicated by the strong peak intensity and
narrow width of the (004) reflection and relatively lower
Kλ intensity and broader width for the other reflections. With
D) (3) an increase in length of the nanorods, the (004) diffraction
β cos θ
peak became much stronger and sharper, whereas other peaks
remained similar in shape and intensity.129 Similar results
where K is a dimensionless constant, 2θ is the diffraction have been observed by other groups.123,127,177,183
angle, λ is the wavelength of the X-ray radiation, and β is
the full width at half-maximum (fwhm) of the diffraction
peak.385 Crystallite size is determined by measuring the
3.4. Raman Vibration Properties of TiO2
broadening of a particular peak in a diffraction pattern
Nanomaterials
associated with a particular planar reflection from within the As the size of TiO2 nanomaterials decreases, the featured
crystal unit cell. It is inversely related to the fwhm of an Raman scattering peaks become broader.255,318,370,386-395 The
individual peaksthe narrower the peak, the larger the size effect on the Raman scattering in nanocrystalline
crystallite size. The periodicity of the individual crystallite TiO2 is interpreted as originating from phonon confine-
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2913

selection rule for the excitation of Raman active optical


phonons with long-range order and crystallite size.318,370 In
a perfect “infinite” crystal, conservation of phonon momen-
tum requires that only optic phonons near the Brillouin zone
(BZ) center (q ≈ 0) are involved in first-order Raman
scattering. In an amorphous material lacking long-range
order, the q vector selection rule breaks down and the Raman
spectrum resembles the phonon density of states. For
nanocrystals, the strict “infinite” crystal selection rule is
replaced by a relaxed version. This results in a range of
accessible q vectors (as large as ∆q ≈ 1/L (L diameter))
due to the uncertainty principle.
The anatase TiO2 has six Raman-active fundamentals in
the vibrational spectrum: three Eg modes centered around
144, 197, and 639 cm-1 (designated here Eg(1), Eg(2), and Eg(3),
respectively), two B1g modes at 399 and 519 cm-1 (desig-
nated B1g(1) and B1g(2d)), and an A1g mode at 513 cm-1.370
As the particle size decreases, the Raman peaks show
increased broadening and systematic frequency shifts (Figure
42).370 The most intense Eg(1) mode shows the maximum blue
shift and significant broadening with decreasing crystallite
size. A small blue shift is seen for the Eg(2) mode, while the
B1g(1) mode and the B1g(2)+A1g modes show very small blue
shifts and red shifts (the latter peak represents a combined
effect of two individual modes), respectively. Whereas the
frequency shifts for the A1g and B1g modes are not pro-
nounced, increased broadening with decreasing crystallite
size is clearly seen for these modes. The Eg(3) mode shows
significant broadening and a red shift with decreasing
crystallite size.
Choi et al. found a volume contraction effect in anatase
TiO2 nanoparticles due to increasing radial pressure as
particle size decreases, and they suggested that the effects
of decreasing particle size on the force constants and
vibrational amplitudes of the nearest neighbor bonds con-
tributed to both broadening and shifts of the Raman bands
with decreasing particle diameter.388

3.5. Electronic Properties of TiO2 Nanomaterials


The DOS of TiO2 is composed of Ti eg, Ti t2g (dyz, dzx,
and dxy), O pσ (in the Ti3O cluster plane), and O pπ (out of
the Ti3O cluster plane), as shown in Figure 43A.396 The upper
valence bands can be decomposed into three main regions:
the σ bonding in the lower energy region mainly due to O
pσ bonding; the π bonding in the middle energy region; and
O pπ states in the higher energy region due to O pπ
nonbonding states at the top of the valence bands where the
hybridization with d states is almost negligible. The contri-
bution of the π bonding is much weaker than that of the σ
Figure 41. (A) Powder XRD patterns of TiO2 samples of different bonding. The conduction bands are decomposed into Ti eg
diameters: (a) 5 nm; (b) 7 nm; (c) 13 nm. Reprinted with permission (>5 eV) and t2g bands (<5 eV). The dxy states are dominantly
from Niederberger, M.; Bartl, M. H.; Stucky, G. D. Chem. Mater. located at the bottom of the conduction bands (the vertical
2002, 14, 4364. Copyright 2002 American Chemical Society. (B) dashed line in Figure 43A). The rest of the t2g bands are
Powder XRD patterns of TiO2 samples of diameter 2.3 nm: (a) antibonding with p states. The main peak of the t2g bands is
spherical particles; (b) 16-nm nanorods; (c) 30-nm nanorods. identified to be mostly dyz and dzx states.
Reprinted with permission from Zhang, Z.; Zhong, X.; Liu, S.; Li,
D.; Han, M. Angew. Chem., Int. Ed. 2005, 44, 3466. Copyright In the molecular-orbital bonding diagram in Figure 43B,
2005 Wiley-VCH. a noticeable feature can be found in the nonbonding states
near the band gap: the nonbonding O pp orbital at the top
ment,255,318,370,386,387,395 nonstoichiometry,391,392 or internal of the valence bands and the nonbonding dxy states at the
stress/surface tension effects.390 Among these theories, the bottom of the conduction bands. A similar feature can be
most convincing is the three-dimensional confinement of seen in rutile; however, it is less significant than in anatase.397
phonons in nanocrystals.255,318,370,386,387,394,395 The phonon In rutile, each octahedron shares corners with eight neighbors
confinement model is also referred to as the spatial correla- and shares edges with two other neighbors, forming a linear
tion model or q vector relaxation model. It links the q vector chain. In anatase, each octahedron shares corners with four
2914 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 42. (A) Ambient pressure Raman spectra of anatase with an average crystallite size of 4 ( 1 nm (A), 8 ( 2 nm (B), 20 ( 8 nm
(C), and 34 ( 5 nm (D). The spectrum marked “E” is from a bulk anatase. (B) The Raman line width (fwhm) of the Eg(1) mode versus
crystallite size. Reprinted with permission from Swamy, V.; Kuznetsov, A.; Dubrovinsky, L. S.; Caruso, R. A.; Shchukin, D. G.; Muddle,
B. C. Phys. ReV. B 2005, 71, 184302/1 (http://link.aps.org/abstract/PRB/v71/p184302). Copyright 2005 by the American Physical Society.
neighbors and shares edges with four other neighbors, surrounded by six oxygen atoms in an elongated octahedral
forming a zigzag chain with a screw axis. Thus, anatase is geometry (D2d). The further splitting of the 3d levels of Ti3+
less dense than rutile. Also, anatase has a large metal-metal due to the asymmetric crystals is shown for rutile and anatase
distance of 5.35 Å. As a consequence, the Ti dxy orbitals at structures. The fine electronic structure of TiO2 can be
the bottom of the conduction band are quite isolated, while directly probed by Ti K-edge X-ray-absorption near-edge
the t2g orbitals at the bottom of the conduction band in rutile structure (XANES), and the right panel of Figure 44B
provide the metal-metal interaction with a smaller distance contains O K-edge experimental electron-energy-loss near-
of 2.96 Å. edge structure (ELNES) spectra.398
The electronic structure of TiO2 has been studied with Hwu et al. found that the crystal field splitting of
various experimental techniques, i.e., with X-ray photoelec- nanocrystal TiO2 was approximately 2.1 eV, slightly smaller
tron and X-ray absorption and emission spectroscop- than that of bulk TiO2, as shown in Figure 45A.379 Luca et
ies.379,398-405 Figure 44 shows a schematic energy level al. found that 1s f np transitions broadened as particle size
diagram of the lowest unoccupied MOs of a [TiO6]8- cluster (increased or decreased) in the postedge region in the X-ray
with Oh, D2h (rutile), and D2d (anatase) symmetry and the Ti absorption spectroscopy for TiO2 nanoparticles.403 Also, a
K-edge XANES and O K-edge ELNES spectra for rutile and clear trend in the X-ray absorption spectroscopy for different
anatase.398 The anatase structure is a tetragonally distorted sized TiO2 nanoparticles was observed, as shown in Figure
octahedral structure in which every titanium cation is 45B from the study by Choi et al.401
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2915

nanoparticle falls below the Bohr radius of the first excitation


state or becomes comparable to the de Broglie wavelength
of the charge carriers, the charge carriers begin to behave
quantum mechanically and the charge confinement leads to
a series of discrete electronic states.408 However, there is a
discrepancy in this critical size below which quantization
effects are observed for TiO2 nanomaterials with indirect
band gaps. The estimated critical diameter depends critically
on the effective masses of the charge carriers.409 Kormann
et al. estimated the excitation radii for titania particles to be
between 7.5 and 19 Å.84 Quantum confinement size effects
were observed for TiO2 nanoparticles with a small apparent
band gap blue shift (<0.1-0.2 eV) caused by quantum size
effects for spherical particles sizes down to 2 nm.58,60 Such
small effects are mainly due to the relatively high effective
mass of carriers in TiO2 and an exciton radius in the
approximate range 0.75-1.90 nm.84 On the other hand,
Serpone et al. suggested that the blue shifts in the effective
band gap of TiO2 with particle sizes of 21, 133, and 267 Å
may in fact not be a quantum confinement effect.410 Mon-
ticone et al. did an excellent study on the quatum size effects
in anatase nanoparticles and found no quantum size effect
in anatase TiO2 nanoparticles for sizes 2R g 1.5 nm, but
they did find unusual variation of the oscillator strength of
the first allowed direct transition with particle size.411

3.6. Optical Properties of TiO2 Nanomaterials


The main mechanism of light absorption in pure semi-
conductors is direct interband electron transitions. This
absorption is especially small in indirect semiconductors, e.g.,
TiO2, where the direct electron transitions between the band
centers are prohibited by the crystal symmetry. Braginsky
and Shklover have shown the enhancement of light absorp-
tion in small TiO2 crystallites due to indirect electron
transitions with momentum nonconservation at the inter-
face.412 This effect increases at a rough interface when the
share of the interface atoms is larger. The indirect transitions
are allowed due to a large dipole matrix element and a large
density of states for the electron in the valence band.
Considerable enhancement of the absorption is expected in
small TiO2 nanocrystals, as well as in porous and micro-
crystalline semiconductors, when the share of the interface
atoms is sufficiently large. A rapid increase in the absorption
takes place at low (hν < Eg + Wc, where Wc is the width of
the conduction band) photon energies. Electron transitions
to any point in the conduction band become possible when
hν ) Eg + Wc. Further enhancement of the absorption occurs
due to an increase of the electron density of states in only
Figure 43. (A) Total and projected densities of states (DOSs) of the valence band. The interface absorption becomes the main
the anatase TiO2 structure. The DOS is decomposed into Ti eg, Ti mechanism of light absorption for the crystallites that are
t2g (dyz, dzx, and dxy), O pσ (in the Ti3O cluster plane), and O pπ smaller than 20 nm.412
(out of the Ti3O cluster plane) components. The top of the valence Sato and Sakai et al. showed through calculation and
band (the vertical solid line) is taken as the zero of energy. The
vertical dashed line indicates the conduction-band minimum as a measurement that the band gap of TiO2 nanosheets was larger
guide to the eye. (B) Molecular-orbital bonding structure for anatase than the band gap of bulk TiO2, due to lower dimensionality,
TiO2: (a) atomic levels; (b) crystal-field split levels; (c) final i.e., a 3D to 2D transition, as shown in Figure 46. 360,413 From
interaction states. The thin-solid and dashed lines represent large the measurement, it was found that the lower edge of the
and small contributions, respectively. Reprinted with permission conduction band for the TiO2 nanosheet was approximately
from Asahi, R.; Taga, Y.; Mannstadt, W.; Freeman, A. J. Phys. 0.1 V higher, while the upper edge of the valence band was
ReV. B 2000, 61, 7459 (http://link.aps.org/abstract/PRB/v61/p7459).
Copyright 2000 by the American Physical Society. 0.5 V lower than that of anatase TiO2.360 The absorption of
the TiO2 nanosheet colloid blue shifted (>1.4 eV) relative
to that of bulk TiO2 crystals (3.0-3.2 eV), due to a size-
It is well-known that for nanoparticles the band gap energy quantization effect, accompanied with a strong photolumi-
increases and the energy band becomes more discrete with nescence of well-developed fine structures extending into
decreasing size.84,406,407 As the size of a semiconductor the visible light regime.362,363 The band gap energy shift, ∆Eg,
2916 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 44. (A) Schematic energy level diagram of the lowest unoccupied MOs of a [TiO6]8- cluster with Oh, D2h (rutile), and D2d (anatase)
symmetry. (B) Ti K-edge XANES and O K-edge ELNES spectra for rutile (a) and anatase (b). Reprinted with permission from Wu, Z. Y.;
Ouvrared, G.; Gressier, P.; Natoli, C. R. Phys. ReV. B 1997, 55, 10382 (http://link.aps.org/abstract/PRB/v55/p10382). Copyright 1997 by
the American Physical Society.

by exciton confinement in anisotropic two-dimensional crystallite dimensions in the parallel and perpendicular
crystallites is formulated as follows: directions with respect to the sheet, respectively. Since the

( )
first term can be ignored, the blue shift is predominantly
h2 1 1 h2 governed by the sheet thickness. The onset of a 270 nm peak
∆Eg ) + + (4)
8µxz L 2 L 2 8µyLy2 in the photoluminescence of TiO2 nanosheets was assigned
x z
to resonant luminescence. The series of peaks extending into
where h is Plank’s constant, µxz and µy are the reduced a longer wavelength region were attributed to interband levels
effective masses of the excitons, and Lx, Ly, and Lz are the generated by the intrinsic Ti site vacancies. The contrasting
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2917

sharp peaks were also attributed to the subnanometer


thickness and its uniformity.362
Bavykin et al. studied the optical absorption and photolu-
minescence of colloidal TiO2 nanotubes with internal diam-
eter in the range of 2.5-5 nm, and they found that, in spite
of the different diameters, all the TiO2 nanotubes had similar
optical properties.158 They attributed this to the complete
smearing of all 1-dimensional effects due to the large
effective mass of charge carriers in TiO2, which resulted in
an apparent 2D behavior of TiO2 nanotubes. Figure 47 shows
the absorption, photoluminescence, and luminescence excita-
tion spectra of TiO2 nanotubes of different mean diameters.158
Within the effective mass model, the energy spectrum of
2D TiO2 nanosheets can be described by eq 10, where the
“plus” and “minus” signs correspond to the conduction and
valence bands, respectively, EG is the energy gap, p is
Planck’s constant, and me and mh are the effective masses
of the electrons and holes, respectively.
EG p2k2
E(
2D ) ( ( (5)
2 2me,h

The electronic band structure of a TiO2 nanotube can be


obtained from this relation by zone-folding and is given by
a series of quasi-1D sub-bands with different indices n
(Figure 48b):
EG p2 2n 2
(
En1D )(
2
(
2me,h [
k|2 +
d ( )] (6)

This transition from the 2D to the quasi-1D energy spectrum


has a dramatic effect on the energy density of states. In the
2D case, the density of states, G2D ) mc.h/πp2, has a constant
value for energies outside the energy gap (see Figure 48c).
In the quasi-1D case, however, the density of states of each
sub-band

{ }
Figure 45. (A) Ti L2.3 absorption of nanocrystal and bulk TiO2.
me,h 1/2 Reprinted from Hwu, Y.; Yao, Y. D.; Cheng, N. F.; Tung, C. Y.;
Gn,1D(E) ) ( (7) Lin, H. M. Nanostruct. Mater. 1997, 9, 355, Copyright 1997, with
2π2p2[E - En(0)] permission from Elsevier. (B) Ti L2.3 absorption of TiO2 nano-
crystals with different sizes. Reprinted with permission from Choi,
diverges at the band edge En(0), leading to van Hove H. C.; Ahn, H. J.; Jung, Y. M.; Lee, M. K.; Shin, H. J.; Kim, S.
B.; Sung, Y. E. Appl. Spectrosc. 2004, 58, 598. Copyright 2004
singularities. The resulting density of state is formed by a Society for Applied Spectroscopy.
series of sharp peaks with long overlapping tails (Figure 48c).
The energy gap between the valence and conductance bands corresponding bulk phase.376 The valence band of both bulk
in the quasi-1D case is larger than that in the parental 2D TiO2 and their nanostructures was composed of 3d Ti-2p
material, and the difference increases with decreasing O states, and the lower part of the conduction band was
diameter of the nanotube. The change in the energy gaps formed by 3d Ti states. The differences between these
between a nanosheet and a nanotube is nanostructures were insignificant. All anatase systems were

( )
semiconductors with a wide direct band gap (∼4.2 eV), while
2p2 1 1 the lepidocrocite nanotubes were semiconductors with an
∆EG ) E1D
G - EG )
2D
+ (8)
d2 me mh indirect band gap (∼4.5 eV). Independent from the specific
topology of the titania nanostructures, the band gap ap-
In TiO2, the effective masses of electrons me can vary proached the band gap of the corresponding nanocrystals with
between 5m0 and 30m0, and the mass of holes mh is more radii of about 25 Å.376
than 3m0. With me ) 9m0 and mh ) 3m0, the difference In addition to the above investigation on the bulk electronic
between energy gaps of nanotubes with diameters 2.5 and 5 structures for various TiO2 nanomaterials, Mora-Seró and
nm is 8 meV. The energy difference between the two first Bisquert investigated the Fermi level of surface states in TiO2
peaks in the density of states G1D(E) (Figure 48) is less than nanoparticles by the nonequilibrium steady-state statistics of
24 meV for d ) 2.5 nm and 6 meV for d ) 5 nm, which are electrons.414 They found that the electrons trapped in surface
too small to be resolved in room-temperature experiments states did not generally equilibrate to the free electrons’ Fermi
due to the thermal fluctuations of kT ) 26 meV.158 level, EFn, and a distinct Fermi level for surface states, EFs,
In the theoretical study conducted by Enyashin and Seifert could be defined consistent with Fermi-Dirac statistics,
recently, the band structures for anatase nanotubes, nano- determining the surface states’ occupancy far from equilib-
strips, and nanorolls were similar to the DOS of the rium. The difference between the free electrons’ Fermi level
2918 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 46. (A) Total and partial densities of states for (a) stacked TiO2 sheets, (b) a single-layered TiO2, (c) rutile, and (d) anatase.
Reprinted with permission from Sato, H.; Ono, K.; Sasaki, T.; Yamagishi, A. J. Phys. Chem. B 2003, 107, 9824. Copyright 2003 American
Chemical Society. (B) Schematic illustration of electronic band structure: (a) TiO2 nanosheets; (b) anatase. Reprinted with permission
from Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. J. Am. Chem. Soc. 2004, 126, 5851. Copyright 2004 American Chemical Society. (C)
UV-visible spectra of (a) TiO2 sheets and (b) a film of nanosheets on a SiO2 glass substrate. The data for the colloidal suspension is
denoted by a dashed trace. Reprinted with permission from Sasaki, T.; Watanabe, M. J. Phys. Chem. B 1997, 101, 10159. Copyright 1997
American Chemical Society.

and the surface Fermi level (∆EFn - EFs) was found to applications of TiO2 nanoparticles. These fundamental
depend on the rate constants for charge transfer and detrap- processes can be expressed as follows:415
ping and could reach several hundred millielectron-
volts.414
TiO2 + hυ 98e- + h+ (9)
3.7. Photon-Induced Electron and Hole Properties
of TiO2 Nanomaterials e- + Ti(IV)O-H f Ti(III)O-H-(X) (10)
After TiO2 nanoparticles absorb, impinging photons with
energies equal to or higher than its band gap (>3.0 eV),
electrons are excited from the valence band into the unoc- h+ + Ti(IV)O-H f Ti(IV)O•-H+(Y) (11)
cupied conduction band, leading to excited electrons in the
conduction band and positive holes in the valence band.
These charge carriers can recombine, nonradiatively or 1 1
radiatively (dissipating the input energy as heat), or get h+ + O2-lattice T O2(g) + vacancy (12)
2 4
trapped and react with electron donors or acceptors adsorbed
on the surface of the photocatalyst. The competition between
these processes determines the overall efficiency for various e-| + O2,s f O2,s- (13)
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2919

O2,s- + H+ T HO2,s (14)

h+ + Ti(III)O-H- f Ti(IV)O-H (15)

e- + Ti(IV)O•-H+ f Ti(IV)O-H (16)

O2,s + Ti(IV)O•-H+ f Ti(IV)O-H + O2,s (17)

Reaction 8 is the photon absorption process. Reactions 10-


14 are photocatalytic redox pathways, whereas reactions 15-
17 represent the recombination channels. Reactions 11 and
12 are the competition pathways for holes, leading to bound
OH radicals and O vacancies, respectively. The reverse of
reaction 12 generates O adatom intermediates upon exposing
defective surfaces to O2-(g).415 Electrons and holes generated
in TiO2 nanoparticles are localized at different defect sites
on the surface and in the bulk. Electron paramagnetic
resonance (EPR) results showed that electrons were trapped
as two Ti(III) centers, while the holes were trapped as
oxygen-centered radicals covalently linked to surface tita-
nium atoms.416-419 Howe and Grätzel found that irradiation
at 4.2 K in vacuo produced electrons trapped at Ti4+ sites
within the bulk and holes trapped at lattice oxide ions
immediately below the surface, which decayed rapidly in
the dark at 4.2 K. In the presence of O2, trapped electrons
were removed and the trapped holes were stable to 77 K.
Warming to room-temperature caused loss of trapped holes
and formation of O2- at the surface.416,417 Hurum et al. found
that, upon band gap illumination, holes appeared at the
surface and preferentially recombined with electrons in
surface trapping sites for mixed-phase TiO2, such as Degussa
P25, and recombination reactions were dominated by surface
reactions that followed charge migration.419
Colombo and Bowman studied the charge carrier dynamics
of TiO2 nanoparticles with femtosecond time-resolved diffuse
reflectance spectroscopy and found a dramatic increase in
the population of trapped charge carriers within the first few
picoseconds.420,421 Skinner et al. found that the trapping time
for photogenerated electrons on 2 nm TiO2 nanoparticles in
acetonitrile by ultrafast transient absorption was about 180
fs.422 Serpone et al. found that localization (trapping) of the
electron as a Ti3+ species occurred with a time scale of about
30 ps and about 90% or more of the photogenerated electron/ Figure 47. (A) (a) Absorption spectrum and (b) luminescence
hole pairs recombined within 10 ns.409 They suggested that excitation spectrum (wavelength of emission light is 400 nm) of
colloidal TiO2 nanotubes of different mean diameters: (1) 2.5 nm;
photoredox chemistry occurring at the particle surface (2) 3.1 nm; (3) 3.5 nm; (4) 5 nm. The curves are shifted vertically
emanated from trapped electrons and trapped holes rather for clarity. (B) Photoluminescence spectra of colloidal TiO2
than from free valence band holes and conduction band nanotubes of different mean diameters: (1) 2.5 nm; (2) 3.1 nm;
electrons. Bahnemann et al. found that, in 2.4 nm TiO2 (3) 3.5 nm; (4) 5 nm. Room temperature, excitation wavelength
nanoparticles, electrons were instantaneously trapped within 237 nm, slits width 5 nm. The range of wavelengths, 455-490
the duration of the laser flash (20 ns). Deeply trapped holes nm, in the spectra is omitted due to the high signal of the second
harmonic from scattered excitation light. The curves are shifted
were rather long-lived and unreactive, and shallowly trapped vertically for clarity. Vertical lines (5) show the positions of the
holes were in a thermally activated equilibrium with free peaks in the PL spectrum of the nanosheets. Reprinted with
holes which exhibited a very high oxidation potential.423 permission from Bavykin, D. V.; Gordeev, S. N.; Moskalenko, A.
Szczepankiewicz and Hoffmann et al. found that O2 was V.; Lapkin, A. A.; Walsh, F. C. J. Phys. Chem. B 2005, 109, 8565.
an efficient scavenger of conduction band electrons at the Copyright 2005 American Chemical Society.
gas/solid interface and the buildup of trapped carriers
eventually resulted in extended surface reconstruction in- types of TiO-H stretch, shallow electron-trapping states
volving Ti-OH functionalities.415 They found that photo- produced a homogeneous electric field and were suggested
generated free conduction band electrons were coupled with not to be associated with localized structures, but rather
acoustic phonons in the lattice and their lifetimes were delocalized across the TiO2 surface.424
lengthened when dehydrated.424 The photoexcited charge Berger et al. studied UV light-induced electron-hole pair
carriers in TiO2 nanoparticles produced Stark effect intensity excitations in anatase TiO2 nanoparticles by electron para-
and wavelength shifts for surface TiO-H stretching vibra- magnetic resonance (EPR) and IR spectroscopy.425 The
tions. Although deep electron-trapping states affected certain localized states such as holes trapped at oxygen anions (O-)
2920 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 48. Schematic presentation of the transformation of the electron band structure of the nanosheet semiconductor accompanying the
formation of nanotubes: (a) band diagram of a 2-dimensional nanosheet; (b) band diagram of quasi-1-D nanotubes; (c) energy density of
states for nanosheets (G2D) and nanotubes (G1D). EG1D and EG2D are the band gaps of the 1D and 2D structures, respectively. kx and ky are
the wave vectors. Reprinted with permission from Bavykin, D. V.; Gordeev, S. N.; Moskalenko, A. V.; Lapkin, A. A.; Walsh, F. C. J. Phys.
Chem. B 2005, 109, 8565. Copyright 2005 American Chemical Society.

trapped at localized sites, giving paramagnetic Ti3+ centers,


or remained in the conduction band as EPR silent species
which may be observed by their IR absorption and that the
EPR-detected holes produced by photoexcitation were O-
species, produced from lattice O2- ions. It was also found
that, under high-vacuum conditions, the majority of photo-
excited electrons remained in the conduction band. At 298
K, all stable hole and electron states were lost.

4. Modifications of TiO2 Nanomaterials


Many applications of TiO2 nanomaterials are closely
related to its optical properties. However, the highly efficient
use of TiO2 nanomaterials is sometimes prevented by its wide
band gap. The band gap of bulk TiO2 lies in the UV regime
(3.0 eV for the rutile phase and 3.2 eV for the anatase phase),
which is only a small fraction of the sun’s energy (<10%),
Figure 49. Scheme of UV-induced charge separation in TiO2.
as shown in Figure 50.11
Electrons from the valence band can either be trapped (a) by defect Thus, one of the goals for improvement of the performance
states, which are located close to the conduction band (shallow of TiO2 nanomaterials is to increase their optical activity by
traps), or (b) in the conduction band, where they produce absorption shifting the onset of the response from the UV to the visible
in the IR region. Electron paramagnetic resonance spectroscopy region.21,426-428 There are several ways to achieve this goal.
detects both electrons in shallow traps, Ti3+, and hole centers, O-.
Reprinted with permission from Berger, T.; Sterrer, M.; Diwald, First, doping TiO2 nanomaterials with other elements can
O.; Knoezinger, E.; Panayotov, D.; Thompson, T. L.; Yates, J. T., narrow the electronic properties and, thus, alter the optical
Jr. J. Phys. Chem. B 2005, 109, 6061. Copyright 2005 American properties of TiO2 nanomaterials. Second, sensitizing TiO2
Chemical Society. with other colorful inorganic or organic compounds can
improve its optical activity in the visible light region. Third,
and electrons trapped at coordinatively unsaturated cations coupling collective oscillations of the electrons in the
(Ti3+ formation) were accessible to EPR spectroscopy. conduction band of metal nanoparticle surfaces to those in
Delocalized and EPR silent electrons in the conduction band the conduction band of TiO2 nanomaterials in metal-TiO2
may be traced by their IR absorption, which results from nanocomposites can improve the performance. In addition,
their electronic excitation within the conduction band in the the modification of the TiO2 nanomaterials surface with other
infrared region (Figure 49). They found that, during continu- semiconductors can alter the charge-transfer properties
ous UV irradiation, photogenerated electrons were either between TiO2 and the surrounding environment, thus im-
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2921

Nd3+-doped and Fe(III)-doped TiO2 nanoparticles with a


hydrothermal method and found that anatase, brookite, and
a trace of hematite coexisted at lower pH (1.8 and 3.6) when
the Fe(III) content was as low as 0.5% and the distribution
of iron ions was nonuniform between particles, but at higher
pH (6.0), the uniform solid solution of iron-titanium oxide
formed.460,463
Anpo et al. prepared TiO2 nanoparticles doped with Cr
and V ions with an ion-implantation method.466-471 Bessekh-
ouad et al. investigated alkaline (Li, Na, K)-doped TiO2
nanoparticles prepared by sol-gel and impregnation technol-
Figure 50. Solar spectrum at sea level with the sun at its zenith. ogy and found that the crystallinity level of the products was
Reprinted with permission from Linsebigler, A. L.; Lu, G.; Yates, largely dependent on both the nature and the concentration
J. T., Jr. Chem. ReV. 1995, 95, 735. Copyright 1995 American of the alkaline, with the best crystallinity obtained for Li-
Chemical Society. doped TiO2 and the lowest for K-doped TiO2.430 Cao et al.
prepared Sn4+-doped TiO2 nanoparticle films by the plasma-
enhanced CVD method and found that, after doping by Sn,
proving the performance of TiO2 nanomaterials-based de- more surface defects were present on the surface.433 Gracia
vices. et al. synthesized M (Cr, V, Fe, Co)-doped TiO2 by ion beam
induced CVD and found that TiO2 crystallized into the
4.1. Bulk Chemical Modification: Doping anatase or rutile structures depending on the type and amount
of cations present with partial segregation of the cations in
The optical response of any material is largely determined
the form of M2On after annealing.438 Wang et al. synthesized
by its underlying electronic structure. The electronic proper-
Fe(III)-doped TiO2 nanoparticles using oxidative pyrolysis
ties of a material are closely related to its chemical
of liquid-feed organometallic precursors in a radiation-
composition (chemical nature of the bonds between the atoms
frequency (RF) thermal plasma and found that the formation
or ions), its atomic arrangement, and its physical dimension
of rutile was strongly promoted with iron doping compared
(confinement of carriers) for nanometer-sized materials. The
to the anatase phase being prevalent in the undoped TiO2.246
chemical composition of TiO2 can be altered by doping.
4.1.1.2. Nonmetal-Doped TiO2 Nanomaterials. Various
Specifically, the metal (titanium) or the nonmetal (oxygen)
nonmetal elements, such as B, C, N, F, S, Cl, and Br, have
component can be replaced in order to alter the material’s
been successfully doped into TiO2 nanomaterials. C-doped
optical properties. It is desirable to maintain the integrity of
TiO2 nanomateirals have been obtained by heating titanium
the crystal structure of the photocatalytic host material and
carbide472-474 or by annealing TiO2 under CO gas flow at
to produce favorable changes in electronic structure. It
high temperatures (500-800 °C)475 or by direct burning of
appears easier to substitute the Ti4+ cation in TiO2 with other
a titanium metal sheet in a natural gas flame.476
transition metals, and it is more difficult to replace the O2-
N-doped TiO2 nanomaterials have been synthesized by
anion with other anions due to differences in charge states
hydrolysis of TTIP in a water/amine mixture and the post-
and ionic radii. The small size of the nanoparticle is beneficial
treatment of the TiO2 sol with amines426,428,477-482 or directly
for the modification of the chemical composition of TiO2
from a Ti-bipyridine complex483 or by ball milling of TiO2
due to the higher tolerance of the structural distortion than
in a NH3 water solution.484 N-doped TiO2 nanomaterials were
that of bulk materials induced by the inherent lattice strain
also obtained by heating TiO2 under NH3 flux at 500-600
in nanomaterials.426,429
°C485,486 or by calcination of the hydrolysis product of
4.1.1. Synthesis of Doped TiO2 Nanomaterials Ti(SO4)2 with ammonia as precipitator487 or by decomposition
of gas-phase TiCl4 with an atmosphere microwave plasma
4.1.1.1. Metal-Doped TiO2 Nanomaterials. Different torch488 or by sputtering/ion-implanting techniques with
metals have been doped into TiO2 nanomaterials.313,430-465 nitrogen489,490 or N2+ gas flux.491
The preparation methods of non-metal-doped TiO2 nanoma- S-doped TiO2 nanomaterials were synthesized by mixing
terials can be divided into three types: wet chemistry, high- TTIP with ethanol containing thiourea492-494 or by heating
temperature treatment, and ion implantation on TiO2 nano- sulfide powder495,496 or by using sputtering or ion-implanting
materials. Wet chemistry methods usually involve hydrolysis techniques with S+ ion flux.497-499 Different doping methods
of a titanium precursor in a mixture of water and other can induce the different valence states of the dopants. For
reagents, followed by heating. Choi et al. performed a example, the incorporated S from thiourea had S4+ or S6+
systematic study of TiO2 nanoparticles doped with 21 metal state,492-494 while direct heating of TiS2 or sputtering with
ions by the sol-gel method and found the presence of metal S+ induced the S2- anion.496-499
ion dopants significantly influenced the photoreactivity, F-doped TiO2 nanomaterials were synthesized by mixing
charge carrier recombination rates, and interfacial electron- TTIP with ethanol containing H2O-NH4F,500-502 or by
transfer rates.434 Li et al. developed La3+-doped TiO2 by the heating TiO2 under hydrogen fluoride503,504 or by spray
sol-gel process and found that the lanthanum doping could pyrolysis from an aqueous solution of H2TiF6505,506 or using
inhibit the phase transformation of TiO2, enhance the thermal ion-implanting techniques with F+ ion flux.507 Cl- and Br-
stability of the TiO2, reduce the crystallite size, and increase co-doped nanomaterials were synthesized by adding TiCl4
the Ti3+ content on the surface.442 Nagaveni et al. prepared to ethanol containing HBr.508
W, V, Ce, Zr, Fe, and Cu ion-doped anatase TiO2 nanopar-
ticles by a solution combustion method and found that the 4.1.2. Properties of Doped TiO2 Nanomaterials
solid solution formation was limited to a narrow range of 4.1.2.1. Electronic Properties of Doped TiO2 Nanoma-
concentrations of the dopant ions.448 Wang et al. prepared terials. 4.1.2.1.1. Metal-Doped TiO2 Nanomaterials. Ac-
2922 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 51. (A) Bonding diagram of TiO2. (B) DOS of the metal-doped TiO2 (Ti1-xAxO2: A ) V, Cr, Mn, Fe, Co, or Ni). Gray solid lines:
total DOS. Black solid lines: dopant’s DOS. The states are labeled (a) to (j). Reprinted from Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai,
K. J. Phys. Chem. Solids 2002, 63, 1909, Copyright 2002, with permission from Elsevier.
cording to Soratin and Schwarz’s study, the electronic states formed by the Cr (Mn) eg and O pσ states occurred within
of TiO2 can be decomposed into three parts: the σ bonding the lower CB. For Fe- and Co-doped TiO2, the localized level
of the O pσ and Ti eg states in the lower energy region; the (e) was situated 0.2 eV above the VB (or at the top of the
π bonding of the O pπ and Ti eg states in the middle energy VB for Co) due to the π antibonding of the Fe eg and O pπ
region; and the O pπ states in the higher energy region states. This level was occupied by four (or five for Co)
(Figure 51A).397,509 The bottom of the lower conduction band electrons. The Fe (Co) eg state was split into dz2 (f) and dx2-y2
(CB) consisting of the Ti dxy orbital contributes to the metal- (g) orbitals in the band gap. For Ni-doped TiO2, the π
metal interactions due to the σ bonding of the Ti t2g-Ti t2g antibonding of the Ni t2g and O pπ states was somewhat
states. At the top of the lower CB, the rest of the Ti2g states delocalized and appeared within the VB (h) due to the Ni eg
are antibonding with the O pπ states. The upper CB consists states from the dz2 and dx2-y2 orbitials situated in the band
of the σ antibonding orbitals between the O pσ and Ti eg gap. The electron densities around the dopant were large in
states. the VB and small in the CB compared to the case of pure
The electronic structures, i.e., the densities of states TiO2. The metal-O interaction strengthened, and the metal-
(DOSs), of V-, Cr-, Mn-, Fe-, Co-, and Ni-doped TiO2 were metal interaction became weak as a result of the 3d metal
analyzed by ab initio band calculations based on the density doping.
functional theory with the full-potential linearized augmented Li et al. found that 1.5 at % Nd3+-doped TiO2 nanoparticles
plane wave (FLAPW) method by Umebayashi et al. (Figure reduced the band gap by as much as 0.55 eV and that the
51B).509 They found that when TiO2 was doped with V, Cr, band gap narrowing was primarily attributed to the substi-
Mn, Fe, or Co, an electron occupied level formed and the tutional Nd3+ ions, which introduced electron states into the
electrons were localized around each dopant. As the atomic band gap of TiO2 to form the new lowest unoccupied
number of the dopant increased, the localized level shifted molecular orbital (LUMO).444 Wang and Doren found that
to lower energy. The energy of the localized level due to Nd 4f electrons changed the electronic structure of Nd-doped
Co doping was low enough that it lay at the top of the valence TiO2 into the half-metallic or the insulating ground state510
band while the other metals produced midgap states. The and that V 3d states were located at the bottom of the
electrons from the Ni dopant were somewhat delocalized, conduction band of the TiO2 host in V-doped TiO2, which
thus significantly contributing to the formation of the valence was shown to be a half-metal or an insulator from their
band with the O p and Ti 3d electrons. The states due to the theoretical studies.511
3d dopants shifted to a lower energy as the atomic number 4.1.2.1.2. Nonmetal-Doped TiO2 Nanomaterials. Recent
of the dopant increased. For Ti1-xVxO2: two localized levels theoretical and experimental studies have shown that the
occurred at 1.5 eV above the VB (a) and between the lower desired band gap narrowing of TiO2 can also be achieved
and upper CBs (b). Level a was occupied by one electron by using nonmetal dopants (refs 385, 428, 444, 489, 481,
consisting of the V t2g and O pπ states and was localized 482, 484, 503, 504, and 512-547). Asahi and co-workers
around V. Level b consisted of the V eg and O pσ states calculated the electronic band structures of anatase TiO2 with
forming the σ antibonding orbital. For Cr- and Mn-doped different substitutional dopants, including C, N, F, P, or S,
TiO2, state c was localized at 1.0 eV (0.5 eV for Mn) above using the FLAPW method in the framework of the local
the VB due to Cr (Mn) t2g and O pπ, the former of which density approximation (LDA) as shown in Figure 52.489 In
was occupied by 2 (3) electrons. The σ antibonding orbital this study, C dopant introduced deep states in the gap.489
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2923

tions of electronic properties of C-doped TiO2, found that


the bands originating from C 2p states appeared in the band
gap of TiO2; however, the mixing of C with O 2p states
was too weak to produce a significant band gap narrowing.517
In Asahi’s study, the substitutional doping of N was the
most effective in the band gap narrowing because its p states
mixed with O 2p states, while the molecularly existing
species, e.g., NO and N2 dopants, gave rise to the bonding
states below the O 2p valence bands and antibonding states
deep in the band gap (Ni and Ni+s), and were well screened
and hardly interacted with the band states of TiO2.489 Di
Valentin et al. found that, for nitrogen doping in both anatase
and rutile polymorphs, N 2p localized states were just above
the top of the O 2p valence band.512,513 In anatase, these
dopant states caused a red shift of the absorption band edge
toward the visible region, while, in rutile, an overall blue
shift was found by the N-induced contraction of the O 2p
band.512 Experimental evidence supported the statement that
nitrogen-doped TiO2 formed nitrogen-induced midgap levels
slightly above the oxygen 2p valence band.486 Lee et al., in
their first-principles density-functional LDA pseudopotential
calculations of electronic properties of N-doped TiO2, found
that the bands originating from N 2p states appeared in the
band gap of TiO2; however, the mixing of N with O 2p states
was too weak to produce a significant band gap narrowing.517
Wang and Doren found that N doping introduced some states
at the valence band edge and thus made the original band
gap of TiO2 smaller, and that a vacancy could induce some
states in the band gap region, which acted as shallow
donors.510 Nakano et al. found that, in N-doped TiO2, deep
levels located at approximately 1.18 and 2.48 eV below the
conduction band were attributed to the O vacancy state as
an efficient generation-recombination center and to the N
doping which contributed to band gap narrowing by mixing
with the O 2p valence band, respectively.523 Okato et al.
found that, at high doping levels, N was difficult to substitute
for O to contribute to the band gap narrowing, instead giving
rise to the undesirable deep-level defects.524
S dopant induced a similar band gap narrowing as
nitrogen,489 and the mixing of the sulfur 3p states with the
valence band was found to contribute to the increased width
of the valence band, leading to the narrowing of the band
gap.495,497 When S existed as S4+, replacing Ti4+, sulfur 3s
states induced states just above the O 2p valence states, and
S 3p states contributed to the conduction band of TiO2 as
shown in Figure 53A.494
When F replaced the O in the TiO2 lattice, F 2p states
were localized below the O 2p valence states without any
Figure 52. (A) Total DOSs of doped TiO2 and (B) the projected mixing with the valence or conduction band as shown in
DOSs into the doped anion sites, calculated by FLAPW, for the Figure 53B, and additional states appeared just below the
dopants F, N, C, S, and P located at a substitutional site for an O
atom in the anatase TiO2 crystal (eight TiO2 units per cell). Ni-
conduction edge, due to the electron occupied level composed
doped stands for N doping at an interstitial site, and Ni+s-doped of the t2g state of the Ti 3d orbital.507 The electronic change
stands for doping at both substitutional and interstitial sites. induced by F dopant was considered to be similar to the O
Reprinted with permission from Asahi, R.; Morikawa, T.; Ohwaki, vacancy, thus reducing the effective band gap and improving
T.; Aoki, K.; Taga, Y. Science 2001, 293, 269 (http://www- visible light photoresponse.507 Li et al. found that F doping
.sciencemag.org). Copyright 2001 AAAS. produced several beneficial effects including the creation of
Nakano et al. found three deep levels located at ap- surface oxygen vacancies, the enhancement of surface
proximately 0.86, 1.30, and 2.34 eV below the conduction acidity, and the increase of Ti3+ ions, and doped N atoms
band in C-doped TiO2, which were attributed to the intrinsic formed a localized energy state above the valence band of
nature of TiO2 for the first one and the two levels newly TiO2, whereas doped F atoms themselves had no influence
introduced by the C doping.522 In particular, the pronounced on the band structure in N-F-co-doped TiO2.519
2.34 eV band contributed to band gap narrowing by mixing 4.1.2.2. Optical Properties of Doped TiO2 Nanomate-
with the O 2p valence band.522 Lee et al., in their first- rials. 4.1.2.2.1. Optical Properties of Metal-Doped TiO2
principles density-functional LDA pseudopotential calcula- Nanomaterials. A red shift in the band gap transition or a
2924 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 53. (A) Total DOS of S-doped TiO2. Reprinted with Figure 54. (A) The UV-vis absorption spectra of TiO2 (a) and
permission from Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, Cr ion-implanted TiO2 photocatalysts (b-d). The amount of
K.; Mitsui, T.; Matsumura, M. Appl. Catal. A 2004, 265, 115, implanted Cr ions (µmol/g) was (a) 0, (b) 0.22, (c) 0.66, or (d) 1.3.
Copyright 2004, with permission from Elsevier. (B) Total DOSs (B) The UV-vis absorption spectra of TiO2 (a) and Cr ion-doped
of F-doped TiO2 calculated by FLAPW. Eg indicates the (effective) TiO2 (b′-d′) photocatalysts prepared by an impregnation method.
band gap energy. The impurity states are labeled (I) and (II). The amount of doped Cr ions (wt%) was (a) 0, (b′) 0.01, (c′) 0.1,
Reprinted from Yamaki, T.; Umebayashi, T.; Sumita, T.; Yama- (d′) 0.5, or (e′) 1. Reprinted from Anpo, M.; Takeuchi, M. J. Catal.
moto, S.; Maekawa, M.; Kawasuso, A.; Itoh, H. Nucl. Instrum. 2003, 216, 505, Copyright 2003, with permission from Elsevier.
Methods Phys. Res., Sect. B 2003, 206, 254, Copyright 2003, with
permission from Elsevier. overlapped in highly impure media. The visible light
absorption for the Cr-doped TiO2 can be attributed to a donor
visible light absorption was observed in metal-doped TiO2 transition from the Cr t2g level into the CB and the acceptor
(refs 433-435, 438, 444, 445, 448, 449, 460-463, 465, 466, transition from the VB to the Cr t2g level.
470, 509, 548, and 549). For V-, Mn-, or Fe-doped TiO2, Stucky et al. found that up to 8 mol % Eu3+ ions could be
the absorption spectra shifted to a lower energy region with doped into mesoporous anatase TiO2, and excitation of the
an increase in the dopant concentration.434,445,460 This red shift TiO2 electrons within their band gap led to nonradiative
was attributed to the charge-transfer transition between the energy transfer to the Eu3+ ions with a bright red lumines-
d electrons of the dopant and the CB (or VB) of TiO2. Metal- cence.287 The mesoporous TiO2 acted as a sensitizer.
ion doped TiO2 prepared by ion implantation with various 4.1.2.2.2. Optical Properties of Nonmetal-Doped TiO2
transition-metal ions such as V, Cr, Mn, Fe, and Ni was Nanomaterials. Nonmetal doped TiO2 normally has a color
found to have a large shift in the absorption band toward from white to yellow or even light gray, and the onset of
the visible light region, with the order of the effectiveness the absorption spectra red shifted to longer wavelengths (refs
in the red shift being V > Cr > Mn > Fe > Ni.466-471 Anpo 385, 426, 478, 483, 489, 494, 495, 497, 498, 505, 506, 512,
et al. found that the absorption band of Cr-ion-implanted 516, 518, 519, 521, and 529). In N-doped TiO2 nanomate-
TiO2 shifted smoothly toward the visible light region, with rials, the band gap absorption onset shifted 600 nm from
the extent of the red shift depending on the amount of metal 380 nm for the undoped TiO2, extending the absorption up
ions implanted as shown in Figure 54A.470 Impregnated or to 600 nm, as shown in Figure 56.426 The optical absorption
chemically Cr-ion-doped TiO2 showed no shift in the of N-doped TiO2 in the visible light region was primarily
absorption edge of TiO2; however, a new absorption band located between 400 and 500 nm, while that of oxygen-
appeared at around 420 nm as a shoulder peak due to the deficient TiO2 was mainly above 500 nm from their density-
formation of an impurity energy level within the band gap, functional theory study.520 N-F-co-doped TiO2 prepared by
with its intensity increasing with the number of Cr ions spray pyrolysis absorbs light up to 550 nm in the visible
(Figure 54B).470 light spectrum.518 The S-doped TiO2 also displayed strong
In the study by Umebayashi et al., visible light absorption absorption in the region from 400 to 600 nm.494 The red shifts
of V-doped TiO2 was due to the transition between the VB in the absorption spectra of doped TiO2 are generally
and the V t2g level.509 The holes in the VB produced an attributed to the narrowing of the band gap in the electronic
anodic photocurrent. The photoexcitation processes under structure after doping.489 C-doped TiO2 showed long-tail
visible light of V-, Cr-, and Mn-doped TiO2 are illustrated absorption spectra in the visible light region.472,543 Cl-, Br-,
in Figure 55. Photoexcitation for V-, Cr-, Mn-, and Fe-doped and Cl-Br-doped TiO2 had increased optical response
TiO2 occurred via the t2g level of the dopant. The visible compared to the case of pure TiO2 in the visible region.508
light absorption for Mn- and Fe-doped TiO2 was due to the Livraghi et al. recently found that N-doped TiO2 contained
optical transitions from the impurity band tail into the CB. single atom nitrogen impurity centers localized in the band
The Mn (Fe) t2g level was close to the VB and easily gap of the oxide which were responsible for visible light
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2925

Figure 56. Reflectance spectra of N-doped TiO2 nanoparticles and


pure TiO2 nanoparticles. Reprinted with permission from Burda,
C.; Lou, Y.; Chen, X.; Samia, A. C. S.; Stout, J.; Gole, J. L. Nano
Lett. 2003, 3, 1049. Copyright 2003 American Chemical Society.

Figure 57. IPCEλ and APCEλ curves for N-doped TiO2 and TiO2.
SE stands for the substrate/electrode (SE) interface. The action
spectra are recorded with light incident onto the SE interface.
Reprinted from Lindgren, T.; Lu, J.; Hoel, A.; Granqvist, C. G.;
Torres, G. R.; Lindquist, S. E. Sol. Energy Mater. Sol. Cells 2004,
84, 145, Copyright 2004, with permission from Elsevier.

4.1.2.3. Photoelectrical Properties of Doped TiO2 Nano-


materials. The photoelectrical properties of a material can
be measured with an “action spectrum” curve using a photo-
to-current conversion setup.385,486,497,521 In this setup, light
from a xenon lamp passing through a monochromator is
radiated onto the electrode, and the photocurrents from the
electrodes are measured as a function of wavelength.385,486,497,521
The incident photo-to-current efficiency as a function of
wavelength, IPCEλ, is called an “action spectrum”. IPCEλ
can be calculated by

hc Iph,λ
IPCEλ ) (18)
e Pλλ
Figure 55. Schematic diagram to illustrate the photoexcitation
process under visible light of metal-doped TiO2: (a) Ti1-xVxO2; where Iph,λ is the photocurrent, Pλ is the power intensity of
(b) Ti1-xFexO2; (c) Ti1-xCrxO2. Reprinted from Umebayashi, T.; the light at wavelength λ, and h, c, and e are Planck’s
Yamaki, T.; Itoh, H.; Asai, K. J. Phys. Chem. Solids 2002, 63, constant, the speed of light, and the elementary charge,
1909, Copyright 2002, with permission from Elsevier. respectively.385 The IPCEλ curve normally has a similar shape
and trend as the absorption spectrum. When the IPCEλ is
absorption with promotion of electrons from the band gap divided by the absorption, the absorbed photon-to-current
localized states to the conduction band.547 Nick Serpone efficiency (APCEλ; also called the quantum yield) is
“proposed that the commonality in all...doped titanias rests obtained.521 Figure 57 shows IPCEλ and APCEλ curves for
with formation of oxygen vacancies and the advent of color N-doped TiO2 and TiO2.521 The photoelectrochemical onset
centers...that absorb the visible light radiation, and he argued for TiO2-xNx is shifted to around 550 nm into the visible
that the red shift of the absorption edge is in fact due to region of the spectrum, and some ultraviolet (UV) efficiency
formation of the color centers.546 for TiO2-xNx is lost compared to that of TiO2, suggesting
2926 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

the TiO2-xNx has a typical photoelectrochemical behavior photocurrent increased drastically after the TiO2 nanopar-
of a material with states in the band gap which act as ticulate thin film was sensitized using CdS nanoparti-
recombination centers for light-induced charge carriers.521 cles.378,555 Sant and Kamat found that quantum size effects
In another study, the action spectrum of N-doped TiO2 also played an important role in interparticle electron transfer in
displayed a higher response in the visible region than that the CdS-TiO2 semiconductor systems in that electron
of pure TiO2.486 The photocurrent spectra for the pure and transfer from photoexcited CdS to TiO2 was found to depend
S-doped crystals showed that the photocurrent spectrum edge on the size of TiO2 nanoparticles.560 Charge transfer occurred
shifted to the low-energy region below 2.9 eV for the only when TiO2 nanoparticles were sufficiently large (>1.2
S-doped crystal, compared to 3.0 eV for pure TiO2, due to nm) that the conduction band of the nanoparticles was located
the transition of electrons across the narrowed band gap below that of CdS nanoparticles.560 Shen et al. studied
between the VB and the CB.497 nanostructured TiO2 electrodes with different nanocrystals
sizes sensitized with CdSe nanoparticles and found that
4.2. Surface Chemical Modifications photoelectrochemical currents in the visible region in the
CdSe-sensitized TiO2 nanostructured electrodes were largely
When a photocurrent is generated with light energy less dependent on both the structure and electron diffusion
than that of the semiconductor band gap, the process is coefficient of the TiO2 electrodes.556 Zaban et al. studied
known as sensitization and the light-absorbing dyes are nanocrystalline TiO2 electrodes sensitized with InP quantum
referred to as sensitizers.9,10 TiO2 is a semiconductor with a dots, and found they exhibited strong photoconduction in
wide band gap, with optical absorption in the UV region the visible region and had a photocurrent action spectrum
(<400 nm). Any materials with a narrower band gap or consistent with the absorption spectrum of the InP QDs,
absorption in the visible or infrared regime can be used as a indicating electron transfers from InP QDs into TiO2 nano-
sensitizer for TiO2 materials. These materials include inor- particles under visible light illumination.559
ganic semiconductors with narrow band gaps, metals, and
organic dyes. How efficiently the sensitized TiO2 can interact Kamat et al. recently reported the sensitization of meso-
with the light depends largely on how efficiently the scopic TiO2 films using bifunctional surface modifiers (SH-
sensitizer interacts with the light. A common and key step R-COOH) linked with CdSe nanoparticles. Upon visible light
in the photosensitization of TiO2 is the efficient charge excitation, CdSe nanoparticles injected electrons into TiO2
transfer from the excited sensitizer to TiO2, and the resulting nanocrystallites.561 The TiO2-CdSe composite exhibited a
charge separation. The match between the electronic struc- photon-to-charge carrier generation efficiency of 12% when
tures of the sensitizer and TiO2 plays a large role in this employed as a photoanode in a photoelectrochemical cell.
process, as does the structure of the interface, including the 4.2.1.2. Sensitization by Metal Nanoparticles. Ohko et
grain boundaries and bonding between the sensitizer and al. found that when the TiO2 nanoparticle films were
TiO2. Careful design is needed to avoid the charge trapping sensitized with Ag nanoparticles, the color of the film could
and recombination which eventually harm the performance be reversely switched back and forth between brownish-gray
of sensitized TiO2.9,10,550 under UV light and the color of illuminating visible light
due to the oxidation of Ag by O2 under visible light and
4.2.1. Inorganic Sensitization reduction of Ag+ under UV light.562 The color of the film
4.2.1.1. Sensitization by Narrow Band Gap Semicon- under visible light could be tuned from green to red and white
ductors. Narrow band gap semiconductors have been used by changing the size of the Ag nanoparticles due to the
as sensitizers to improve the optical absorption properties plasmon-based absorption of Ag and the dielectric confine-
of TiO2 nanomaterials in the visible light region by various ment of the TiO2 nanoparticle film matrix. Figure 58 shows
groups.551-559 The preparation method for these inorganic absorption spectra and photographs of Ag-TiO2 films.
semiconductor sensitized TiO2 nanomaterials systems is Naoi et al. found that the chromogenic properties of the
usually the sol-gel method.551-558 Hoyer et al. reported the Ag-TiO2 films could be improved by simultaneous irradia-
sensitization of a nanocrystalline TiO2 matrix by small PbS tion during Ag deposition with UV and blue lights to
nanoparticles (<2.5 nm), and they found that the photogen- suppress the formation of anisotropic Ag particles and that
erated excess electrons could be directly injected from the nonvolatilization of a color image could be achieved by
PbS to the TiO2, resulting in strong photoconductance in the removing Ag+ that was generated during the irradiation with
visible region.553 Fitzmaurice et al. found that excitation of a colored light.563 The color of the film was further found to
the sensitizer AgI on TiO2 nanoparticles resulted in a be affected by the resonance wavelengths of the Ag particles,
stabilization of electron-hole pairs with a lifetime well the TiO2 film, and the nanopores in the TiO2 film. They
beyond 100 µs and in electron migration from AgI to TiO2.551 found that the photochromism and rewritability of Ag-TiO2
Vogel et al. studied the sensitization of nanoporous TiO2 by films could be deactivated by modification of Ag nanopar-
CdS, PbS, Ag2S, Sb2S3, and Bi2S3 and found that the relative ticles with thiols to make it possible to retain color images
positions of the energetic levels at the interface between the displayed on the films, and that the deactivated properties
quantum size particles and TiO2 could be optimized for could be fully reactivated by UV irradiation (Figure 59A).564
efficient charge separation by using the size quantization Kawahara et al. proposed the mechanism of charge
effect and that the photostability of the electrodes could be separation at the interface between Ag and TiO2 nanoparticles
significantly enhanced by surface modification of the TiO2 shown in Figure 59B.565 They found that, in the multicolor
nanoparticles with CdS nanoparticles.558 photochromism of TiO2 nanoporous films loaded with
Qian et al. found from surface photovoltage spectra (SPS) photocatalytically deposited or electrodeposited and com-
measurements that the large surface state density present on mercially available Ag nanoparticles, visible light-induced
the TiO2 nanoparticles could be efficiently decreased by electron transfer from Ag to oxygen molecules played an
sensitization using CdS nanoparticles and that the slow essential role. Some of the photoexcited electrons on Ag were
photocurrent response disappeared and the steady-state transferred to oxygen molecules via TiO2 and nonexcited
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2927

Figure 59. (A) Schematic illustrations for photochromism of the


Ag-TiO2 film (a, b) and deactivation (c) and reactivation (d) of
the photochromism. From: Naoi, K.; Ohko, Y.; Tatsuma, T. Chem.
Commun. 2005, 1288 (http://dx.doi.org/10.1039/b416139d) s
Reproduced by permission of The Royal Society of Chemistry. (B)
Proposed mechanism of the charge separation at the interface
between Ag and TiO2 nanoparticles. From: Kawahara, K.; Suzuki,
Figure 58. (A) Absorption spectra of a Ag-TiO2 film by K.; Ohko, Y.; Tatsuma, T. Phys. Chem. Chem. Phys. 2005, 7, 3851
ultraviolet light irradiation and after visible light irradiation. (http://dx.doi.org/10.1039/b511489f) s Reproduced by permission
Corresponding photographs are also shown. (B) Photograph of of the PCCP Owner Societies.
multicolored spots on the Ag-TiO2 film on a glass substrate
irradiated successively with monochromatic lights. A xenon lamp nanocrystalline solar cells (DSSCs).18,246,312,568-673Organic
and an ultraviolet-cut filter (blocking light below 400 nm) was used dyes are usually transition metal complexes with low lying
with a 450 nm (blue), 530 nm (green), 560 nm (yellowish-green), excited states, such as polypyridine complexes, phthalocya-
600 nm (orange), or 650 nm (red) bandpass filter (fwhm, 10 nm), nine, and metalloporphyrins.568-673 The metal centers for the
or without any bandpass filter (white). Reprinted with permission dyes include Ru(II), Zn(II), Mg(II), Fe(II), and Al(III), while
from Ohko, Y.; Tatsuma, T.; Fujii, T.; Naoi, K.; Niwa, C.; Kubota, the ligands include nitrogen heterocyclics with delocalized
Y.; Fujishima, A. Nature Mater. 2003, 2, 29. Copyright 2003 Nature
Publishing Group. π or aromatic ring systems.
These organic dyes are normally linked to TiO2 nanopar-
Ag, and replacement of the nonexcited Ag with Pt accelerated ticle surfaces via functional groups by various interactions
the electron transport from the photoexcited Ag to oxygen between the dyes and the TiO2 nanoparticle substrate: (a)
molecules and the photochromic behavior. covalent attachment by directly linking groups of interest or
Tian and Tatsuma found that nanoporous TiO2 films via linking agents, (b) electrostatic interactions via ion
loaded with Ag and Au nanoparticles exhibit negative exchange, ion-pairing, or donor-acceptor interactions, (c)
potential changes and anodic currents in response to visible hydrogen bonding, (d) van der Waals forces, etc. Most of
light irradiation, with potential applications for photovoltaic the dyes of interest link in the first way. Groups such as
cells, photocatalysts, and plasmon sensors.566 They found that silanyl (-O-Si-), amide (-NH-(CdO)-), carboxyl (-O-
for the Au-TiO2 system photoaction spectra for open-circuit (CdO)-), and phosphonato (-O-(HPO2)-) have been
potential and short-circuit current agreed with the absorption shown to from stable linkages with the surface hydroxyl
spectrum of Au nanoparticles in the TiO2 film. After the Au groups on TiO2 substrates.610 Carboxylic and phosphonic acid
nanoparticles were photoexcited due to plasmon resonance, derivatives react with the hydroxyl groups to form esters,
charge separation occurred by the transfer of photoexcited while amide linkages are obtained via the reaction of amine
electrons from the Au particle to the TiO2 conduction band derivatives and dicyclohexyl carbodiimide on TiO2. The most
with the simultaneous transfer of compensating electrons common and successful functional groups are based on
from a donor in solution to the Au particles.567 Cozzoli et carboxylic acids. Qu and Meyer found spectroscopic evi-
al. found that, following UV illumination, TiO2 nanorods dence for ester linkages after carboxylic acids react with the
sensitized with Ag or Au nanoparticles could sustain a higher surface titanol groups dehydratively.674 Metal cyano com-
degree of conduction band electron accumulation than pure pounds in acidic solutions were found to link to TiO2 surfaces
TiO2.126 by a single cyanide ligand with a C4V symmetry, i.e., TiIV-
4.2.1.3. Organic Dye Sensitization. Organic dyes have NC-FeII(CN)5.668,669,675
been widely employed as sensitizers for TiO2 nanomaterial The interfacial charge separation between the adsorbed
to improve its optical properties, i.e., in dye-sensitized dyes and TiO2 nanomaterials involves one of three mecha-
2928 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

nisms, which differ by the nature of the donor that transfers ponent (200 ( 50 fs) and a slow component (20 ps),
the electron to the semiconductor: (1) excited state; (2) attributed to the electron injection from the initially formed
reduced state; or (3) molecule-to-particle charge-transfer and the relaxed dye excited states, respectively.577
complex.550 For complete knowledge of the charge transfer In the reduced sensitizer injection mechanism, the sensi-
between the dye sensitizers and the TiO2 nanomaterials, tizer excited state(s) is first quenched by an external donor,
please refer to other excellent reviews.10,18,19,550,676,677 Ultrafast and subsequently the reduced state of the dye, S-, transfers
electron transfer from metal-to-ligand charge transfer (MLCT) an electron across the semiconductor interface.550,688 A
excited states to anatase TiO2 is the most common category potential advantage of this mechanism is that the reduced
in dye-sensitized TiO2.572,595,606,618,619,677-683 The mechanism sensitizer is a stronger reductant than the MLCT excited state,
of the dye sensitization of TiO2 nanoparticles normally typically by 0.3-0.5 eV. Thus, sensitizers that are weak
involves the excitation of the dye and the charge transfer photoreductants may sensitize TiO2 efficiently after reductive
from the dye to TiO2 nanoparticles. The low-lying MLCT quenching. This mechanism may be exploited to produce
and ligand-centered (π-π*) excited states of these complexes large open-circuit photovoltages or enhanced light harvesting
are fairly long-lived, allowing them to participate in electron- in the near-IR regions. The observation of ultrafast electron
transfer processes. As an efficient photosensitizer, the dye injection coupled with the weak oxidizing power of the
has to meet several requirements. First, the dye should have excited sensitizers currently in use strongly suggests that an
high absorption efficiency and a wide spectral range of excited-state injection mechanism is operative in regenerative
coverage of light absorption in the visible, near-IR, and IR solar cells based on these materials. The reduced sensitizer
regions. Second, the excited states of the dye should have a injection mechanism was reported by Thompson,688 Haque,598
long lifetime and a high quantum yield. Third, the dye should and Wang.659
have matched electronic structures for the ground and excited The metal-to-particle charge-transfer mechanism involves
states with TiO2 nanoparticles to ensure the efficient charge interfacial chemistry between the compounds and the TiO2
transfer between them; that is, the energy level of the excited- surface which produces color changes, observed by Grätzel
state should be well matched to the lower bound of the and identified as molecule-to-particle charge-transfer transi-
conduction band of TiO2 to minimize energetic losses during tions.689 Metal cyanides, [M(CN)x]4- (M ) FeII, RuII, OsII,
the electron-transfer reaction.550 ReIII, MoIV, or WIV, x ) 6, 7, or 8), such as ferrocyanide,
The electron transfer from the dye to TiO2 usually is very FeII(CN)64-, bind to TiO2 through ambidentate cyano ligands.
fast, in the range of tens of femtoseconds. Hannappel et al. For example, FeII(CN)64- does not absorb light above 380
found that electron transfer from the excited electronic singlet nm, but a deep orange color with an absorption maximum
state of chemisorbed ruthenium(II) cis-di(isothiocyanato)- centered at 420 nm was observed for FeII(CN)64-/TiO2, due
bis(2,2′-bipyridyl-4,4′-dicarboxylate) into empty electronic to a MPCT complex formed between FeII(CN)64- and surface
states in a colloidal anatase TiO2 film was on the time scale Ti4+ ions, Fe(II)fTi(IV).550 The metal-to-particle charge-
of <25 fs.595 Rehm et al. found the charge injection from a transfer mechanism was consistent with the subpicosecond
surface-bound coumarin 343 to the conduction band of TiO2 infrared spectroscopy study on the FeII(CN)64-/TiO2 nano-
occurred on a time scale of ∼200 fs due to strong electronic particle by Weng et al., where a mid-infrared absorption was
coupling between the dye and TiO2 energy levels.684 The assigned to TiO2 electrons in the semiconductor.690 The
electron transport and recombination in dye-sensitized TiO2 injection rate constant could not be time-resolved with a 50-
solar cells with different electrolytes had been investigated, fs instrument response function. The MPCT was also found
including iodine-doped ionic liquids (diethylmethylsulfo- by Yang et al. in their study on Fe(bpy)(CN)42--sensitized
nium, dibutylmethylsulfonium, or 3-hexyl-1-methylimida- TiO2, where the absorption spectra were well modeled by a
zolium iodide) and an organic solvent (3-methoxypropion- sum of MLCT (Fe f bpy) and metal-to-particle (Fe(II) f
itrile with LiI, I2, and 1-methylbenzimidazole).685 The most Ti(IV)) bands. The MLCT bands were solvatochromic, while
viscous electrolytes showed a clear limitation in photocurrent the MPCT bands were not.668,669 Benkoe et al. found that
attributed to a low diffusion coefficient for the triiodide that the larger the TiO2 particle and the better its overall
transports positive charge to the counterelectrode. The crystallinity, the faster the process of electron injection from
electron transport of the solar cells appeared to be dominated the dye fluorescein 27 to the anatase TiO2 film.578 Haque et
by the properties of the nanostructured TiO2 film, and the al. found that a supramolecule dye with a remarkably long-
electron lifetime depended on the type of cation used in the lived (4 s) charge-separation state could be obtained by
ionic liquid. Bulky, less absorptive cations seem to give controlling the spatial separation between the cation center
longer lifetimes. Schwarzburg et al. found a time constant of the dye and the electrode surface.597 The dyes were Ru-
of 13 fs for electron transfer from the excited singlet state (II) complexes containing carboxylated polypyridyl chro-
of the chromophore perylene bonded to the surface via a mophores and a bipyridyl ligand with aromatic amine-based
carboxyl group into anatase TiO2.686 The electron-transfer electron donor substituents.597
time of perylene became much longer (3.8 ps) at a distance The kinetics and mechanisms of the injections, transport, re-
of about 1.3 nm. Wenger et al. found that carefully controlled combination, and photovoltaic properties of electrons in nano-
deposition of Ru(II) complex dye molecules onto nanocrys- structured TiO2 solar cells have been thoroughly discussed
talline TiO2 consistently yielded monophasic injection dy- in recent reviews676,691-694 and will only be briefly mentioned
namics with a time constant shorter than 20 fs.687 The below. Considerable effort has been devoted to the kinetics
anchoring of ruthenium dye {(C4H9)4N}[Ru(Htc-terpy)- and energetics of transport and recombination in dye-sensi-
(NCS)3] (tc-terpy ) 4,4′,4′′-tricarboxy-2,2′:6′,2′′-terpyridine), tized solar cells with various techniques, such as intensity
the so-called black dye, onto nanocrystalline TiO2 films modulated photocurrent spectroscopy (IMPS),640,695-702 inten-
occurs by a bidentate binuclear coordination mode.577 The sity modulated photovoltage spectroscopy (IMVS),639,700,703,704
electron injection process from the dye excited state into the electrical impedance spectroscopy (EIS),700,705-710 transient
TiO2 conduction band was biexponential with a fast com- photocurrent,695,706,711-718 and transient photovoltage.715,719 For
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2929

example, Searson and Cao studied the photocurrent response imparted with antifogging functions on various glass prod-
of dye-sensitized, porous nanocrystalline TiO2 cells with ucts, i.e., mirrors and eyeglasses, having superhydrophilic
photocurrent transient measurements and intensity-modulated or superhydrophobic surfaces.332,732-734 For example, Feng
photocurrent spectroscopy and found that the electron et al. found that reversible superhydrophilicity and super-
transport in the TiO2 film can be fitted with a diffusion model hydrophobicity could be switched back and forth for TiO2
where the diffusion coefficient for electrons in the particle nanorod films.142 When the TiO2 nanorod films were
network was a function of the light intensity.695 Using irradiated with UV light, the photogenerated hole reacted
intensity modulated photocurrent spectroscopy, Vanmaekele- with lattice oxygen to form surface oxygen vacancies. Water
bergh et al. found that the electronic transport was controlled molecules kinetically coordinated to these oxygen vacancies,
by trapping and detrapping of photogenerated electrons in and the spherical water droplet filled the grooves along the
interfacial band gap states, distributed in energy, and that nanorods and spread out on the film with a contact angle of
the localization time of a trapped electron was controlled about 0°, resulting in superhydrophilic TiO2 films. After the
by the steady-state light intensity and interfacial kinetics.696,697 hydroxy group adsorption, the surface transformed into an
Peter et al. recently found that the electron transport in dye- energetically metastable state. When the films were placed
sensitized nanocrystalline solar cells appeared to be a slow in the dark, the adsorbed hydroxy groups were gradually
diffusion-controlled process, attributed to multiple trapping replaced by atmospheric oxygen, and the surface evolved
at energy levels distributed exponentially in the band gap of back to its original state. The surface wettability converted
the nanocrystalline TiO2.720 from superhydrophilic to superhydrophobic.142 Stain-proof-
Frank et al. summarized that the electron motion is ing, self-cleaning properties can also be bestowed on many
essentially ambipolarly diffusional and the morphology and different types of surfaces due to the superhydrophilic or
defect structure of the TiO2 film had a strong influence on superhydrophobic surfaces.735-744 TiO2 nanomaterials have
electron transport.676 The recombination predominates at the also been used as sensors for various gases and humidity
interface and depends on the spatial region of photoinjected due to the electrical or optical properties which change upon
charge buildup in the cell, the redox electrolyte, and the adsorption.745-751
surface properties of both the TiO2 nanoparticle film and One of the most important research areas for future clean
the TCO substrate. For the recombination, two mechanisms energy applications is to look for efficient materials for the
assume either a dismutation reaction or an interfacial production of electricity and/or hydrogen. When sensitized
electron-transfer reaction as rate-limiting, while the third with organic dyes or inorganic narrow band gap semicon-
mechanism states electron transport limits recombination.676 ductors, TiO2 can absorb light into the visible light region
The spatial location of the traps limiting electron transport and convert solar energy into electrical energy for solar cell
in nanocrystalline TiO2 has been a long standing issue. These applications.28,30,752 For example, an overall solar to current
traps have been speculated to locate either at the particle conversion efficiency of 10.6% has been reached by the
surface,640,721,722 in the bulk of the particles,723 or at inter- group led by Grätzel with DSSC technology.31 TiO2 nano-
particle grain boundaries.568 Kopidakis et al. recently inves- materials have been widely studied for water splitting and
tigated the dependence of the electron diffusion coefficient hydrogen production due to their suitable electronic band
and the photoinduced electron density on the internal surface structure given the redox potential of water.198,475,476,753-770
area of TiO2 nanoparticle films in dye-sensitized solar cells Another application of TiO2 nanomaterials when sensitized
by photocurrent transient measurements.724 They found that with dyes or metal nanoparticles is to build photochromic
the density of electron traps in the films changed in direct devices.562,565,771-777 Of course, one of the many applications
proportion with the internal surface area, which was varied of TiO2 nanomaterials is the photocatalytic decomposition
by altering the average particle size of the films, and the of various pollutants.
scaling of the electron diffusion coefficient with the internal
surface area. They suggested that the traps were located 5.1. Photocatalytic Applications
predominately at the surface of TiO2 particles instead of in
the bulk of the particles or at interparticle grain boundaries, TiO2 is regarded as the most efficient and environmentally
and that surface traps limited transport in TiO2 nanoparticle benign photocatalyst, and it has been most widely used for
films. Kopidakis et al. found that the traps were located photodegradation of various pollutants.121,127,132,430,442,778-822
predominately at the surface of TiO2 particles instead of in TiO2 photocatalysts can also be used to kill bacteria, as has
the bulk of the particles or at interparticle grain boundaries been carried out with E. coli suspensions.793,799 The strong
and that surface traps limited transport in TiO2 nanoparticle oxidizing power of illuminated TiO2 can be used to kill tumor
films in the dye-sensitized TiO2 solar cell.724 cells in cancer treatment.782,785,820,823-825
The photocatalytic reaction mechanisms are widely stud-
5. Applications of TiO2 Nanomaterials ied.7,12,20,33,406 The principle of the semiconductor photocata-
lytic reaction is straightforward. Upon absorption of photons
The existing and promising applications of TiO2 nanoma- with energy larger than the band gap of TiO2, electrons are
terials include paint, toothpaste, UV protection, photoca- excited from the valence band to the conduction band,
talysis, photovoltaics, sensing, and electrochromics as well creating electron-hole pairs. These charge carriers migrate
as photochromics. TiO2 nanomaterials normally have elec- to the surface and react with the chemicals adsorbed on the
tronic band gaps larger than 3.0 eV and high absorption in surface to decompose these chemicals. This photodecom-
the UV region. TiO2 nanomaterials are very stable, nontoxic, position process usually involves one or more radicals or
and cheap. Their optical and biologically benign properties intermediate species such as •OH, O2-, H2O2, or O2, which
allow them to be suitable for UV protection applications.725-730 play important roles in the photocatalytic reaction mecha-
A surface is defined as superhydrophilic or superhydro- nisms. The photocatalytic activity of a semiconductor is
phobic if the water-surface contact angle is larger than 130° largely controlled by (i) the light absorption properties, e.g.,
or less than 5°, respectively.731 TiO2 nanomateirals can be light absorption spectrum and coefficient, (ii) reduction and
2930 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

oxidation rates on the surface by the electron and hole, (iii)


and the electron-hole recombination rate. A large surface
area with a constant surface density of adsorbents leads to
faster surface photocatalytic reaction rates. In this sense, the
larger the specific surface area, the higher the photocatalytic
activity is. On the other hand, the surface is a defective site;
therefore, the larger the surface area, the faster the recom-
bination. The higher the crystallinity, the fewer the bulk
defects, and the higher the photocatalytic activity is. High-
temperature treatment usually improves the crystallinity of
TiO2 nanomaterials, which in turn can induce the aggregation
of small nanoparticles and decrease the surface area. Judging
from the above general conclusions, the relation between the
physical properties and the photocatalytic activities is
complicated. Optimal conditions are sought by taking these Figure 60. Photocatalytic properties of mesoporous TiO2 samples
considerations into account and may vary from case to case.20 as prepared and calcined at different temperature as well as TiO2
P25 nanoparticles (RB, c0 ) 1.0 × 10-5 M, pH ) 6.0) under UV-
5.1.1. Pure TiO2 Nanomaterials: First Generation light radiation. Reprinted with permission from Peng, T.; Zhao,
D.; Dai, K.; Shi, W.; Hirao, K. J. Phys. Chem. B 2005, 109, 4947.
As the size of the TiO2 particles decreases, the fraction of Copyright 2005 American Chemical Society.
atoms located at the surface increases with higher surface
area to volume ratios, which can further enhance the catalytic 60 shows the photocatalytic properties of mesoporous TiO2
activity. The increase in the band gap energy with decreasing samples as prepared and calcined at different temperatures
nanoparticle size can potentially enhance the redox potential compared to those of TiO2 P25 nanoparticles. All mesopo-
of the valence band holes and the conduction band electrons, rous TiO2 showed better activity than Deguessa P25 TiO2.
allowing photoredox reactions, which might not otherwise The optimum reactivity was obtained with the sample
proceed in bulk materials, to occur readily. One disadvantage calcined at 400 °C, and the photoactivity gradually decreased
of TiO2 nanoparticles is that they can only use a small with further increases in calcination temperature.
percentage of sunlight for photocatalysis. Practically, there Yang et al. found that TiO2 nanotubes treated with H2-
exists an optimal size for a specific photocatalytic reaction. SO4 solutions showed photocatalytic activity on degradation
Anpo et al. investigated the photocatalytic activity of TiO2 of acid orange II in the following order: TiO2 nanotubes
nanoparticles on hydrogenation reactions of CH3CCH with treated with 1.0 mol/L H2SO4 solution > TiO2 nanotubes
H2O, and they found the activity increased as the diameter treated with 0.2 mol/L H2SO4 solution > untreated TiO2
of the TiO2 particles decreased, especially below 10 nm.406 nanotubes > TiO2 nanoparticles, since TiO2 nanotubes treated
They suggested that the dependence of the yields on the with H2SO4 were composed of smaller particles and had
particle size arose from the differences in the chemical higher specific surface areas.818
reactivity and not from the physical properties of these TiO2 aerogels were also suggested as promising candidates
catalysts. for photocatalysts.316,317,319 Degan et al. prepared TiO2
Wang et al. found that there was an optimal size for TiO2 aerogels with a porosity of 90% and surface areas of 600
nanoparticles for maximum photocatalytic efficiency in the m2/g, and they found that the photodegradation of salicylic
decomposition of chloroform.815 They observed an improve- acid on TiO2 aerogels, after 1 h of near-UV illumination,
ment in activity when the particle size was decreased from was about 10 times faster than that on the Degussa TiO2.316,317
21 to 11 nm, but the activity decreased when the size was Figure 61 shows photodegradation profiles for the aerogel
reduced further to 6 nm. They concluded that for this before and after annealing, as compared to the commercial
particular reaction the optimum particle size was about 10 Degussa P25 powder.
nm. In large TiO2 nanoparticles, bulk recombination of the
charge carriers was the dominant process, which could be 5.1.2. Metal-Doped TiO2 Nanomaterials: Second
reduced by a decrease in particle size; as the particle size Generation
was lowered below a certain limit, surface recombination Over the past decades, metal-doped TiO2 nanomaterials
processes became dominant, since most of the electrons and have been widely studied for improved photocatalytic
holes were generated close to the surface and surface performance on the degradation of various organic pollutants,
recombination was faster than interfacial charge carrier i.e., under visible light irradiation (refs 21, 430, 433-435,
transfer processes.826 444, 446, 450-452, 455-457, 490, 515, 548, 810, 827-
Chae et al. studied the photocatalytic activity of four sizes 836). Choi et al. conducted a systematic study on the
of TiO2 nanoparticles on the decomposition of 2-propanol, photocatalytic activity of TiO2 nanoparticles doped with 21
and they found that 7-nm particles showed 1.6 times better transition metal elements on the oxidation of CHCl3 and the
photocatalytic activity than TiO2 P25 and that 15- and 30- reduction of CCl4 and found that the photocatalytic activity
nm particles showed lower photocatalytic efficiencies.132 was related to the electron configuration of the dopant ion
Mesoporous TiO2, TiO2 nanorods, and nanotubes have in that dopant ions with closed electron shells had little or
been demonstrated to have high photocatalytic performance no effect on the activity.434,435 Doping with Fe3+, Mo5+, Ru3+,
under suitable conditions.127,187,265,281,296,818 Peng et al. pre- Os3+, Re5+, V4+, and Rh3+ at 0.1-0.5 at % significantly
pared mesoporous TiO2 with a high specific surface area, increased the photoreactivity, while Co3+ and Al3+ doping
which showed significant activity on the oxidation of decreased the photoreactivity. The presence of metal ion
Rhodamine B due to the large surface area, small crystal dopants in the TiO2 matrix significantly influenced the charge
size, and well-crystallized anatase mesostructure.296 Figure carrier recombination rates and interfacial electron-transfer
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2931

Figure 61. Photodegradation profiles of salicylic acid on annealed


(Ela) and nonannealed (El) TiO2 aerogels as compared to a
commercial Degussa P25. Reprinted with permission from Dagan,
G.; Tomkiewicz, M. J. Phys. Chem. 1993, 97, 12651. Copyright
1993 American Chemical Society.

rates. The photoreactivity of doped TiO2 appeared to be a


complex function of the dopant concentration, the energy
level of dopants within the TiO2 lattice, their d electronic
configurations, the distribution of dopants, the electron donor
Figure 62. Variation of phenol concentration with reaction time
concentrations, and the light intensity. under (A) UV and (B) visible light: (a) pure TiO2 catalyst; (b)
Sn4+ ion-doped TiO2 nanoparticle films prepared by the Sn4+-doped TiO2. From: Cao, Y.; Yang, W.; Zhang, W.; Liu, G.;
plasma-enhanced CVD method displayed a higher photo- Yue, P. New J. Chem. 2004, 28, 218 (http://dx.doi.org/10.1039/
catalytic activity for photodegradation of phenol than pure b306845e) s Reproduced by permission of The Royal Society of
TiO2 under both UV and visible light, and the Sn4+ dopant Chemistry (RSC) on behalf of the Centre National de la Recherche
Scientifique (CNRS).
was found profitable to the separation of photogenerated
carriers under both UV and visible light excitation.433 Figure
62 shows the photocatalytic decomposition of phenol with when the doping ions were in the shallow surface, the doping
reaction time under UV and visible light using Sn4+-doped was beneficial, while, in the deep bulk, the doping was
TiO2 nanoparticles as photocatalyst.433 detrimental.451
Fe-doped nanocrystalline TiO2 was shown to display However, not all the metal-doped TiO2 nanomaterials
higher photocatalytic activity with lower Fe content (optimal showed higher photocatalytic activities than pure TiO2
0.05% mass fraction) than TiO2 in the treatment of paper- nanomaterials. Martin found V-doped TiO2 nanoparticles had
making wastewater,837 and it was shown to be more efficient reduced photocatalytic activity on the photooxidation of
in the photoelectrocatalytic disinfection of E. coli than pure 4-chlorophenol compared to pure TiO2 nanoparticles. Va-
TiO2.827 V-doped TiO2 photocatalyst photooxidized ethanol nadium appeared to reduce the photoreactivity of TiO2 by
under visible radiation and had comparable activity under promoting charge-carrier recombination with electron trap-
UV radiation to that of pure TiO2.548 Pt4+ ion-doped TiO2 ping at VO2+ centers or with hole trapping at V4+ impurity
nanoparticles exhibited higher visible light photocatalytic centers, which shunted charge carriers away from the solid/
activities on the degradations of dichloroacetate and 4-chlo- solution interface.446 Hermann et al. found that although Cr-
rophenol,830 and Ag-TiO2 nanocatalysts displayed enhanced doped (0.85 atomic %) TiO2 absorbed in the visible region,
photocatalytic activity in the degradation of 2,4,6-trichlo- its activity for oxidation of oxalic acid, propene, and
rophenol due to a better separation of photogenerated charge 2-propanol and for O isotope exchange was null under visible
carriers and improved oxygen reduction inducing a higher illumination and was smaller under UV light than that of
extent of degradation of atoms.809 pure TiO2, due to an increase in electron-hole recombination
at the Cr3+ ion sites.440 Luo et al. reported that the
Wei et al. synthesized La- and N-co-doped TiO2 nano-
photoactivity of TiO2 doped with 1.5 mol % Mo, 1 mol %
particles with superior catalytic activity under visible light,
V, 0.1 mol % V plus 1 mol % Al, or 0.1 mol % V plus 1
where N doping was responsible for the band gap narrowing
mol % Pb decreased, since the d electrons of Mo(4d) and
of TiO2 and La3+ doping prevented the aggregation of
V(3d), as majority carriers in TiO2, could effectively quench
nanoparticles.833 Chang et al. reported Cr- and N-co-doped
the high-energy photogenerated holes at the impurity levels
TiO2 nanomaterials with visible light absorbance generally
introduced by doping within the band gap of TiO2.445
led to a reduction in photocatalytic efficacy in the decolori-
zation of methylene blue, except at the low nitrogen doping 5.1.3. Nonmetal-Doped TiO2 Nanomaterials: Third
concentration.490 Bessekhouad et al. found that low concen- Generation
tration alkaline (Li, Na, K)-doped TiO2 nanoparticles were
promising materials for organic pollutants degradation.430 Nonmetal-doped TiO2 nanomaterials have been regarded
Peng et al. found that in Be2+-doped TiO2 nanomaterials, as the third generation photocatalyst. Various nonmetal-
2932 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

different photocatalytic activity under visible light due the


different carrier behavior in these samples.849
A noticeable photocatalytic activity on decompositions of
methylene blue and isopropanal in the visible region was
demonstrated for C-doped TiO2 made from a TiC precur-
sor.472,473 C-doped TiO2 made by pyrolyzing Ti metal in a
natural gas flame displayed a much higher photoactivity in
water splitting than pure TiO2.476 C-doped TiO2 nanoparticles
also displayed high photoactivity in degradation of trichlo-
roacetic acid under visible light.474
Yu et al. found that F-doped TiO2 showed higher photo-
catalytic activity on the oxidation of acetone into CO2 than
did Degeussa P25 in the photodecomposition study of
acetone under proper preparation conditions.502 N/F-doped
Figure 63. Photocatalytic properties of TiO2-xNx and TiO2 based
on decomposition rates [measuring the change in absorption of the
TiO2 nanomaterials had high visible light photocatalytic
reference light (∆abs)] of methylene blue as a function of the cutoff activities for decompositions of both acetaldehyde and
wavelength of the optical high-path filters under fluorescent light. trichloroethylene due to the creation of surface oxygen
The inset shows the decomposition rates of methylene blue in the vacancies rather than the improvement of optical absorption
aqueous solution under visible light as a function of the ratio of properties.505,506,518,519 Luo et al. found that chlorine- and
the decomposed area in the XPS spectra with the peak at 396 eV bromine-co-doped TiO2 displayed a much higher photocata-
to the total area of N 1s. The total N concentrations were 1.0 atom lytic activity than chlorine- or bromine-doped TiO2.508
% (a), 1.1 atom % (b), 1.4 atom % (c), 1.1 atom % (d), and 1.0
atom % (e). Reprinted with permission from Asahi, R.; Morikawa,
T.; Ohwaki, T.; Aoki, K.; Taga, Y. Science 2001, 293, 269 (http:// 5.2. Photovoltaic Applications
www.sciencemag.org). Copyright 2001 AAAS.
5.2.1. The TiO2 Nanocrystalline Electrode in DSSCs
doped TiO2 nanomaterials have been widely studied for their Photovoltaics based on TiO2 nanocrystalline electrodes
visible light photocatalytic activities (refs 21, 385, 426, 428, have been widely studied.9,28-32 A schematic presentation
452, 472-474, 481-487, 490, 492-494, 496, 505, 518, 520, of the structure and operating principles of the DSSC is given
524, 525, 527, 532, 533, 802, 838-847). Nonmetal-doped in Figure 64. At the heart of the system is a nanocrystalline
TiO2 nanomaterials have been demonstrated to have im- mesoporous TiO2 film with a monolayer of the charge-
proved photocatalytic activities compared to those for pure transfer dye attached to its surface. The film is placed in
TiO2 nanomaterials, especially in the visible light re- contact with a redox electrolyte or an organic hole conductor.
gion.426,428,485,489,833,848 Photoexcitation of the dye injects an electron into the
Figure 63 shows the decomposition of methylene blue conduction band of TiO2. The electron can be conducted to
using N-doped TiO2 as measured by Asahi and co-wokers.489 the outer circuit to drive the load and make electric power.
It was found that N-doped TiO2 had much higher photo- The original state of the dye is subsequently restored by
catalytic activity than pure TiO2 in the visible light region, electron donation from the electrolyte, usually an organic
while displaying lower activity in the UV-light region. A solvent containing a redox system, such as the iodide/
nitrogen concentration dependent performance of the pho- triiodide couple. The regeneration of the sensitizer by iodide
tocatalytic activity of the nitrogen-doped TiO2 was found in prevents the recapture of the conduction band electron by
the visible region, and the active sites of N for photocatalysis the oxidized dye. The iodide is regenerated in turn by the
under visible light were identified with the atomic β-N states reduction of triiodide at the counterelectrode, with the circuit
peaking at 396 eV in the XPS spectra.489 In the study of Irie being completed via electron migration through the external
and co-workers, the concentration dependent photocatalytic load. The voltage generated under illumination corresponds
activity of the N-doped TiO2 was attributed to the fact that to the difference between the Fermi level of TiO2 and the
the band structure of the N-doped TiO2 with lower nitrogen redox potential of the electrolyte. Overall, the device
concentration (<2%) was different from that with higher generates electric power from light without suffering any
concentration.483 It was found that the significant increase permanent chemical transformation.9,28-32
in photocatalytic activity in N-doped TiO2 nanoparticles was Cahen et al. explained the cause for the photocurrent and
due to the O-Ti-N bond formation as oxynitride during photovoltage in nanocrystalline mesoporous dye-sensitized
the substitutional doping process.428,477 The photocatalytic solar cells in terms of the separation, recombination, and
oxidation of organic compounds by N-doped TiO2 under transport of electronic charge as well as in terms of electron
visible illumination was mainly via reactions with surface energetics.721 The basic cause for the photovoltage is the
intermediates of water oxidation or oxygen reduction, not change in the electron concentration in the nanocrystalline
by direct reactions with holes trapped at the N-induced electron conductor that results from photoinduced charge
midgap level.486 N-doped TiO2 nanotubes also exhibited high injection from the dye. Pichot and Gregg found that the
photocatalytic oxidation activity for decomposition of gas- photovoltage was determined by photoinduced chemical
eous isopropanol into acetone and carbon dioxide when potential gradients, not by equilibrium electric fields.635 The
illuminated with visible light.528 maximum photovoltage is given by the difference in electron
The photocatalytic activity of sulfur-doped TiO2 has also energies between the redox level and the bottom of the
been studied.492-494,496 The S-doped TiO2 was found to conduction band of the electron conductor, rather than by
display a higher photocatalytic activity in the visible region any difference in electrical potential in the cell, in the dark.
but a lower photocatalytic activity in the UV region.492-494 Charge separation occurs because of the enthalpic and
S-doped TiO2 prepared with different methods showed entropic driving forces that exist at the dye/electron conductor
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2933

under solar illumination with polypyridyl ruthenium and


osmium sensitizers with the general structure ML2(X)2, where
L stands for 2,2′-bipyridyl-4,4′-dicarboxylic acid, M is Ru
or Os, and X presents a halide, cyanide, thiocyanate, acetyl
acetonate, thiacarbamate, or water substituent.88 The dye-
sensitized solar cells with cis-dithiocyanatobis(4,4′-dicar-
boxylic acid-2,2′-bipyridine)ruthenium(II) (N3) displayed
absorption maxima at 518 and 380 nm and emission at 750
nm with a lifetime of 60 ns.625,850 In 2001 the “black dye”
tri(cyanato)-2,2′,2′′-terpyridyl-4,4′,4′′-tricarboxylate) ruthe-
nium(II) was found to achieve 10.4% conversion efficiency
in full sunlight.631 Amphiphilic heteroleptic N3 equivalent
dyes were recently applied to solar cells.673 These amphiphilic
heteroleptic sensitizers had several advantages compared to
the N3 complex: (a) The ground-state pKa of the 4,4′-
dicarboxy-2,2′-bpy was higher to enhance the binding of the
complex onto the TiO2 surface. (b) The decreased charge
on the sensitizer attenuated the electrostatic repulsion and
increased the dye loading. (c) The presence of the hydro-
phobic moiety on the ligand increased the stability of solar
cells toward water-induced desorption. (d) The oxidation
potential of these complexes was cathodically shifted com-
pared to that of the N3 sensitizer, which increased the
reversibility of the ruthenium III/II couple, leading to
enhanced stability. Combining the N3 dye with guanidinium
thiocyanate brought a further increase in the open-circuit
voltage of the solar cell.30,31
Unlike the large amount of effort put forth to optimize
the organic dyes in DSSCs in the past decades, attention has
only recently been paid to the TiO2 nanocrystalline electrode,
and some important results have been obtained. In the
following, various research efforts on the use of the TiO2
Figure 64. (A) Structure and (B) principle of operation and energy nanocrystalline electrode for DSSCs are briefly summarized.
level scheme of the dye-sensitized nanocrystalline solar cell. 5.2.1.1. Mesoporous TiO2 Nanocrystalline Electrodes.
Photoexcitation of the sensitizer (S) is followed by electron injection Zukalova et al. found that ordered mesoporous TiO2 nano-
into the conduction band of an oxide semiconductor film. The dye
molecule is regenerated by the redox system, which itself is
crystalline films showed enhanced solar conversion efficiency
regenerated at the counterelectrode by electrons passed through the by about 50% compared to traditional films of the same
load. Potentials are referred to the normal hydrogen electrode thickness made from randomly oriented anatase nanocrys-
(NHE). The open circuit voltage of the solar cell corresponds to tals.312 The TiO2 nanocrystalline film was prepared via layer-
the difference between the redox potential of the mediator and the by-layer deposition with Pluronic P123 as template. The
Fermi level of the nanocrystalline film indicated with a dashed line. sensitizer used was cis-dithiocyanato(4,4′-dicarboxy-2,2′-
The energy levels drawn for the sensitizer and the redox mediator bipyridine)(4,4′-di-(2-(3,6-dimethoxyphenyl)ethenyl)-2,2′-bi-
match the redox potentials of the doubly deprotonated N3 sensitizer
ground state and the iodide/triiodide couple. Reprinted from Grätzel, pyridine) ruthenium(II), N945. Figure 65 shows the photo-
M. J. Photochem. Photobiol. A: Chem. 2004, 164, 3, Copyright current-voltage characteristics for solar cells based on ordered
2004, with permission from Elsevier. and nonordered TiO2 films. When sensitized by N945, the
0.95-µm-thick nonorganized anatase film gave a conversion
interface, with charge transport aided by such driving forces efficiency of only 2.21%, which increased to 2.74% with
at the electron conductor/contact interface. The mesoporosity surface treatment by TiCl4 prior to dye deposition. Under
and nanocrystallinity of the semiconductor are important not standard global AM 1.5 solar conditions, the cell with an
only because of the large amount of dye that can be adsorbed ordered mesoporous TiO2 nanocrystallinne film gave a
on the very large surface but also for two additional photocurrent density of Ip ) 7 mA/cm2, an open circuit
reasons: (a) they allow the semiconductor small particles potential of UOC ) 0.799 V, and a fill factor of ff ) 0.72,
to become almost totally depleted upon immersion in the yielding 4.04% conversion efficiency. This improvement
electrolyte (allowing for large photovoltages), and (b) the resulted from a remarkable enhancement of the short circuit
proximity of the electrolyte to all particles makes screening photocurrent, due to the huge surface area accessible to both
of injected electrons, and thus their transport, possible.721 the dye and the electrolyte.312
Many ruthenium complexes containing anchoring groups 5.2.1.2. TiO2 Nanotube Electrode. Adachi et al. found
such as carboxylic acid, dihydroxy, and phosphonic acid on that dye-sensitized solar cells with electrodes made of
pyridyl ligands have been used as dyes in the DSSCs. Grätzel disordered single-crystalline TiO2 nanotubes (10-nm diam-
et al. have been leading the research in this field since their eter, 30-300-nm length) displayed an efficiency of 4.88%,
breakthrough in the early 1990s. Tris(2,2′-bipyridyl-4,4′- showing more than double the short-circuit current density
carboxylate) ruthenium(II) was used in DSSCs until the compared to those made of TiO2 nanoparticles of Deguessa
announcement in 1991 of a sensitized electrochemical P-25 in a similar thin-film thickness region.569 Macak et al.
photovoltaic device with a conversion efficiency of 7.1% found that, for Ru-dye (N3) sensitization of self-organized
2934 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 66. Photocurrent-photovoltage characteristics of a TiO2


nanotube array DSSC under 100% AM-1.5 illumination. The inset
Figure 65. Photocurrent-voltage characteristics of a solar cell, shows an SEM image of TiO2 nanotubes. Reprinted with permission
based on TiO2 films sensitized by N945, cis-dithiocyanato(4,4′- from Mor, G. K.; Shankar, K.; Paulose, M.; Varghese, O. K.;
dicarboxy-2,2′-bipyridine)(4,4′-di-(2-(3,6-dimethoxyphenyl)ethenyl)- Grimes, C. A. Nano Lett. 2006, 6, 215. Copyright 2006 American
2,2′-bipyridine) ruthenium(II). (1) Pluronic-templated three-layer Chemical Society.
film, 1.0-µm-thick; (2) nonorganized anatase treated by TiCl4, 0.95-
µm-thick; (3) nonorganized anatase nontreated by TiCl4, 0.95-µm-
thick. The inset shows the SEM image of Pluronic-templated three-
layer TiO2 films. Reprinted with permission from Zukalova, M.;
Zukal, A.; Kavan, L.; Nazeeruddin, M. K.; Liska, P.; Grätzel, M.
Nano Lett. 2005, 5, 1789. Copyright 2005 American Chemical
Society.

TiO2 nanotubes grown by Ti anodization, IPCEmax values


(at 540 nm) of 3.3% and 1.6% (at 530 nm) were obtained
for 2.5-µm- and 500-nm-long nanotubes, respectively.219
Ohsaki et al. found that the higher efficiency of solar cells
with TiO2 nanotube-based electrodes resulted from an
increase in electron density in nanotube electrodes compared
to P25 electrodes.851
Grimes et al. fabricated highly ordered nanotube arrays
(46-nm pore diameter, 17-nm wall thickness, and 360-nm
length) grown perpendicular to an F-doped SnO2-coated glass
substrate by anodic oxidization.201 After crystallization by
oxygen annealing and treatment with TiCl4, the nanotube
arrays were integrated into a DSC structure using a com-
mercially available ruthenium-based dye N719. The cell
generated a photocurrent of 7.87 mA/cm2 with a photocurrent
efficiency of 2.9%, using a 360-nm-thick electrode under
AM 1.5 illumination. They found that the highly ordered
TiO2 nanotube arrays had superior electron lifetimes and
provided excellent pathways for electron percolation in
comparison to nanoparticulate systems. Figure 66 shows the
photocurrent-photovoltage characteristics of the TiO2 nano-
tube DSSC.201 They also found that backside illuminated
solar cells based on 6-µm-long highly ordered nanotube-
array films sensitized by bis(tetrabutylammonium)-cis- Figure 67. (A) Current-voltage (I-V) curves for an inversed opal
cell obtained in the dark and under white light illumination using
(dithiocyanato)-N,N-bis(4-carboxylato-4-carboxylic acid-2,2- an AM 1.5 simulator (Isc ) 1.8 × 10-7 A/cm2, Voc ) 0.78 V, FF
bipyridine)ruthenium(II) (commonly called “N719”) showed ) 0.33). The inset shows an SEM image of an inverse opal TiO2
a power conversion efficiency of 4.24% under AM 1.5 film. (B) Current-voltage (I-V) characteristic of a nanocrystalline
illumination.203 TiO2 cell in the dark and under white light illumination using an
5.2.1.3. Inversed TiO2 Opal. The relatively low efficiency AM 1.5 simulator (Isc ) 8.5 × 10-9 A/cm2, Voc ) 0.87 V, FF )
0.40). Reprinted from Somani, P. R.; Dionigi, C.; Murgia, M.;
obtained in solid-state DSSCs is attributed to the poor Palles, D.; Nozar, P.; Ruani, G. Sol. Energy Mater. Sol. Cells 2005,
penetration of the material into pores of the thick TiO2 films 87, 513, Copyright 2005, with permission from Elsevier.
and the consequent noncontact of the hole transport layer
with the titania electrode. A novel approach to increase the indicated that light conversion efficiency increased by at least
efficiency of solid-state Grätzel solar cells was presented by 1 order of magnitude by the usage of the inversed opal TiO2
Somani et al., using large-surface titania inverse opal films films rather than nanocrystalline TiO2 films (Figure 67). The
as electrodes in fabricating solid-state dye-sensitized organic- better performance of inversed opal cells was due to the wide
inorganic hybrid Grätzel solar cells.352 Direct comparison and well-connected pores in mesoporous TiO2 films that
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2935

Figure 69. Schematic band diagrams of working electrodes


consisting of a TiO2-WO3 buffer layer between TCO and a P25
layer. From: Kang, T. S.; Moon, S. H.; Kim, K. J. J. Electrochem.
Soc. 2002, 149, E155. Copyright 2002. Reproduced by permission
of The Electrochemical Society, Inc.

Figure 68. Photocurrent density versus voltage for the photoelec- both open-circuit photovoltage and short-circuit photocurrent
trochemical cells based on the pure anatase (TiO2 II) and anatase- were enhanced. In the case of the electrode having a buffer
rutile (TiO2 I) nanocrystalline TiO2 electrodes sensitized by N3. layer of less than about 10 mol % WO3, due to the large
The effective area for illumination is 0.5 cm2. The thicknesses of negative VFB, a potential barrier to the conduction band
the sputter deposited layer and the nanocrystalline layer are 20 nm electrons from TiO2 emerged at the TiO2-WO3/TiO2 junc-
and 6 µm, respectively. Conditions: electrolyte, 0.5 M LiI + 0.04 tion. This resulted in a drop in photoinjection efficiency and
M I2 in propylene carbonate (PC); room temperature; light intensity,
98 mW/cm2; AM1.5 spectral radiation. Inset: Performance param- subsequently in the photocurrent. For electrodes having more
eters of solar cells. Reprinted Figure 2 from Han, H.; Zan, L.; than about 75 mol % WO3, the conduction band edge of the
Zhong, J.; Zhao, X. J. Mater. Sci. 2005, 40, 4921, Copyright 2005, buffer layer lay close to or lower than that of TCO, and the
with kind permission of Springer Science and Business Media. relative conduction band energy of the buffer layer was not
particularly beneficial for the electron injection from the
allowed easy penetration of the hole transporting material, conduction band of TiO2.854
allowing good contact with the dye and hence the best 5.2.1.4.3. Core-Shell Structured Nanocrystalline Elec-
efficiency of the cell. trode. Under the operating conditions of a DSSC, the
5.2.1.4. Hybrid TiO2 Nanocrystalline Electrode. 5.2.1.4.1. electrons need to diffuse several micrometers into the TiO2
Anatase-Rutile TiO2 Nanocrystalline Electrode. Han et al. layer surrounded by electron acceptors at a distance of only
found that a hybrid TiO2 electrode composed of a mixture several nanometers. The nanoporous structure of the TiO2
of anatase and rutile phases showed a higher solar-to-electric layer provides a large surface area, allowing absorption of
energy conversion efficiency than one made of pure ana- enough dye molecules to achieve significant optical den-
tase.852,853 Figure 68 shows the performance of the photo- sity.10,855 However, the structure also enhances the recom-
electrochemical cells built with pure anatase (TiO2 II) and bination processes and decreases the total conversion effi-
anatase-rutile (TiO2 I) nanocrystalline TiO2 electrodes ciency of DSSC.856-858 The recombination processes are
sensitized by N3. These electrodes had the same crystalline completely prohibited due to the lack of a significant electric
sizes and surface areas (26 nm, BET 57 m2/g).852 TiO2 I had field that could assist the separation of electrons from holes
71% anatase phase with 29% rutile phase, while TiO2 II had in the TiO2 layer, since small TiO2 nanoparticles allow only
pure anatase phase. The anatase-rutile-based DSSC showed limited band bending at the electrode surface.721,856,857,859
higher performance (efficiency η ) 6.8%, short-circuit Core-shell TiO2 electrodes consisting of a nanoporous TiO2
photocurrents density Jsc ) 19.4 mA/cm2, open-circuit covered with a shell of another metal oxide have been shown
photovoltage VOC ) 652 mV, fill factor ff ) 0.53) than the to slow the recombination processes by the formation of an
pure TiO2 (η ) 5.3%, Jsc ) 18.4 mA/cm2, VOC ) 582 mV, energy barrier at the TiO2 surface.560,648,649,860-865 The con-
ff ) 0.51). duction band potential of the shell should be more negative
5.2.1.4.2. Nanocrystalline Electrode with a Buffer Layer. than that of TiO2 in order to generate an energy barrier for
In a standard nanoporous electrode during DSSC operation, the reaction of the electrons present in TiO2 with the oxidized
two main problems are associated with the porous geom- dye or the redox mediator in solution. Two approaches are
etry: (a) the high-area cross section for recombination of employed to fabricate the nanoporous core-shell electrodes.
photoinjected electrons with holes that are transferred to the The first approach involves synthesis of core-shell nano-
electrochemical mediator and (b) the image field opposing particles that are applied onto the conducting sub-
the separation process that is distributed inside the TiO2 strate.560,648,649,863,866 An energy barrier forms not only at the
nanoporous electrode. The conversion efficiency of a DSSC electrode/electrolyte interface but also between the individual
decreases due to recombination losses of photoinjected TiO2 nanoparticles. The second approach involves a nano-
electrons with oxidized dye molecules or a redox couple at porous TiO2 electrode coated with the thin shell lay-
the surface of nanocrystalline TiO2. Various methods have er.670,860-862,864,865,867 The TiO2 nanoparticles are connected
been adopted to prevent this loss. Kang et al. added a buffer directly to each other allowing electron transport through
layer of a TiO2-WO3 composite material between a TCO TiO2.
substrate (Figure 69) and a TiO2 layer and found that the The approach involving nanoporous electrodes in a well-
buffer layer effectively isolated dye molecules and electro- defined core-shell configuration is usually a TiO2 core
lytes from directly contacting the conducting substrate.854 In coated with Al2O3,596,865,868-870 MgO,871 SiO2,865 ZrO2,865 or
the presence of the buffer layer having 15-75 mol % WO3, Nb2O5.860,670 For example, Zaban et al. found that TiO2/Nb2O5
2936 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 71. Schematic view of the collector-shell electrode. This


core shell electrode consists of a conductive nanoporous matrix
that is coated with TiO2. Reprinted with permission from Chappel,
S.; Grinis, L.; Ofir, A.; Zaban, A. J. Phys. Chem. B 2005, 109,
1643. Copyright 2005 American Chemical Society.

which were induced by the properties of the two materials


at the core/shell interface.862
Palomares et al. found that the conformal growth of an
overlayer of Al2O3 on a nanocrystalline TiO2 film resulted
in a 4-fold retardation of interfacial charge recombination
Figure 70. (A) Schematic view of the new bilayer nanoporous
and a 30% improvement in photovoltaic device efficiency.870
electrode which consists of a nanoporous TiO2 matrix covered with Fabregat-Santiago et al. found that the alumina barrier
a thin layer of Nb2O5. The Nb2O5 coating forms an inherent energy reduced the recombination of photoinjected electrons to both
barrier at the electrode/electrolyte interface, which reduces the the dye cations and the oxidized redox couple, due to two
recombination rate of the photoinjected electrons. From: Zaban, effects: (a) almost complete passivation of surface trap states
A.; Chen, S. G.; Chappel, S.; Gregg, B. A. Chem. Commun. 2000, in TiO2 that were able to inject electrons to acceptor species
2231 (http://dx.doi.org/10.1039/b005921h) s Reproduced by per- and (b) slowing down by a factor of 3-4 of the rate of
mission of The Royal Society of Chemistry. (B) I-V curves of
four DSSCs differing by the nanoporous electrodes used to fabricate interfacial charge transfer from conduction band states.868
them: the TiO2 reference electrode (a), and three bilayer electrodes O’Regan found that the Al2O3 layer acted as a tunnel barrier,
(b-d). The Nb2O5 coating was made by a 30 s dipping of a 6 µm thus increasing Voc and the fill factor.869 Palomares et al.
TiO2 matrix in a 5 mM solution of NbCl5 in dry ethanol (b), Nb- prepared SiO2, Al2O3, and ZrO2 overlayers by dipping
(isopropoxide)5 in 2-propanol (c), and Nb(ethoxide)5 in ethanol (d). mesoporous nanocrystalline TiO2 films in organic solutions
Reprinted with permission from Chen, S. G.; Chappel, S.; Diamant, of their respective alkoxides, followed by sintering at 435
Y.; Zaban, A. Chem. Mater. 2001, 13, 4629. Copyright 2001
American Chemical Society. °C.865 The metal oxide overlayers acted as barrier layers for
interfacial electron-transfer processes. The most basic over-
nanoporous electrodes could improve the performance of layer coating, Al2O3 (pzc ) 9.2), was optimal for retarding
interfacial recombination losses under negative applied bias,
dye-sensitized solar cells by >35%.670,860 Figure 70 shows
with an increase in open-circuit voltage of up to 50 mV and
a bilayer nanoporous electrode which consists of a nanopo-
a 35% improvement in overall device efficiency. Diamant
rous TiO2 matrix covered with a thin layer of Nb2O5 and
et al. found that SrTiO3-coated nanoporous TiO2 electrodes
the performance of three TiO2 electrodes coated with Nb2O5.
increased the open circuit photovoltage while reducing the
For the best coating condition, the photocurrent increased short circuit photocurrent and resulting in a 15% improve-
from 10.2 to 11.4 mA/cm2, the photovoltage from 661 to ment of the overall conversion efficiency of the solar cell.861
730 mV, and the fill factor from 51.0 to 56.5%. As a result, The SrTiO3 layer shifted the conduction band of the TiO2 in
the conversion efficiency of the solar cell increased by 35% the negative direction due to a surface dipole rather than
from 3.62 to 4.97%.860 They also found that sometimes the forming an energy barrier at the TiO2/electrolyte inter-
shell material shifted the conduction band potential of the face.861,862 The shell having a more negative conduction band
core rather than forming an energy barrier. For example, potential acted as an energy barrier that slowed recombination
coating of TiO2 with a SrTiO3 shell resulted in a shift of the reactions. Photoexcitation of dye molecules anchored to
TiO2 conduction band in the negative direction.861,862 Con- ultrathin (e1 nm) outer shells of insulators or semiconductors
sequently, introduction of a SrTiO3-coated TiO2 electrode on n-type semiconductor crystallites resulted in electron
to a DSSC increased the open circuit photovoltage while transfer to the inner core material.
reducing the short circuit photocurrent compared to that of However, there is still considerable recombination that
the noncoated TiO2 electrode.861,862 Diamant et al. found that increases with the distance between the electron injection
the mechanism by which the shell affected the electrode point and the current collector. In other words, the limited
properties depended on the coating material. Coating materi- lifetime of the injected electron and the slow diffusion rate
als included Nb2O5, ZnO, SrTiO3, ZrO2, Al2O3, and SnO2.862 inside the porous structure limit the effective thickness of
The coating Nb2O5 formed a surface energy barrier, which the nanoporous electrode. Chappel et al. proposed a electrode
slowed the recombination reactions, while the other shell design, shown in Figure 71, with a core shell configuration
materials each formed a surface dipole layer that shifted the based on a conductive ITO or Sb-doped SnO2 matrix coated
conduction band potential of the core TiO2. The shift with TiO2.872 In principle, the conducting core extended the
direction and magnitude depended on the dipole parameters current collector into the nanoporous network and was
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2937

Figure 72. (a) SEM of a cross section of the bilayer photonic crystal-nano-TiO2 photoelectrode. The conductive glass is at the top of the
image in part a. The photonic crystal layer and the nanocrystalline TiO2 layer are enlarged in parts b and c, respectively. Reprinted with
permission from Nishimura, S.; Abrams, N.; Lewis, B. A.; Halaoui, L. I.; Mallouk, T. E.; Benkstein, K. D.; van de Lagemaat, J.; Frank,
A. J. J. Am. Chem. Soc. 2003, 125, 6306. Copyright 2003 American Chemical Society.
denoted the nanoporous “collector shell electrode”. Conse- Nishimura347 and Halaoui333 reported an enhancement in
quently, the distance between the injection spot and the the light conversion efficiency of dye-sensitized TiO2 solar
current collector should decrease to several nanometers cells by coupling a conventional nanocrystalline TiO2 film
throughout the nanoporous electrode, in contrast to several to a TiO2 inverse opal, with a 26% increase in the IPCE
micrometers with the standard electrode. All electrons relative to that of a nanocrystalline film of the same overall
injected into the electrode, including those generated several thickness in the 550-800 nm spectral range. They found
micrometers away from the substrate, had to travel a very that the bilayer architecture, rather than enhanced light
short distance before reaching the current collector. As shown harvesting within the inverse opal structures, was responsible
by several studies, transport shorter than 1 µm provides 100% for the bulk of the gain in the IPCE.333 Figure 72 shows an
collection efficiency. In addition, the new collector-shell SEM image of a cross section of the bilayer photonic
electrode contained inherent screening capability due to the crystal-nano-TiO2 photoelectrode.347
high doping level of the conducting matrix. Theoretically, Figure 73 shows the sketch for the mechanism of the
the new design should enable efficient charge separation and photonic crystal in enhancing absorption in certain regimes.347
collection for thick nanoporous layers and solid electro- The fact that light waves were localized in different parts of
chemical mediators. They found that, unless the TiO2 coating the structure, depending on their energy, implied that an
was thicker than 6 nm, the electrode performance was very absorber in the high dielectric medium should interact more
low due to fast recombination.872 strongly with light at wavelengths to the red of the stop band,
5.2.1.4.4. Electrode Coupled with Photonic Crystals. and less strongly to the blue. Effectively, the red part of the
Development of photosensitizers with improved spectral spectrum of this absorber would “borrow” intensity from the
response at the low-energy end of the solar spectrum has blue part.
not proven so successful because dye molecules with high Figure 74A shows the effect of the TiO2 photonic crystal
red absorbance have lower excited-state excess free energy, as compared to a film of nanocrystalline TiO2 on the
thus lowering the quantum yield for charge injection. absorption spectra when dye is adsorbed to the surface.347
Increasing the thickness of the film beyond 10-12 µm in In a comparison of the spectrum of dye molecules adsorbed
order to increase the absorbance in the red results in an to the TiO2 photonic crystal film with that of a conventional
increase in the electron transport length and the recombina- nanocrystalline TiO2 film, there was a substantial enhance-
tion rate, and a decrease in the photocurrent. An alternative ment absorbance on the red side of the stop band, as well as
approach to improving efficiency was to increase the path a slight attenuation of absorbance on the blue side of the
length of light by enhancing light scattering in the TiO2 stop band. The enhanced absorbance was most pronounced
films.873-878 While the small size of TiO2 nanoparticles (10- between 500 and 550 nm, but it persisted to a lesser degree
30 nm) employed to ensure a high surface area makes at longer wavelengths. Figure 74B shows the enhancement
conventional nanocrystalline TiO2 films poor light scatterers, of the performance of a bilayer electrode compared to a
mixing the nanoparticles with larger particles or applying a conventional nanocrystalline TiO2 photoelectrode.347 Between
scattering layer to the nanocrystalline film has been shown 400 and 530 nm, there was little difference between the two
to increase light harvesting by enhancing the scattering of kinds of electrodes. The close similarity in the maximum
light.873-878 photocurrent from the two electrodes was consistent with
2938 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 73. (A) Simplified optical band structure of a photonic


crystal. Near the Brillouin zone center, light travels with velocity Figure 74. (A) Absorption spectra of the TiO2 photonic crystal
c0/n, where c0 is the speed of light in a vacuum and n is the average (a), the N719 dye adsorbed on the photonic crystal (b), and the
refractive index. At photon energies approaching a full band gap dye adsorbed on a film of nanocrystalline TiO2 (c). The position
or a stop band from the red side, the group velocity of light of the stop band at 486 nm is indicated by the arrow. (B)
decreases and light can be increasingly described as a sinusoidal Wavelength dependence of the short-circuit photocurrent in the
standing wave that has its highest amplitude in the high-refractive- bilayer electrode (a) and the conventional nanocrystalline TiO2
index part of the structure. At energies above the band gap or stop photoelectrode (b). The position of the stop band maximum in the
band, the standing wave is predominantly localized in the low index bilayer electrode was 610 nm. Reprinted with permission from
part of the photonic crystal, i.e., in the air voids. (B) Illustration of Nishimura, S.; Abrams, N.; Lewis, B. A.; Halaoui, L. I.; Mallouk,
the effect of standing wave localization on dye absorbance. In an T. E.; Benkstein, K. D.; van de Lagemaat, J.; Frank, A. J. J. Am.
isotropic medium, the dye absorbs strongly in the blue but weakly Chem. Soc. 2003, 125, 6306. Copyright 2003 American Chemical
in the red (heavy line). If the stop band is tuned to the position Society.
shown by the arrow, the blue absorbance is diminished and the red
absorbance is increased when the dye is confined to the high- steps. First, light absorption occurred in the surface-absorbed
refractive-index part of the photonic crystal (dotted line). Reprinted photoreceptors, giving rise to energetic electrons. Second,
with permission from Nishimura, S.; Abrams, N.; Lewis, B. A.; electrons from the photoreceptor excited state were injected
Halaoui, L. I.; Mallouk, T. E.; Benkstein, K. D.; van de Lagemaat, into the conduction levels of the adjacent conductor, where
J.; Frank, A. J. J. Am. Chem. Soc. 2003, 125, 6306. Copyright 2003 they travelled ballistically through the metal at an energy,
American Chemical Society.
1e, above the Fermi energy, Ef. Third, provided that 1e was
the fact that both contain the same amount of dye. Between greater than the Schottky barrier height, f, and the carrier
540 and 750 nm, the short circuit photocurrent was substan- mean-free path was long compared to the metal thickness,
tially increased in the bilayer electrode. The overall gain, the electrons traversed the metal and entered the conduction
integrated over the visible spectrum (400-750 nm), was levels of the semiconductor (internal electron emission). The
about 30%. Localization of heavy photons at the edges of absorbed photon energy was preserved in the remaining
the photonic stop band347,879,880 from Bragg diffraction in the excess electron free energy when it was collected at the back
periodic lattice and multiple scattering events at disordered ohmic contact, giving rise to the photovoltage, V. The
regions in the photonic crystal or at disordered films led photooxidized dye was reduced by transfer of thermalized
ultimately to enhanced backscattering.333 This largely ac- electrons from states near Ef in the adjacent metal. Devices
counted for the enhanced light conversion efficiency in the fabricated by using a fluorescein photoreceptor on an Au/
red spectral range (600-750 nm), where the sensitizer was TiO2/Ti multilayer structure had typical open-circuit photo-
a poor absorber.333 voltages of 600-800 mV and short-circuit photocurrents of
10-18 mA cm-2 under 100 mW cm-2 visible light illumina-
5.2.2. Metal/Semiconductor Junction Schottky Diode Solar tion: the internal quantum efficiency (electrons measured
Cell per photon absorbed) was 10%. This alternative approach
McFarland and Tang reported a multilayer photovoltaic to photovoltaic energy conversion might provide the basis
device structure in which photon absorption occurred in for durable low-cost solar cells using a variety of materials.
photoreceptors deposited on the surface of an ultrathin metal/
semiconductor junction Schottky diode.881 The device struc- 5.2.3. Doped TiO2 Nanomaterials-Based Solar Cell
ture was a solid-state multilayer with a photoreceptor layer Lindgren et al. found that N-doped TiO2 nanocrystalline
deposited on a 10-50 nm Au film, which capped 200 nm porous thin films showed visible light absorption in the
of TiO2 on an ohmic metal back contact (Figure 75). The wavelength range from 400 to 535 nm and generated an
photon-to-electron conversion in this device occurred in four incident photon-to-current efficiency response in good agree-
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2939

and oxidized by the holes to form O2, leading to overall water


splitting.883-885 The width of the band gap and the potentials
of the conduction and valence bands are important. The
bottom level of the conduction band has to be more negative
than the reduction potential of H+/H2 (0 V vs NHE), while
the top level of the valence band has to be more positive
than the oxidation potential of O2/H2O (1.23 V). The potential
of the band structure of TiO2 is just the thermodynamical
requirement. Other factors such as charge separation, mobil-
ity, and lifetime of photogenerated electrons and holes also
affect the photocatalytic properties of TiO2. These factors
are strongly affected by the bulk properties of the material
such as crystallinity. Surface properties such as surface states,
surface chemical groups, surface area, and active reaction
sites are also important.768 The water-splitting process in
return affects the local pH environment and surface structures
of the TiO2 electrode.769
Figure 75. Electron transfer in the operating photovoltaic device: Salvador conducted a thermodynamic and kinetic consid-
(process A) photon absorption and electron excitation from the eration of water-splitting and competitive reactions in the
chromophore ground state, S, to the excited state, S*; (process B) photoelectrochemical cell, and they found that the overvolt-
energetic electron transfer from S* into and (ballistically) through age for evolution of O must be minimized, which was on
the conducting surface layer and over the potential energy barrier
into the semiconductor; (process C) conduction of electrons as the order of 0.6 eV for n-TiO2 electrodes loaded with
majority carriers within the semiconductor to the ohmic back-contact RuO2.767 Cocatalysts such as Pt and NiO are often loaded
and through the load; (process D) reduction of the oxidized on the surface in order to introduce active sites for H2
chromophore, S, by a thermal electron from the conductor surface. evolution. Thus, suitable bulk and surface properties and
Shown schematically are the relative energies of the electron levels energy structure are demanded for photocatalysts.
within the device structures, the Schottky barrier, f, the Fermi
energy, Ef, and the semiconductor band gap, Eg. Reprinted with
Laser-induced photocatalytic oxidation/splitting of water
permission from McFarland, E. W.; Tang, J. Nature 2003, 421, over TiO2 catalysts was studied.883,886,887 Sayama and Ara-
616. Copyright Nature Publishing Group. kawa found that addition of carbonate salts to Pt-loaded TiO2
suspensions led to highly efficient water splitting.888 The
carbonate ions affected both the Pt particles and the TiO2
surface. The Pt was covered with some titanium hydroxide
compounds and the rate of the back reaction on the Pt was
suppressed effectively in the presence of carbonate ions. The
carbonate species aided desorption of O2 from the TiO2
surface.888 Khan and Akikusa found that bare n-TiO2
nanocrystalline film electrodes were unstable during water-
splitting reactions under illumination of light and their
Figure 76. Reaction schemes for semiconductor photocatalysts. stability could be significant improved when covered with
Reprinted Figure 2 from Kudo, A. Catal. SurV. Asia 2003, 7, 31, Mn2O3.759
Copyright 2003, with kind permission of Springer Science and
Business Media. 5.3.2. Use of Reversible Redox Mediators
It has been reported that pure TiO2 could not easily split
ment with the optical spectra.385 For the best nitrogen-doped
water into H2 and O2 in the simple aqueous suspension
TiO2 electrodes, the photoinduced current due to visible light
system.413,754,889 The main problem is the fast, undesired
and at moderate bias increased around 200 times compared
electron-hole recombination reaction.762 Therefore, it is
to the behavior of pure TiO2 electrodes.
important to prevent the electron-hole recombination pro-
cess. The Pt-TiO2 system could be illustrated as a “short-
5.3. Photocatalytic Water Splitting circuited” photoelectrochemical cell, where a TiO2 semi-
5.3.1. Fundamentals of Photocatalytic Water Splitting conductor electrode and a platinum-metal counterelectrode
are brought into contact. Well-dispersed metal particles act
An enormous research effort has been dedicated to the as miniphotocathodes, trapping electrons, which reduces
study of the properties and applications of TiO2 under light water to hydrogen.
illumination since the discovery of photocatalytic splitting The role of sacrificial reagents is shown in Figure 77.761
of water on a TiO2 electrode in 1972 (Fujishima and When the photocatalytic reaction is carried out in aqueous
Honda).6-8 Photocatalytic splitting of water into H2 and O2 solutions including easily oxidizable reducing reagents,
using TiO2 nanomaterials continues to be a dream for clean photogenerated holes irreversibly oxidize the reducing
and sustainable energy sources.882 reagents instead of water. This makes the photocatalyst
Figure 76 shows the principle of water splitting using a electron-rich, and a H2 evolution reaction is enhanced as
TiO2 photocatalyst.761 When TiO2 absorbs light with energy shown in Figure 77a. On the other hand, in the presence of
larger than the band gap, electrons and holes are generated electron acceptors such as Ag+ and Fe3+, the photogenerated
in the conduction and valence bands, respectively. The electrons in the conduction band are consumed by them and
photogenerated electrons and holes cause redox reactions. an O2 evolution reaction is enhanced as shown in Figure
Water molecules are reduced by the electrons to form H2 77b. These reactions using sacrificial reagents are regarded
2940 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

I- oxidation to IO3- over Pt-TiO2-anatase, and (b) IO3-


reduction to I- and water oxidation to O2 over TiO2-rutile.
IO3- reduction to I- over Pt-TiO2-anatase is an undesirable
reaction. If this reaction is suppressed, the total water-splitting
reaction will take place more efficiently. The advantage of
this system is that H2 gas is evolved over the Pt-TiO2-
anatase photocatalyst only and that O2 gas is evolved over
the TiO2-rutile photocatalyst only, even from a mixture of
IO3- and I- in a basic aqueous solution. Therefore, another
undesirable backward reaction, H2O formation from H2 and
O2 on Pt particles, was suppressed.753 They found that
addition of a small amount of iodide anion, I-, into the
aqueous suspension of Pt-TiO2-anatase photocatalyst sig-
nificantly improved the splitting into H2 and O2 with a
stoichiometric ratio. The iodide anion was adsorbed prefer-
entially onto the Pt cocatalyst as iodine atom. This iodine
layer effectively suppressed the backward reaction of water
formation from H2 and O2 to H2O over the Pt surface.754
Fujihara et al. studied the photochemical splitting of water
Figure 77. Photocatalytic H2 (a) or O2 (b) evolution in the presence by combining the reduction of water to hydrogen using
of sacrificial reagents. Reprinted Figure 5 from Kudo, A. Catal. bromide ions and the oxidation of water to oxygen using
SurV. Asia 2003, 7, 31, Copyright 2003, with kind permission of FeIII ions.892 The bromide ions were oxidized to bromine on
Springer Science and Business Media. Pt-loaded TiO2 nanoparticles, and the FeIII ions were reduced
to FeII ions on TiO2 nanoparticles. These two reactions were
carried out in separated compartments and combined via
platinum electrodes and cation-exchange membranes as
shown in Figure 79. At the electrodes, FeII ions were oxidized
by bromine, and protons were transported through the
membranes to maintain the electrical neutrality and pH of
the solutions in the two compartments. As a result, water
was continuously split into hydrogen and oxygen under
radiation. The reversible reactions on photocatalysts which
often suffered from the effects of back reactions were largely
prevented due to the low concentration of the products in
Figure 78. Proposed reaction mechanism for overall photocatalytic solution.
water splitting using a IO3-/I- redox mediator and a mixture of Lee et al. found that a considerable amount of photocata-
Pt-TiO2-antase and TiO2-rutile photocatalysts. Reprinted with lytic H2 was produced from water over NiO/TiO2 in
permission from Abe, R.; Sayama, K.; Domen, K.; Arakawa, H. proportion to the hole scavenger CN-.890 Galinska and
Chem. Phys. Lett. 2001, 344, 339. Copyright 2001 Elsevier.
Walendziewski studied water splitting over a Pt-TiO2
catalyst with various sacrificial reagents, such as methanol,
as half reactions and are often employed for test reactions
Na2S, EDTA, and I- and IO3- ions, and they found that the
of photocatalytic H2 or O2 evolution. However, one should
sacrificial reagents had a key role in hydrogen production
realize that the results do not guarantee a photocatalyst to
via the photocatalyzed water-splitting reaction.757 Photocata-
be active for overall water splitting into H2 and O2 in the
lytic water splitting was obtained when EDTA and Na2S were
absence of sacrificial reagents.
used. They acted as effective hole scavengers, preventing
A sacrificial reagent helps to control the electron-hole
oxygen formation and the recombination reaction of oxygen
recombination process. The photoefficiency of the process
with hydrogen.
can be improved by the addition of sacrificial reagents.754,889,890
The sacrificial reagents help separation of the photoexcited 5.3.3. Use of TiO2 Nanotubes
electrons and holes. Various compounds such as methanol,
ethanol, EDTA (an ethylenediaminetetraacetic derivative), Mor et al. found that highly ordered TiO2 nanotube arrays
Na2S, and Na2SO4 or ions such as I-, IO3-, CN-, and Fe3+ efficiently decomposed water under UV radiation.198 The
have been used as sacrificial reagents.753-755,757,890,891 authors found that the nanotube wall thickness was a key
Abe et al. conducted a series of experiments on water parameter influencing the magnitude of the photoanodic
splitting under sunlight.753-755 They designed a new photo- response and the overall efficiency of the water-splitting
catalytic reaction that split water into H2 and O2 by a two- reaction. For TiO2 nanotubes with 22-nm pore diameter and
step photoexcitation system composed of an IO3-/I- shuttle 34-nm wall thickness (Figure 80A), upon 320-400 nm
redox mediator and two different TiO2 photocatalysts: Pt- illumination at an intensity of 100 mW/cm2, hydrogen gas
loaded TiO2-anatase for H2 evolution and TiO2-rutile for O2 was generated at the power-time normalized rate of 960
evolution (Figure 78).753 Simultaneous gas evolution of H2 mmol/h W (24 mL/h W) at an overall conversion efficiency
(180 mmol/h) and O2 (90 mmol/h) was observed from a basic of 6.8% as shown in Figure 80B.198,199 They also claimed
(pH ) 11) NaI aqueous suspension of two different TiO2 that, for illumination at 320-400 nm (98 mW/cm2), the TiO2
photocatalysts under UV radiation. The overall water splitting nanotube-array photoanodes could generate H2 by H2O
proceeded by the redox cycle between IO3- and I- under photoelectrolysis with a photoconversion efficiency of
basic conditions as follows: (a) water reduction to H2 and 12.25%.212 Park et al. further found that, when doped with
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2941

Figure 80. (A) SEM images, top view, of 20 V TiO2 nanotube


arrays anodized at 5 °C. (B) Photoconversion efficiency as a
function of measured potential [vs Ag/AgCl] for 10 V samples
anodized at four temperatures [i.e., 5, 25, 35, and 50 °C]. Reprinted
with permission from Mor, G. K.; Shankar, K.; Paulose, M.;
Varghese, O. K.; Grimes, C. A. Nano Lett. 2005, 5, 191. Copyright
2005 American Chemical Society.
Figure 79. (A) Schematic of the photocatalytic reaction cell for
splitting water. (B) Energy diagram of splitting of water by
combined photocatalytic reactions. From: Fujihara, K.; Ohno, T.;
Matsumura, M. Faraday Trans. 1998, 94, 3705 (http://dx.doi.org/
10.1039/a806398b) s Reproduced by permission of The Royal
Society of Chemistry.

carbon, TiO2-xCx nanotube arrays showed more efficient


water splitting under UV and visible light illumination (>420
nm) than pure TiO2 nanotube arrays.475
5.3.4. Water Splitting under Visible Light Figure 81. Strategy of the development of photocatalysts with a
5.3.4.1. Water Splitting over Doped TiO2 Nanomate- visible light response. Reprinted Figure 6 from Kudo, A. Catal.
rials. In general, the conduction bands of stable oxide SurV. Asia 2003, 7, 31, Copyright 2003, with kind permission of
semiconductor photocatalysts consisting of metal cations with Springer Science and Business Media.
a d0 and d10 configuration consist of empty orbitals (LUMO) narrow. New photocatalysts having the band structure shown
of the metal cations. On the other hand, the valence bands in Figure 81 are necessary in order to develop materials for
consist of O2p orbitals. The potential of this valence band splitting water into H2 and O2 under visible light.761 The
(about +3 eV) is considerably more positive than the created levels have to possess not only the thermodynamical
oxidation potential of H2O to O2 (E0 ) 1.23 V). Therefore, potential for oxidation of H2O but also the catalytic properties
the band gaps of oxide semiconductor photocatalysts with for the four-electron oxidation reaction. The following
the potential for H2 evolution inevitably become wide. strategies can be considered for the development of visible
Accordingly, a valence band or an electron donor level light-driven photocatalysts: (i) forming a donor level above
consisting of orbitals of some element, except for O2p, has a valence band by doping some element into conventional
to be formed to make the band gaps or the energy gaps photocatalysts with wide band gaps such as TiO2; (ii) creating
2942 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

a new valence band employing some element; and (iii)


controling the band structure by making a solid solution.761
Borgarello et al. found that water cleavage could be
induced with visible light in colloidal solutions of Cr-doped
TiO2 nanoparticles deposited with ultrafine Pt or RuO2.431
A pronounced synergistic effect in catalytic activity was
noted when both RuO2 and Pt were co-deposited onto the
particle. Jin and Lu found that Pt/B-doped TiO2 was a good
system for water splitting under a B4O72- environment
without sacrificial electron donor reagents.893 Luo et al. found
that Br-- and Cl--co-doped nanocrystalline TiO2 with the
absorption edge shifted to a lower energy region displayed
higher efficiency for water splitting than pure TiO2.508 Jing
et al. found that a Ni-doped mesoporous TiO2 photocatalyst Figure 82. Energy diagram of H2 production from water over dye-
with 0.2 wt % Pt accomplished hydrogen evolution at nearly sensitized Pt/TiO2 photocatalysts in the presence of I- or EDTA
125.6 lmol/h compared to 81.2 lmol/h for TiO2 P25.894 N-, as an electron donor. Reprinted with permission from Abe, R.;
B-doped TiO2 nanomaterials have displayed higher activity Sayama, K.; Sugihara, H. J. Sol. Energy Eng. 2005, 127, 413.
than pure TiO2 in water splitting, i.e., under visible light Copyright 2005 by ASME.
illumination.529,889 Khan et al. found that a C-doped TiO2
nanocrystalline film with visible light response obtained by combination of single-crystal p-SiC and nanocrystalline
controlled combustion of Ti metal in a natural gas flame n-TiO2 photoelectrodes.899 Both photoelectrodes (p-SiC and
had a high water-splitting performance with a total conver- n-TiO2) were placed side by side facing the light source and
sion efficiency of 11% and a maximum photoconversion in contact with an electrolyte of 0.5 M H2SO4. The open
efficiency of 8.35% when illuminated at 40 mW/cm2,476 circuit potential was found to be 1.24 V between the n-TiO2
although there were questions about its solar-to-hydrogen and p-SiC photoelectrodes, with a maximum photocurrent
conversion efficiency by other researchers.895-897 density of 0.05 mA cm-2 under a closed circuit potential of
Matsuoka et al. developed visible light responsive TiO2 0.23 V, corresponding to an efficiency of 0.06%. The low
nanocrystalline thin films by the radio frequency magnetron cell photocurrent density and the photoconversion efficiency
sputtering method, which decomposed water when Pt-loaded for the p-SiC/n-TiO2 self-driven system for the water-splitting
and in the presence of a sacrificial reagent such as methanol reaction were due to the high band gap energies of both
or silver nitrate under visible light.763,764 semiconductors and high recombination of photogenerated
5.3.4.2. Water Splitting over Dye-Sensitized TiO2. carriers mainly in the covalently bonded p-SiC.
Duonghong et al. found that TiO2 loaded simultaneously with Takabayashi et al. proposed a solar water-splitting system
ultrafine Pt and RuO2 displayed extremely high activity as based on a composite polycrystalline-Si/doped TiO2 thin-
an H2O decomposition catalyst under band gap excitation film electrode for high-efficiency and low-cost by combining
of the TiO2 and that, when Ru(bipy)32+ or rhodamine B was the advantages of Si and doped TiO2: (1) an n-Si electrode
used as a sensitizer, H2O was decomposed under visible with surface alkylation and a metal nanodot coating gave
light.898 an efficient and stable photovoltaic characteristic, and (2)
Abe et al. investigated H2 production over merocyanine TiO2 doped with other elements, such as nitrogen and sulfur,
or coumarin dye C343 or Ru complex dye N3 dye-sensitized could cause water photooxidation (oxygen photoevolution)
Pt/TiO2 photocatalysts under visible light in a water- under visible light illumination.770 The structure and working
acetonitrile solution containing iodide as an electron donor.756 mechanism of solar water splitting with this system is shown
They found that the rates of H2 evolution decreased with in Figure 83. Although a high solar-to-chemical conversion
increasing proportion of water in the solutions because of a efficiency of more than 10% was calculated for this system,
decrease in the energy gap between the redox potential of several major problems needed to be solved before the real
I3-/I- and the HOMO levels of the dyes, which decreases device could show promising performance.770
the efficiency of electron transfer from I- to dye. The energy
diagram and the mechanism for the H2 production from water 5.4. Electrochromic Devices
over the dye-sensitized Pt/TiO2 photocatalyst system are
TiO2 nanomaterials have been widely explored as elec-
shown in Figure 82. The two key electron-transfer steps,
trochromic devices, such as electrochromic windows and
electron injection from an excited state of the dye to the TiO2
displays.611,634,772,900-916 Electrochromism can be defined as
conduction band and oxidation of I- to I3- (steps 2 and 5),
the ability of a material to undergo color change upon
occurred efficiently in acetonitrile solvent. The increased ratio
oxidation or reduction. Electrochromic devices are able to
of water hindered electron transfer from I- to the HOMO
vary their throughput of visible light and solar radiation upon
level of the oxidized dye (step 5).
electrical charging and discharging using a low voltage. A
In addition, Park and Bard designed two different kinds small voltage applied to the windows will cause them to
of cells with bipolar dye-sensitized TiO2/Pt panels connected darken; reversing the voltage causes them to lighten. Thus,
so that their photovoltages added to provide vectorial electron one can regulate the amount of energy entering through a
transfer for unassisted water splitting to yield the separated “smart window” so that the need for air conditioning in a
products H2 and O2.765 cooled building decreases. The energy efficiency inherent
in this technology can be large, provided that the control
5.3.5. Coupled/Composite Water-Splitting System
strategy is adequate. Additionally, the transmittance regula-
Akikusa et al. found that a self-driven system for water tion can impart glare control as well as user control of the
splitting under illumination could be achieved with the indoor environment. The absorbance, rather than the reflec-
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2943

The second type is the electrochromism of nanocrys-


talline TiO2 electrodes modified with viologens and/or
anthrachinons equipped with a surface anchoring
group.902-904,906-908,910,917-919 This category usually has fast
switching times and considerable optical dynamics, due to
the combination of good conductivity between the TiO2
nanoparticles and the fast electron exchange between TiO2
and the monolayer of the electrochromic compound covering
each particle.904 Bach et al. demonstrated high-quality paper-
like electrochromic displays based on nanostructured TiO2
films modified with electrochromophores with excellent ink-
on-paper optical qualities, fast response times, and low power
consumption.901 Moeller et al. demonstrated electrochromic
pictures with unprecedented resolution (360 dpi) in transpar-
ent and reflective electrochromic displays (ECD) based on
ink-jet printing technology and cascade-type crosslinking
reactions of viologens in the mesopores of a TiO2 electrode,
with a completely transparent counterelectrode based on
mesoporous antimony tin oxide coated with CeO2.913

5.4.1. Fundamentals of Electrochromic Devices


Figure 84A shows the principle of the electrochromism
of a molecular monolayer adsorbed on TiO2.902 A molecule,
which functions as the electrochromophore and exhibits
different colors in different oxidation states, must be chosen
such that its redox potential lies above the conduction band
edge of the TiO2 nanocrystalline electrode at the liquid/solid
interface. In this way, electrons can be transferred reversibly
from the conduction band to the molecule. The TiO2
electrode in fact behaves like a conductor for the adsorbed
electroactive species. If the redox potential is situated below
the conduction band edge, the reduction process is irrevers-
ible. Figure 84B shows the TiO2 nanocrystalline electro-
chromic devices based on viologen (solvent: glutarodinitrile)
with a counterelectrode made of Prussian blue.902 The device
could be switched back and forth between the colorless and
the colored states within 1 s.
Figure 83. (A) Schematic illustration and (B) the operation The nanocrystalline structure of the TiO2 film makes
principle of solar water splitting with a composite polycrystalline- possible 100- to 1000-fold amplification compared to a flat
Si/doped TiO2 semiconductor electrode. Reprinted from Takaba-
yashi, S.; Nakamura, R.; Nakato, Y. J. Photochem. Photobiol., A: surface as shown in Figure 85.902 The combination of high
Chem. 2004, 166, 107, Copyright 2004, with permission from conductivity of the nanocrystalline TiO2 particles, fast
Elsevier. electron exchange with the molecular monolayer, optical
tance, is modulated so that the electrochromic devices tend amplification by the porous structure, and fast charge
to heat up in their low-transparent state.752 compensation by ions in the contacting liquid makes the
Two types of electrochromism of nanocrystalline thin film nanocrystalline electrodes highly attractive electrochromic
TiO2 electrodes have been reported. The first type is the elements. The principle of efficiency relies on fast interfacial
electrochromism of nanocrystalline TiO2 electrodes in Li- electron transfer between the nanocrystalline TiO2 and the
containing electrolytes related to the reversible insertion of adsorbed modifier as well as on the high surface area of the
Li+ into the anatase lattice of the nanoparticles.912 Hagfeldt TiO2 support that amplifies optical phenomena by 2 or 3
et al. found that forward biasing of transparent nanocrystal- orders of magnitude.902 The investigated TiO2 nanocrystalline
line TiO2 films in lithium ion-containing organic electrolytes electrodes include ordered905 and disordered902-904 mesopo-
led to rapid and reversible coloration due to electron rous films. Ordered mesoporous nanocrystalline TiO2 elec-
accumulation and Li+ intercalation in the anatase lattice.912 trodes were found to display enhanced color contrast yet have
Absorption of >90% light throughout the visible and near similar conduction band edge energy levels and electron
IR could be switched on and off within a few seconds. The percolation ability as electrodes made from nanocrystalline
nanocrystalline morphology of the film played a role in TiO2, attributed to the uniform and ordered mesopore
enhancing the electrochromic process. Ottaviani et al. found architecture and the large accessible surface area for tethering
that the rate of the electrochromic process was controlled viologen molecules.905
by the diffusion of the Li+ ions throughout the TiO2 lattice.914
5.4.2. Electrochromophore for an Electrochromic Device
It was convenient to drive the electrochromic process with
potentiostatic pulses, and under these conditions, many cycles The viologen group (N,N′-disubstituted-4,4′-bipyridinium)
with initially good color contrast and efficiencies which has been commonly chosen as an electrochromophore, for
approached 100% were obtained with TiO2 thin film its remarkable stability in both the oxidized and the reduced
electrodes. (radical cation) states (Figure 86).902,904,906,907 Oxidized vi-
2944 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 84. (A) Principle of the electrochromism of a molecular monolayer adsorbed on a semiconductor surface. Electrons are injected
from the conducting substrate into the conduction band of the semiconductor and from there reduce the adsorbed electroactive molecule.
Provided the redox potential of that molecule lies above the conduction band edge, the process is reversible by application of a positive
potential to the conductive substrate. (B) Nanocrystalline electrochromic devices based on viologen (solvent: glutarodinitrile) with a
counterelectrode made of Prussian blue, in the colorless and in the colored state. Reprinted from Bonhote, P.; Gogniat, E.; Campus, F.;
Walder, L.; Grätzel, M. Displays 1999, 20, 137, Copyright 1999, with permission from Elsevier.

ologen is colorless, while the radical cation can be blue, triarylamminum radical cation, which is accompanied by a
violet, purple, or green, depending on the substituents. The blue coloration with the absorption band at 730 nm.903
associated first reduction potential is between 0.2 and -0.6 Vayssieres et al. studied bis(phthalocyaninato)lutetium(III)
V (vs NHE). The typical absorption spectrum of reduced complexes (Pc2Lu) as electrochromophores, and they found
N,N′-dialkylviologen in an organic solvent has a maximum that the typical neutral green state of Pc2Lu was reduced to
around 600 nm. With N,N′-diarylviologens, the absorption a brown state at potentials < -0.3 V vs Ag/AgCl at neutral
band is shifted by about 50 nm to the red. In concentrated pH when Pc2Lu was adsorbed onto a nanostructured TiO2
solution or in the solid state, viologen radical cations form electrode.916
dimers, with their blue-shifted absorption maximum in the Ag-TiO2 films, prepared by loading nanoporous films
550 nm region. A second reduced state can be reached at with Ag nanoparticles by photocatalytic means, exhibited
potentials which are more negative by 0.2-0.4 V. This state multicolor photochromism, which was related to the oxida-
is neutral and almost colorless (yellowish). This second tion and reduction of Ag nanoparticles under UV and visible
reduction is reversible in organic solvents like acetonitrile radiation.773 Please also see section 4.2.1.2 on Sensitization
but not in water. The anchoring groups with strong affinity by Metal Nanoparticles.
toward TiIV include carboxylates, salicylates, or phospho-
nates.902 5.4.3. Counterelectrode for an Electrochromic Device
Bonhote et al. examined phosphonated triarylamine as an Closed cells are built by combining a transparent nano-
electrochromophore due to its oxidation by the stable crystalline electrode with a counterelectrode able to provide
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2945

Figure 85. Principle of signal amplification by a TiO2 nanocrystalline film. Sintered 20-nm particles of TiO2 form a several millimeter
thick film characterized by a very high surface area. Once derivatized with a molecular adsorbate, the structure contains the equivalent of
hundreds of superposed monolayers. Reprinted from Bonhote, P.; Gogniat, E.; Campus, F.; Walder, L.; Grätzel, M. Displays 1999, 20, 137,
Copyright 1999, with permission from Elsevier.

ing glass and modified by the electrochromophore [β-(10-


phenothiazyl)propoxy]phosphonic acid, which displayed
cycles-switching times of <250 ms, a coloration efficiency
of 270 cm2 C-1, and steady-state currents of <6 mA cm-2.908
Zinc can be used as a counterelectrode instead of PB for
displays applications.902 When the two electrodes are short-
circuited, the electrons flow from the zinc, which oxidizes
to Zn2+ ions, to the viologens of the nanocrystalline electrode.
The process can be reversed under a potential of 1-2 V.
Figure 86. Electrochromism mechanism of a viologen chro-
mophore. Reprinted from Bonhote, P.; Gogniat, E.; Campus, F.; 5.4.4. Photoelectrochromic Devices
Walder, L.; Grätzel, M. Displays 1999, 20, 137, Copyright 1999,
with permission from Elsevier. Pichot et al. demonstrated a photoelectrochromic smart
window with flexible substrates and solid-state electrolytes
enough electrons to allow complete reduction of viologen. based on a dye-sensitized TiO2 electrode spin-coated onto
The simplest counterelectrode is conducting glass. In-Sn oxide-coated polyester substrates coupled with a WO3
electrochronic counterelectrode, separated by a cross-linked
Prussian blue (PB) is an inorganic polymeric material (iron
polymer electrolyte containing LiI (Figure 87).634 The devices
phexacyanoferrate) that is commonly used on a conducting
typically transmitted 75% of visible light in the bleached
glass substrate as a counterelectrode. Being blue in the
state. After a few minutes of exposure to white light, the
oxidized state and colorless in the reduced state, it is a
windows turned dark blue, transmitting only 30% of visible
suitable complementary electrochromic material to the
light. They spontaneously bleached back to their initial
nanocrystalline viologen electrode. When the latter electrode
noncolored state upon removal of the light source. The
turns blue by reduction, the PB counterelectrode turns blue
photoelectrochromic device ideally behaved like a capaci-
by oxidation.902,904 Bonhote et al. studied nanocrystalline
tor: There was initially no mobile oxidized species (i.e., I2)
WO3 films as counterelectrodes for electrochromic applica-
present in the electrolyte. A schematic representation of the
tions since they turn from colorless to blue by reduction and
components and the electron and ion transfers in the solid-
lithium ion insertion.903
state photoelectrochromic device is shown in Figure 87. The
Fitzmaurice et al. constructed an electrochromic window ultimate electron acceptor (WO3) is localized as an insoluble
based on a modified transparent nanostructured TiO2 film material on the back electrode. Only the electron donor (I-),
supported on conducting glass and modified with the which serves as a regenerator to the oxidized dye, is initially
electrochromophore bis(2-phosphonoethyl)-4,4′-bipyridinium present in the electrolyte introduced as LiI. Upon coloration
dichloride, the electrolyte LiClO4, and ferrocene in γ-butyro- of the device at short-circuiting under illumination, I2 is
lactone.906 They used a counterelectrode of conducting glass, generated at the TiO2 electrode and Li+ intercalates in WO3.
which had excellent electrochromic performance with a
coloration efficiency of 170 cm2 C-1 at 608 nm, a switching 5.5. Hydrogen Storage
time of 1 s, and stability over 10,000 steady test cycles. They
upgraded this system with a counterelectrode based on a Lim et al. found that TiO2 nanotubes could reproducibly
transparent nanostructured SnO2 film supported on conduct- store up to approximately 2 wt % H2 at room temperature
2946 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Figure 87. Schematic representation of the components and the


electron and ion transfers in a solid-state photoelectrochromic
device. Upon illumination, electrons are injected into the TiO2
conduction band (CB), travel through the external circuit, and reduce
the WO3 counterelectrode. Upon lithium ion intercalation into the
reduced WO3 film, a “tungsten bronze” is formed that absorbs
visible and near-IR radiation, presumably via an intervalence
charge-transfer absorption. TCO stands for transparent conducting
oxide, which is typically ITO or fluorine-doped tin oxide. From:
Pichot, F.; Ferrere, S.; Pitts, R. J.; Gregg, B. A. J. Electrochem.
Soc. 1999, 146, 4324, Copyright 1999. Reproduced by permission
of The Electrochemical Society, Inc.

and 6 MPa.165 About 75% of this stored hydrogen could be Figure 88. (a) Pressure-concentration isotherms of TiO2 nanotubes
released when the hydrogen pressure was lowered to ambient and bulk TiO2 at room temperature. (b) Pressure-concentration
conditions due to physisorption. Approximately 13% was isotherms of TiO2 nanotubes at 24, 70, and 120 °C. Reprinted with
weakly chemisorbed and could be released at 70 °C as H2, permission from Lim, S. H.; Luo, J.; Zhong, Z.; Ji, W.; Lin, J.
Inorg. Chem. 2005, 44, 4124. Copyright 2005 American Chemical
and approximately 12% was strongly bonded to oxide ions Society.
and released only at temperatures above 120 °C as H2O. The
P-C isotherms of TiO2 nanotubes are shown in Figure 88.
At room temperature and a pressure of ∼900 psi (6 MPa),
the atomic ratio H/TiO2 was ∼1.6, corresponding to ∼2.0
wt % H2 for TiO2 nanotubes, compared to a much lower
hydrogen concentration of ∼0.8 wt % for bulk TiO2. When
the pressure was reduced, only ∼75% of the stored hydrogen
could be released, whereas 25% of adsorbed hydrogen
molecules were retained due to chemical adsorption.
Bavykin et al. studied the sorption of hydrogen between
the layers of the multilayered wall of nanotubular TiO2 in
the temperature range of -195 to 200 °C and at pressures
of 0 to 6 bar.157 Hydrogen could intercalate between layers
in the walls of TiO2 nanotubes forming host-guest com-
pounds TiO2‚xH2, where x e 1.5 and decreases at higher
temperature. The rate of hydrogen uptake increased with
temperature, and the characteristic time for hydrogen sorption Figure 89. Isotherm for (9) hydrogen sorption into and (O)
in TiO2 nanotubes was several hours at 100 °C. The hydrogen desorption out of the pores of TiO2 nanotubes at -196 °C. Reprinted
adsorption isotherm for TiO2 nanotubes at -195 °C is shown with permission from Bavykin, D. V.; Lapkin, A. A.; Plucinski, P.
in Figure 89. Almost 1.5 hydrogen molecules per one Ti K.; Friedrich, J. M.; Walsh, F. C. J. Phys. Chem. B 2005, 109,
19422. Copyright 2005 American Chemical Society.
atom could be adsorbed at a hydrogen partial pressure of 2
bar. During the desorption of hydrogen, a large hysteresis of the square of nanotube length from their proposed
was observed; even at 0 bar of pressure, the uptake of diffusion model.
hydrogen achieved a 1.25 molar ratio (point B). The Recently, Xu et al. studied the hydrogen storage properties
adsorption of hydrogen was a reversible process. Heating of a series of five pristine micro- and mesoporous Ti oxide
the sample in a vacuum to 200 °C led to a complete materials, synthesized from C6, C8, C10, C12, and C14
desorption of hydrogen, returning the weight of the sample amine templates possessing BET surface areas ranging from
to its initial value (point A). The author found that the 643 to 1063 m2/g, and they found that at 77 K the isotherms
diffusion of hydrogen molecules in the axial direction for all materials gently rose sharply at low pressure and
between the layers in multilayered walls of TiO2 nanotubes continued to rise in a linear fashion from 10 atm onward to
was the rate-limiting step of the process of intercalation and 65 atm and then return on desorption without significant
the rate of hydrogen intercalation depended on the inverse hysteresis. Extrapolation to 100 atm could yield total storage
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2947

Figure 90. Plot of real-time variation of resistance before, during, and after cleaning the contaminant, motor oil 10W-30, with UV exposure.
The plot, broken into four parts for clarity, shows (a) the original sensor behavior from time 10 to 1000 s, (b) the behavior of the sensor
over time 100-6000 s, during which the sensor is contaminated with oil, losing its hydrogen-sensing capabilities, and is initially exposed
to UV light, and (c) the behavior of the sensor from time 5000 s to 45,000 s. At time 7000 s, the UV is turned off, with the sensor regaining
its nominal starting resistance of approximately 100,000 Ω, at which point it is exposed to 1000 ppm hydrogen and its resistance changes
by a factor of approximately 50. The sensor is then again exposed to UV, from roughly time 15,000 s to 29,000 s. After this second UV
exposure, the sensor is again exposed to 1000 ppm hydrogen, showing an approximate factor of 500 change in electrical resistance. The
sensor is once again exposed to UV, from time 36,000 s. (d) Sensor behavior from time 45,000 to 70,000 s continues with UV exposure
of the sensor until time 52,000 s, after which the sensor is repeatedly cycled between air and 1000 ppm hydrogen, showing a relative
change in impedance of approximately 1000×. Compared to the hydrogen sensitivity of a noncontaminated sensor, the relative response of
the “recovered” sensor is within a factor of 2. Reprinted with permission from Mor, G. K.; Carvalho, M. A.; Varghese, O. K.; Pishko, M.
V.; Grimes, C. A. J. Mater. Res. 2004, 19, 628. Copyright 2004 Materials Research Society.

values as high as 5.36 wt % and 29.37 kg/m3, and surface that the resistance of anatase TiO2 varied in the presence of
Ti reduction by the appropriate organometallic reagent CO and H2 at temperatures above 500 °C, but on doping
provided an increase in performance, possibly because of a with 10% alumina it became selective for hydrogen.922
Kubas-type interaction.920 Shimizu et al. reported that anodized nanoporous titania films
with a Pd Schottky barrier were sensitive to hydrogen at 250
5.6. Sensing Applications °C.951,952 Kobayashi et al. investigated the mechanism of
hydrogen sensing by Pd/TiO2 Schottky diodes, and they
TiO2 nanocrystalline films have been widely studied as found that the formation of adsorbed water from adsorbed
sensors for various gases (refs 194, 196, 202, 206-209, 211, oxygen at the Pd/TiO2 interface was the dominant reaction
322, 745-747, 750, 751, and 921-961). Grimes et al.
for the Pd/TiO2(001) diodes throughout the hydrogen con-
conducted a series of excellent studies on sensing using TiO2
centration range of 0-3000 ppm; for the Pd/TiO2(100)
nanotubes.194,196,206-209,211 They found that TiO2 nanotubes
diodes, this reaction was dominant only for hydrogen
were excellent room-temperature hydrogen sensors not only
concentrations below 100 ppm and the hydrogen adsorption
with a high sensitivity of 104 but also with an ability to self-
clean photoactively after environmental contamination.196 At on bare Pd atoms became dominant for higher hydrogen
24 °C, in response to 1000 ppm of hydrogen, the sensors concentrations.939,940 Carney et al. found that sensors based
showed a fully reversible change in electrical resistance of on SnO2-TiO2 with higher surface areas were more sensitive
approximately 175,000%. The hydrogen-sensing capabilities to H2 in the presence of O2 by measuring the change in the
of the sensors were largely recovered by ultraviolet (UV) electrical resistance of the sensor upon exposure to different
light exposure after being completely extinguished by a rather hydrogen concentrations under a constant hydrogen gas flow
extreme means of sensor contamination: immersion of the rate.923 Devi et al. found that ordered mesoporous TiO2
sensor in motor oil. Figure 90 shows a plot of real-time exhibited higher H2 and CO sensitivities than sensors made
variation of resistance before, during, and after cleaning the from common TiO2 powders due to increased surface area,
contaminant, motor oil 10W-30, with UV exposure. and the sensitivity could be further improved by loading the
Many types of TiO2 nanomaterial-based room-temperature sensor with 0.5 mol % Nb2O5.929 Gao et al. found that
hydrogen sensors are based on Schottky barrier modulation nanoscale TiO2 displayed higher performance in H2 sensing
of devices like Pd/TiO2 or Pt/TiO2.922,947,954 Elevated tem- than microscale TiO2 due to larger surface area.932
perature hydrogen sensors examine the electrical resistance Oxygen sensors based on TiO2 nanomaterials include
change with hydrogen concentration. Birkefeld et al. found TiO2-x,961 TiO2-Nb2O5,928 CeO2-TiO2,957 and Ta-,935
2948 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

Ruiz et al. found that La-doped TiO2 nanoparticles were good


sensing materials for ethanol based on electrical resistance,746
while Cu- or Co-doped TiO2 nanoparticles were good
candidates for CO sensing.747 Garzella et al. found W-doped
TiO2 displayed better performance for ethanol sensing than
pure TiO2.934 The addition of Ta and Nb to TiO2 was
beneficial for stabilization of the nanophase, resulting in
selectivity enhancement toward CO930,931 and NO2.930,960
Comini et al. found that the sensitivity enhancement toward
ethanol and methanol of TiO2 films could be improved when
doped with Pt and Nb.926
Benkstein and Semancik found that mesoporous TiO2
nanoparticle thin films prepared on MEMS micro-hot-plate
platforms could be used as high-sensitivity conductometric
gas sensor materials.921 The nanoparticle films were deposited
onto selected micro-hot-plates in a multielement array via
microcapillary pipet and were sintered using the micro-hot-
plate. Figure 91A shows the conductometric response of four
TiO2 nanoparticle films. The relative thickness of the films
was varied by using one, two, three, or four drops of 6%
mass fraction TiO2 to cast the film. Sensitivity was defined
as the ratio of the film conductance in the presence of an
analyte to the baseline conductance measured in dry air (S
) G/G0). The thicker films showed a higher baseline
conductance and a higher overall sensitivity to methanol (G/
G0(1 drop) ) 4.1, G/G0(4 drops) ) 7.5). Shown in Figure
91B are sensitivity responses to µmol/mol levels of methanol
of a mesoporous TiO2 nanoparticle film and a CVD TiO2
film. The nanoparticle films were found to demonstrate
higher sensitivity to target analytes, attributed to the high
internal surface area of the porous nanoparticle films.
Montesperelli et al. found K-doped TiO2 nanocrystalline
films showed high sensitivity of magnitude of 107 with great
stability over time.943 Yadav et al. fabricated TiO2 nano-
Figure 91. (A) Conductance response of TiO2 nanoparticle films
to methanol in concentrations going from zero (baseline) to 50 crystalline films as optical humidity sensors based on the
µmol/mol in 10 µmol/mol steps, sequentially, at a sensor temper- variations in the intensity of light with in humidity changes.750
ature of 450 °C. Between the steps, the analyte concentration was The sensor element consisted of a thin U-shaped borosil glass
returned to 0 µmol/mol. The thickness of the films was controlled rod with a film of TiO2 deposited on it. Both the ends of the
by the number of drops of the nanoparticle dispersion used to glass rod were coupled to optical fibers. Light from a He-
deposit the film. (B) Comparison of the sensitivity of a TiO2 thin Ne laser was launched into the sensing element through one
film deposited by CVD (solid line) and that of a TiO2 thin film
composed of 15-nm-diameter anatase nanoparticles (dotted line). of them. Light received from the other fiber was fed into an
Reprinted from Benkstein, K. D.; Semancik, S. Sens. Actuators, B optical power meter.
2006, 113, 445, Copyright 2006, with permission from Else-
vier. 6. Summary
Nb-,949 949 950
Cr-, and Pt-doped TiO2. Pt-doped TiO2 sensors Over the past decades, the tremendous effort put into TiO2
showed improved gas sensitivity, low operation temperature nanomaterials has resulted in a rich database for their
(350-800 °C), and short response time (<0.1 s).941,950,958 The synthesis, properties, modifications, and applications. The
oxygen-sensing mechanism was the combination of Pt/TiO2 continuing breakthroughs in the synthesis and modifications
interfaces in a Schottky-barrier mechanism and an oxygen- of TiO2 nanomaterials have brought new properties and new
vacancy bulk effect mechanism.959 At high temperatures, applications with improved performance. Accompanied by
TiO2 devices can be used as thermodynamically controlled the progress in the synthesis of TiO2 nanoparticles are new
bulk defect sensors to detect oxygen over a large range of findings in the synthesis of TiO2 nanorods, nanotubes,
partial pressures; at low temperatures, Pt/TiO2 Schottky nanowires, as well as mesoporous and photonic structures.
diodes make extremely sensitive oxygen detection possible.938 Besides the well-know quantum-confinement effect, these
In Ta-doped TiO2 sensors, oxygen vacancies formed by new nanomaterials demonstrate size-dependent as well as
photoirradiation acted as oxygen-sensing sites.935 Sotter et shape- and structure-dependent optical, electronic, thermal,
al. found Nb-doped TiO2 nanomaterials to be good sensor and structural properties. TiO2 nanomaterials have continued
materials for O2.953 Nb- and Cr-doped TiO2 nanocrystalline to be highly active in photocatalytic and photovoltaic
films displayed higher O2 sensitivity than pure TiO2 films949 applications, and they also demonstrate new applications
in that Nb5+-doped TiO2 showed 65 times enhancement in including electrochromics, sensing, and hydrogen storage.
the sensitivity compared to undoped material at a lower This steady progress has demonstrated that TiO2 nanoma-
operating temperature.474 terials are playing and will continue to play an important
TiO2 nanomaterials are promising candidates for CO role in the protections of the environment and in the search
sensing927,945 and for methanol and ethanol sensing.933,946,955,956 for renewable and clean energy technologies.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2949

7. Acknowledgment (51) Bessekhouad, Y.; Robert, D.; Weber, J. V. J. Photochem. Photobiol.,


A 2003, 157, 47.
The authors acknowledge the financial support from (52) Chemseddine, A.; Moritz, T. Eur. J. Inorg. Chem. 1999, 235.
(53) Kim, K. D.; Kim, H. T. Powder Technol. 2001, 119, 164.
Lawrence Berkeley National Laboratory and the U.S. (54) Kim, K. D.; Kim, H. T. Colloids Surf., A 2002, 207, 263.
Department of Energy. We thank M. C. Schriver and K. R. (55) Kuznetsova, I. N.; Blaskov, V.; Stambolova, I.; Znaidi, L.; Kanaev,
Carrington for critical reading of the manuscript. A. Mater. Lett. 2005, 59, 3820.
(56) Lee, J. H.; Yang, Y. S. Mater. Chem. Phys. 2005, 93, 237.
(57) Lee, J. H.; Yang, Y. S. J. Eur. Ceram. Soc. 2005, 25, 3573.
8. References (58) Li, Y.; White, T. J.; Lim, S. H. J. Solid State Chem. 2004, 177, 1372.
(59) Liu, P.; Bandara, J.; Lin, Y.; Elgin, D.; Allard, L. F.; Sun, Y. P.
(1) Pfaff, G.; Reynders, P. Chem. ReV. 1999, 99, 1963. Langmuir 2002, 18, 10398.
(2) Salvador, A.; Pascual-Marti, M. C.; Adell, J. R.; Requeni, A.; March, (60) Reddy, K. M.; Reddy, C. V. G.; Manorama, S. V. J. Solid State
J. G. J. Pharm. Biomed. Anal. 2000, 22, 301. Chem. 2001, 158, 180.
(3) Zallen, R.; Moret, M. P. Solid State Commun. 2006, 137, 154. (61) Manorama, S. V.; Reddy, K. M.; Reddy, C. V. G.; Narayanan, S.;
(4) Braun, J. H.; Baidins, A.; Marganski, R. E. Prog. Org. Coat. 1992, Raja, P. R.; Chatterji, P. R. J. Phys. Chem. Solids 2001, 63, 135.
20, 105. (62) Moritz, T.; Reiss, J.; Diesner, K.; Su, D.; Chemseddine, A. J. Phys.
(5) Yuan, S. A.; Chen, W. H.; Hu, S. S. Mater. Sci. Eng. C 2005, 25, Chem. B 1997, 101, 8052.
479. (63) Oskam, G.; Nellore, A.; Penn, R. L.; Searson, P. C. J. Phys. Chem.
(6) Fujishima, A.; Honda, K. Nature 1972, 37, 238. B 2003, 107, 1734.
(7) Fujishima, A.; Rao, T. N.; Tryk, D. A. J. Photochem. Photobiol. C (64) Pottier, A.; Chaneac, C.; Tronc, E.; Mazerolles, L.; Jolivet, J. P. J.
2000, 1, 1. Mater. Chem. 2001, 11, 1116.
(8) Tryk, D. A.; Fujishima, A.; Honda, K. Electrochim. Acta 2000, 45, (65) Pottier, A.; Cassaignon, S.; Chaneac, C.; Villain, F.; Tronc, E.; Jolivet,
2363. J. P. J. Mater. Chem. 2003, 13, 877.
(9) Grätzel, M. Nature 2001, 414, 338. (66) Sugimoto, T. AdV. Colloid Interface Sci. 1987, 28, 65.
(10) Hagfeldt, A.; Grätzel, M. Chem. ReV. 1995, 95, 49. (67) Sugimoto, T.; Okada, K.; Itoh, H. J. Colloid Interface Sci. 1997,
(11) Linsebigler, A. L.; Lu, G.; Yates, J. T., Jr. Chem. ReV. 1995, 95, 193, 140.
735. (68) Sugimoto, T.; Zhou, X. J. Colloid Interface Sci. 2002, 252, 347.
(12) Millis, A.; Le Hunte, S. J. Photochem. Photobiol., A 1997, 108, 1. (69) Sugimoto, T.; Zhou, X.; Muramatsu, A. J. Colloid Interface Sci. 2002,
(13) Alivisatos, A. P. J. Phys. Chem. 1996, 100, 13226. 252, 339.
(14) Alivisatos, A. P. Science 1996, 271, 933. (70) Sugimoto, T.; Zhou, X.; Muramatsu, A. J. Colloid Interface Sci. 2003,
(15) Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chem. ReV. 259, 53.
2005, 105, 1025. (71) Sugimoto, T.; Zhou, X.; Muramatsu, A. J. Colloid Interface Sci. 2003,
(16) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Annu. ReV. Mater. 259, 43.
Sci. 2000, 30, 545. (72) Uekawa, N.; Kajiwara, J.; Kakegawa, K.; Sasaki, Y. J. Colloid
(17) Yin, Y.; Alivisatos, A. P. Nature 2005, 437, 664. Interface Sci. 2002, 250, 285.
(18) Adams, D. M.; Brus, L.; Chidsey, C. E. D.; Creager, S.; Creutz, C.; (73) Zhang, H.; Banfield, J. F. J. Mater. Chem. 1998, 8, 2073.
Kagan, C. R.; Kamat, P. V.; Lieberman, M.; Lindsay, S.; Marcus, (74) Zhang, H.; Banfield, J. F. J. Phys. Chem. B 2000, 104, 3481.
R. A.; Metzger, R. M.; Michel-Beyerle, M. E.; Miller, J. R.; Newton, (75) Zhang, H.; Finnegan, M.; Banfield, J. F. Nano Lett. 2001, 1, 81.
M. D.; Rolison, D. R.; Sankey, O.; Schanze, K. S.; Yardley, J.; Zhu, (76) Zhang, H.; Banfield, J. F. Chem. Mater. 2002, 14, 4145.
X. J. Phys. Chem. B 2003, 107, 6668. (77) Zhang, H.; Banfield, J. F. Chem. Mater. 2005, 17, 3421.
(19) Asbury, J. B.; Hao, E.; Wang, Y.; Ghosh, H. N.; Lian, T. J. Phys. (78) Znaidi, L.; Seraphimova, R.; Bocquet, J. F.; Colbeau-Justin, C.;
Chem. B 2001, 105, 4545. Pommier, C. Mater. Res. Bull. 2001, 36, 811.
(20) Beydoun, D.; Amal, R.; Low, G.; McEvoy, S. J. Nanopart. Res. 1999, (79) Anderson, M. A.; Gieselmann, M. J.; Xu, Q. J. Membr. Sci. 1988,
1, 439. 39, 243.
(21) Chen, X.; Lou, Y.; Dayal, S.; Qiu, X.; Krolicki, R.; Burda, C.; Zhao, (80) Barbe, C. J.; Arendse, F.; Comte, P.; Jirousek, M.; Lenzmann, F.;
C.; Becker, J. Nanosci. Nanotechnol. 2005, 5, 1408. Shklover, V.; Grätzel, M. J. Am. Ceram. Soc. 1997, 80, 3157.
(22) Dai, H. Acc. Chem. Res. 2002, 35, 1035. (81) Barringer, E. A.; Bowen, H. K. Langmuir 1985, 1, 414.
(23) Dresselhaus, M. S.; Dresselhaus, G. Annu. ReV. Mater. Sci. 1995, (82) Barringer, E. A.; Bowen, H. K. Langmuir 1985, 1, 420.
25, 487. (83) Jean, J. H.; Ring, T. A. Langmuir 1986, 2, 251.
(24) Dresselhaus, M. S.; Dresselhaus, G.; Jorio, A.; Souza Filho, A. G.; (84) Kormann, C.; Bahnemann, D. W.; Hoffmann, M. R. J. Phys. Chem.
Pimenta, M. A.; Saito, R. Acc. Chem. Res. 2002, 35, 1070. 1988, 92, 5196.
(25) Dresselhaus, M. S.; Dresselhaus, G.; Jorio, A. Annu. ReV. Mater. (85) Livage, J.; Henry, M.; Sanchez, C. Prog. Solid State Chem. 1988,
Sci. 2004, 34, 247. 18, 259.
(26) Efros, Al. L.; Rosen, M. Annu. ReV. Mater. Sci. 2000, 30, 475. (86) Look, J. L.; Zukoski, C. F. J. Am. Ceram. Soc. 1995, 78, 21.
(27) El-Sayed, M. A. Acc. Chem. Res. 2004, 37, 326. (87) Look, J. L.; Zukoski, C. F. J. Am. Ceram. Soc. 1992, 75, 1587.
(28) Grätzel, M. Prog. PhotoVolt. 2000, 8, 171. (88) O’Regan, B.; Grätzel, M. Nature 1991, 353, 737.
(29) Grätzel, M. J. Sol-Gel Sci. Technol. 2001, 22, 7. (89) Penn, R. L.; Banfield, J. F. Geochim. Cosmochim. Acta 1999, 63,
(30) Grätzel, M. J. Photochem. Photobiol., C 2003, 4, 145. 1549.
(31) Grätzel, M. J. Photochem. Photobiol., A 2004, 164, 3. (90) Vorkapic, D.; Matsoukas, T. J. Am. Ceram. Soc. 1998, 81, 2815.
(32) Grätzel, M. MRS Bull. 2005, 30, 23. (91) Vorkapic, D.; Matsoukas, T. J. Colloid Interface Sci. 1999, 214, 283.
(33) Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W. Chem. (92) Miao, L.; Tanemura, S.; Toh, S.; Kaneko, K.; Tanemura, M. J. Cryst.
ReV. 1995, 95, 69. Growth 2004, 264, 246.
(34) Link, S.; El-Sayed, M. A. Annu. ReV. Phys. Chem. 2003, 54, 331. (93) Miao, L.; Tanemura, S.; Toh, S.; Kaneko, K.; Tanemura, M. Appl.
(35) Mao, S. S. Int. J. Nanotechnol. 2004, 1, 42. Surf. Sci. 2004, 238, 175.
(36) Nirmal, M.; Brus, L. Acc. Chem. Res. 1999, 32, 407. (94) Lin, Y.; Wu, G. S.; Yuan, X. Y.; Xie, T.; Zhang, L. D. J. Phys.:
(37) Nozik, A. J. Inorg. Chem. 2005, 44, 6893. Condens. Matter 2003, 15, 2917.
(38) Nozik, A. J. Annu. ReV. Phys. Chem. 2001, 52, 193. (95) Chen, Y.; Crittenden, J. C.; Hackney, S.; Sutter, L.; Hand, D. W.
(39) Ouyang, M.; Huang, J.-L.; Lieber, C. M. Acc. Chem. Res. 2002, 35, EnViron. Sci. Technol. 2005, 39, 1201.
1018. (96) Lee, S.; Jeon, C.; Park, Y. Chem. Mater. 2004, 16, 4292.
(40) Vayssieres, L. Int. J. Nanotechnol. 2004, 1, 1. (97) Liu, S. M.; Gan, L. M.; Liu, L. H.; Zhang, W. D.; Zeng, H. C. Chem.
(41) Cushing, B. L.; Kolesnichenko, V. L.; O’Connor, C. J. Chem. ReV. Mater. 2002, 14, 1391.
2004, 104, 3893. (98) Sander, M. S.; Cote, M. J.; Gu, W.; Kile, B. M.; Tripp, C. P. AdV.
(42) Eustis, S.; El-Sayed, M. A. Chem. Soc. ReV. 2006, 35, 209. Mater. 2004, 16, 2052.
(43) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2005, 109, 12663. (99) Jung, J. H.; Kobayashi, H.; van Bommel, K. J. C.; Shinkai, S.;
(44) El-Sayed, M. A. Acc. Chem. Res. 2001, 34, 257. Shimizu, T. Chem. Mater. 2002, 14, 1445.
(45) Fox, M. A.; Dulay, M. T. Chem. ReV. 1993, 93, 341. (100) Jung, J. H.; Shimizu, T.; Shinkai, S. J. Mater. Chem. 2005, 15, 3979.
(46) Pierre, A. C.; Pajonk, G. M. Chem. ReV. 2002, 102, 4243. (101) Qiu, J. J.; Yu, W. D.; Gao, X. D.; Li, X. M. Nanotechnology 2006,
(47) Lu, Z. L.; Lindner, E.; Mayer, H. A. Chem. ReV. 2002, 102, 3543. 17, 4695.
(48) Wight, A. P.; Davis, M. E. Chem. ReV. 2002, 102, 3589. (102) Hong, S. S.; Lee, M. S.; Park, S. S.; Lee, G. D. Catal. Today 2003,
(49) Schwarz, J. A.; Contescu, C.; Contescu, A. Chem. ReV. 1995, 95, 87, 99.
477. (103) Kim, K. D.; Kim, S. H.; Kim, H. T. Colloids Surf., A 2005, 254, 99.
(50) Hench, L. L.; West, J. K. Chem. ReV. 1990, 90, 33. (104) Li, G. L.; Wang, G. H. Nanostruct. Mater. 1999, 11, 663.
2950 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

(105) Li, Y.; Lee, N. H.; Hwang, D. S.; Song, J. S.; Lee, E. G.; Kim, S. J. (153) Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K.
Langmuir 2004, 20, 10838. Langmuir 1998, 14, 3160.
(106) Lim, K. T.; Hwang, H. S.; Ryoo, W.; Johnston, K. P. Langmuir 2004, (154) Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. AdV.
20, 2466. Mater. 1999, 11, 1307.
(107) Lim, K. T.; Hwang, H. S.; Hong, S. S.; Park, C.; Ryoo, W.; Johnston, (155) Bavykin, D. V.; Milsom, E. V.; Marken, F.; Kim, D. H.; Marsh, D.
K. P. Stud. Surf. Sci. Catal. 2004, 153, 569. H.; Riley, D. J.; Walsh, F. C.; El-Abiary, K. H.; Lapkin, A. A.
(108) Lin, J.; Lin, Y.; Liu, P.; Meziani, M. J.; Allard, L. F.; Sun, Y. P. J. Electrochem. Commun. 2005, 7, 1050.
Am. Chem. Soc. 2002, 124, 11514. (156) Bavykin, D. V.; Parmon, V. N.; Lapkin, A. A.; Walsh, F. C. J. Mater.
(109) Yu, J. C.; Yu, J.; Ho, W.; Zhang, L. Chem. Commun. 2001, 1942. Chem. 2004, 14, 3370.
(110) Zhang, D.; Qi, L.; Ma, J.; Cheng, H. J. Mater. Chem. 2002, 12, 3677. (157) Bavykin, D. V.; Lapkin, A. A.; Plucinski, P. K.; Friedrich, J. M.;
(111) Niederberger, M.; Bartl, M. H.; Stucky, G. D. Chem. Mater. 2002, Walsh, F. C. J. Phys. Chem. B 2005, 109, 19422.
14, 4364. (158) Bavykin, D. V.; Gordeev, S. N.; Moskalenko, A. V.; Lapkin, A. A.;
(112) Parala, H.; Devi, A.; Bhakta, R.; Fischer, R. A. J. Mater. Chem. Walsh, F. C. J. Phys. Chem. B 2005, 109, 8565.
2002, 12, 1625. (159) Bavykin, D. V.; Lapkin, A. A.; Plucinski, P. K.; Friedrich, J. M.;
(113) Tang, J.; Redl, F.; Zhu, Y.; Siegrist, T.; Brus, L. E.; Steigerwald, M. Walsh, F. C. J. Catal. 2005, 235, 10.
L. Nano Lett. 2005, 5, 543. (160) Chien, S. H.; Liou, Y. C.; Kuo, M. C. Synth. Met. 2005, 152, 333.
(114) Arnal, P.; Corriu, R. J. P.; Leclercq, D.; Mutin, P. H.; Vioux, A. J. (161) Du, G. H.; Chen, Q.; Che, R. C.; Yuan, Z. Y.; Peng, L. M. Appl.
Mater. Chem. 1996, 6, 1925. Phys. Lett. 2001, 79, 3702.
(115) Arnal, P.; Corriu, R. J. P.; Leclercq, D.; Mutin, P. H.; Vioux, A. (162) Gundiah, G.; Mukhopadhyay, S.; Tumkurkar, U. G.; Govindaraj, A.;
Chem. Mater. 1997, 9, 694. Maitra, U.; Rao, C. N. R. J. Mater. Chem. 2003, 13, 2118.
(116) Hay, J. N.; Raval, H. M. J. Sol-Gel Sci. Technol. 1998, 13, 109. (163) Kukovecz, A.; Hodos, M.; Konya, Z.; Kiricsi, I. Chem. Phys. Lett.
(117) Hay, J. N.; Raval, H. M. Chem. Mater. 2001, 13, 3396. 2005, 411, 445.
(118) Lafond, V.; Mutin, P. H.; Vioux, A. Chem. Mater. 2004, 16, 5380. (164) Lan, Y.; Gao, X.; Zhu, H.; Zheng, Z.; Yan, T.; Wu, F.; Ringer, S.
(119) Trentler, T. J.; Denler, T. E.; Bertone, J. F.; Agrawal, A.; Colvin, V. P.; Song, D. AdV. Funct. Mater. 2005, 15, 1310.
L. J. Am. Chem. Soc. 1999, 121, 1613. (165) Lim, S. H.; Luo, J.; Zhong, Z.; Ji, W.; Lin, J. Inorg. Chem. 2005,
(120) Scolan, E.; Sanchez, C. Chem. Mater. 1998, 10, 3217. 44, 4124.
(121) Cozzoli, P. D.; Comparelli, R.; Fanizza, E.; Curri, M. L.; Agostiano, (166) Ma, R.; Fukuda, K.; Sasaki, T.; Osada, M.; Bando, Y. J. Phys. Chem.
A. Mater. Sci. Eng., C 2003, C23, 707. B 2005, 109, 6210.
(122) Cozzoli, P. D.; Fanizza, E.; Curri, M. L.; Laub, D.; Agostiano, A. (167) Qian, L.; Du, Z. L.; Yang, S. Y.; Jin, Z. S. J. Mol. Struct. 2005, 749,
Chem. Commun. 2005, 942. 103.
(123) Cozzoli, P. D.; Kornowski, A.; Weller, H. J. Am. Chem. Soc. 2003, (168) Seo, D. S.; Lee, J. K.; Kim, H. J. Cryst. Growth 2001, 229, 428.
125, 14539. (169) Tian, Z. R.; Voigt, J. A.; Liu, J.; McKenzie, B.; Xu, H. J. Am. Chem.
(124) Cozzoli, P. D.; Fanizza, E.; Comparelli, R.; Curri, M. L.; Agostiano, Soc. 2003, 125, 12384.
A.; Laub, D. J. Phys. Chem. B 2004, 108, 9623. (170) Wang, M.; Guo, D. J.; Li, H. L. J. Solid State Chem. 2005, 178,
(125) Cozzoli, P. D.; Comparelli, R.; Fanizza, E.; Curri, M. L.; Agostiano, 1996.
A.; Laub, D. J. Am. Chem. Soc. 2004, 126, 3868. (171) Wang, Y. Q.; Hu, G. Q.; Duan, X. F.; Sun, H. L.; Xue, Q. K. Chem.
(126) Cozzoli, P. D.; Curri, M. L.; Agostiano, A. Chem. Commun. 2005, Phys. Lett. 2002, 365, 427.
3186. (172) Yao, B. D.; Chan, Y. F.; Zhang, X. Y.; Zhang, W. F.; Yang, Z. Y.;
(127) Joo, J.; Kwon, S. G.; Yu, T.; Cho, M.; Lee, J.; Yoon, J.; Hyeon, T. Wang, N. Appl. Phys. Lett. 2003, 82, 281.
J. Phys. Chem. B 2005, 109, 15297. (173) Yuan, Z. Y.; Su, B. L. Colloids Surf., A 2004, 241, 173.
(128) Jun, Y. W.; Casula, M. F.; Sim, J. H.; Kim, S. Y.; Cheon, J.; (174) Miyauchi, M.; Tokudome, H.; Toda, Y.; Kamiya, T.; Hosono, H.
Alivisatos, A. P. J. Am. Chem. Soc. 2003, 125, 15981. Appl. Phys. Lett. 2006, 89.
(129) Zhang, Z.; Zhong, X.; Liu, S.; Li, D.; Han, M. Angew. Chem., Int.
(175) Bavykin, D. V.; Friedrich, J. M.; Walsh, F. C. AdV. Mater. 2006,
Ed. 2005, 44, 3466.
18, 2807.
(130) Buonsanti, R.; Grillo, V.; Carlino, E.; Giannini, C.; Curri, M. L.;
(176) Wang, W.; Varghese, O. K.; Paulose, M.; Grimes, C. A.; Wang, Q.;
Innocenti, C.; Sangregorio, C.; Achterhold, K.; Parak, F. G.;
Dickey, E. C. J. Mater. Res. 2004, 19, 417.
Agostiano, A.; Cozzoli, P. D. J. Am. Chem. Soc. 2006, 128, 16953.
(177) Li, X. L.; Peng, Q.; Yi, J. X.; Wang, X.; Li, Y. D. Chem.sEur. J.
(131) Andersson, M.; Oesterlund, L.; Ljungstroem, S.; Palmqvist, A. J.
2006, 12, 2383.
Phys. Chem. B 2002, 106, 10674.
(132) Chae, S. Y.; Park, M. K.; Lee, S. K.; Kim, T. Y.; Kim, S. K.; Lee, (178) Xu, J.; Ge, J. P.; Li, Y. D. J. Phys. Chem. B 2006, 110, 2497.
W. I. Chem. Mater. 2003, 15, 3326. (179) Wang, X.; Zhuang, J.; Peng, Q.; Li, Y. D. Nature 2005, 437, 121.
(133) Cot, F.; Larbot, A.; Nabias, G.; Cot, L. J. Eur. Ceram. Soc. 1998, (180) Wen, B.; Liu, C.; Liu, Y. Inorg. Chem. 2005, 44, 6503.
18, 2175. (181) Wen, B.; Liu, C.; Liu, Y. New J. Chem. 2005, 29, 969.
(134) Yang, J.; Mei, S.; Ferreira, J. M. F. Mater. Sci. Eng., C 2001, C15, (182) Wen, B.; Liu, C.; Liu, Y. J. Phys. Chem. B 2005, 109, 12372.
183. (183) Kim, C. S.; Moon, B. K.; Park, J. H.; Choi, B. C.; Seo, H. J. J.
(135) Yang, J.; Mei, S.; Ferreira, J. M. F. J. Am. Ceram. Soc. 2000, 83, Cryst. Growth 2003, 257, 309.
1361. (184) Kim, C. S.; Moon, B. K.; Park, J. H.; Chung, S. T.; Son, S. M. J.
(136) Yang, J.; Mei, S.; Ferreira, J. M. F. J. Am. Ceram. Soc. 2001, 84, Cryst. Growth 2003, 254, 405.
1696. (185) Yang, S. W.; Gao, L. Mater. Chem. Phys. 2006, 99, 437.
(137) Yang, J.; Mei, S.; Ferreira, J. M. F. J. Mater. Res. 2002, 17, 2197. (186) Wu, J. M. J. Cryst. Growth 2004, 269, 347.
(138) Yang, J.; Mei, S.; Ferreira, J. M. F. J. Colloid Interface Sci. 2003, (187) Wu, J. M.; Zhang, T. W.; Zeng, Y. W.; Hayakawa, S.; Tsuru, K.;
260, 82. Osaka, A. Langmuir 2005, 21, 6995.
(139) Yang, J.; Mei, S.; Ferreira, J. M. F. J. Eur. Ceram. Soc. 2003, 24, (188) Wu, J. M.; Hayakawa, S.; Tsuru, K.; Osaka, A. Cryst. Growth Des.
335. 2002, 2, 147.
(140) Yang, J.; Mei, S.; Ferreira, J. M. F. Mater. Sci. Forum 2004, 455- (189) Wu, J. M.; Hayakawa, S.; Tsuru, K.; Osaka, A. Scripta Mater. 2002,
456, 556. 46, 101.
(141) Zhang, Q.; Gao, L. Langmuir 2003, 19, 967. (190) Wu, J. M.; Hayakawa, S.; Tsuru, K.; Osaka, A. Scripta Mater. 2002,
(142) Feng, X.; Zhai, J.; Jiang, L. Angew. Chem., Int. Ed. 2005, 44, 5115. 46, 705.
(143) Huang, Q.; Gao, L. Chem. Lett. 2003, 32, 638. (191) Wu, J. M.; Zhang, T. W. J. Photochem. Photobiol., A 2004, 162,
(144) Yang, S.; Gao, L. Chem. Lett. 2005, 34, 972. 171.
(145) Yang, S.; Gao, L. Chem. Lett. 2005, 34, 1044. (192) Peng, X.; Chen, A. J. Mater. Chem. 2004, 14, 2542.
(146) Yang, S.; Gao, L. Chem. Lett. 2005, 34, 964. (193) Gong, D.; Grimes, C. A.; Varghese, O. K.; Hu, W.; Singh, R. S.;
(147) Armstrong, A. R.; Armstrong, G.; Canales, J.; Garcı́a, R.; Bruce, P. Chen, Z.; Dickey, E. C. J. Mater. Res. 2001, 16, 3331.
G. AdV. Mater. 2005, 7, 862. (194) Mor, G. K.; Varghese, O. K.; Paulose, M.; Grimes, C. A. Sens. Lett.
(148) Armstrong, A. R.; Armstrong, G.; Canales, J.; Garcı́a, R.; Bruce, P. 2003, 1, 42.
G. Angew. Chem., Int. Ed. 2004, 43, 2286. (195) Mor, G. K.; Varghese, O. K.; Paulose, M.; Mukherjee, N.; Grimes,
(149) Yoshida, R.; Suzuki, Y.; Yoshikawa, S. J. Solid State Chem. 2005, C. A. J. Mater. Res. 2003, 18, 2588.
178, 2179. (196) Mor, G. K.; Carvalho, M. A.; Varghese, O. K.; Pishko, M. V.; Grimes,
(150) Zhang, Y. X.; Li, G. H.; Jin, Y. X.; Zhang, Y.; Zhang, J.; Zhang, L. C. A. J. Mater. Res. 2004, 19, 628.
D. Chem. Phys. Lett. 2002, 365, 300. (197) Mor, G. K.; Shankar, K.; Varghese, O. K.; Grimes, C. A. J. Mater.
(151) Nian, J. N.; Teng, H. S. J. Phys. Chem. B 2006, 110, 4193. Res. 2004, 19, 2989.
(152) Wei, M.; Konishi, Y.; Zhou, H.; Sugihara, H.; Arakawa, H. Chem. (198) Mor, G. K.; Shankar, K.; Paulose, M.; Varghese, O. K.; Grimes, C.
Phys. Lett. 2004, 400, 231. A. Nano Lett. 2005, 5, 191.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2951

(199) Mor, G. K.; Varghese, O. K.; Paulose, M.; Shankar, K.; Grimes, C. (243) Oh, S. M.; Kim, S. S.; Lee, J. E.; Ishigaki, T.; Park, D. W. Thin
A. Mater. Res. Soc. Symp. Proc. 2005, 836, 29. Solid Films 2003, 435, 252.
(200) Mor, G. K.; Varghese, O. K.; Paulose, M.; Grimes, C. A. AdV. Funct. (244) Oh, S. M.; Ishigaki, T. Thin Solid Films 2004, 457, 186.
Mater. 2005, 15, 1291. (245) Oh, S. M.; Li, J. G.; Ishigaki, T. J. Mater. Res. 2005, 20, 529.
(201) Mor, G. K.; Shankar, K.; Paulose, M.; Varghese, O. K.; Grimes, C. (246) Wang, X. H.; Li, J. G.; Kamiyama, H.; Katada, M.; Ohashi, N.;
A. Nano Lett. 2006, 6, 215. Moriyoshi, Y.; Ishigaki, T. J. Am. Chem. Soc. 2005, 127, 10982.
(202) Paulose, M.; Varghese, O. K.; Mor, G. K.; Grimes, C. A.; Ong, K. (247) Nedeljkovic, J. M.; Saponjic, Z. V.; Rakocevic, Z.; Jokanovic, V.;
G. Nanotechnology 2006, 17, 398. Uskokovic, D. P. Nanostruct. Mater. 1997, 9, 125.
(203) Paulose, M.; Shankar, K.; Varghese, O. K.; Mor, G. K.; Hardin, B.; (248) Grujic-Brojcin, M.; Scepanovic, M. J.; Dohcevic-Mitrovic, Z. D.;
Grimes, C. A. Nanotechnology 2006, 17, 1. Hinic, I.; Matovic, B.; Stanisic, G.; Popovic, Z. V. J. Phys. D 2005,
(204) Ruan, C.; Paulose, M.; Varghese, O. K.; Mor, G. K.; Grimes, C. A. 38, 1415.
J. Phys. Chem. B 2005, 109, 15754. (249) Scepanovic, M.; Dohcevic-Mitrovic, Z. D.; Hinic, I.; Grujic-Brojcin,
(205) Shankar, K.; Paulose, M.; Mor, G. K.; Varghese, O. K.; Grimes, C. M.; Stanisic, G.; Popovic, Z. V. Mater. Sci. Forum 2005, 494, 265.
A. J. Phys. D 2005, 38, 3543. (250) Oh, C. W.; Seong, G. D. L.; Park, S.; Ju, C. S.; Hong, S. S. React.
(206) Varghese, O. K.; Gong, D.; Paulose, M.; Ong, K. G.; Grimes, C. A. Kinet. Catal. Lett. 2005, 85, 261.
Sens. Actuators, B 2003, B93, 338. (251) Oh, C. W.; Lee, G. D.; Park, S. S.; Ju, C. S.; Hong, S. S. Korean J.
(207) Varghese, O. K.; Grimes, C. A. J. Nanosci. Nanotechnol. 2003, 3, Chem. Eng. 2005, 22, 547.
277. (252) Wu, J. M.; Shih, H. C.; Wu, W. T. Chem. Phys. Lett. 2005, 413,
(208) Varghese, O. K.; Gong, D.; Paulose, M.; Ong, K. G.; Dickey, E. C.; 490.
Grimes, C. A. AdV. Mater. 2003, 15, 624.
(253) Wu, J. M.; Shih, H. C.; Wu, W. T.; Tseng, Y. K.; Chen, I. C. J.
(209) Varghese, O. K.; Gong, D.; Dreschel, W. R.; Ong, K. G.; Grimes, Cryst. Growth 2005, 281, 384.
C. A. Sens. Actuators, B 2003, B94, 27.
(254) Xiang, B.; Zhang, Y.; Wang, Z.; Luo, X. H.; Zhu, Y. W.; Zhang, H.
(210) Varghese, O. K.; Gong, D.; Paulose, M.; Grimes, C. A.; Dickey, E.
Z.; Yu, D. P. J. Phys. D 2005, 38, 1152.
C. J. Mater. Res. 2003, 18, 156.
(211) Varghese, O. K.; Mor, G. K.; Grimes, C. A.; Paulose, M.; Mukherjee, (255) Lei, Y.; Zhang, L. D.; Fan, J. C. Chem. Phys. Lett. 2001, 338, 231.
N. J. Nanosci. Nanotechnol. 2004, 4, 733. (256) Liu, S.; Huang, K. Sol. Energy Mater. Sol. Cells 2004, 85, 125.
(212) Varghese, O. K.; Paulose, M.; Shankar, K.; Mor, G. K.; Grimes, C. (257) Blesic, M. D.; Saponjic, Z. V.; Nedeljkovic, J. M.; Uskokovic, D. P.
A. J. Nanosci. Nanotechnol. 2005, 5, 1158. Mater. Lett. 2002, 54, 298.
(213) Paulose, M.; Shankar, K.; Varghese, O. K.; Mor, G. K.; Grimes, C. (258) Guo, W.; Lin, Z.; Wang, X.; Song, G. Microelectron. Eng. 2003,
A. J. Phys. D 2006, 39, 2498. 66, 95.
(214) Paulose, M.; Shankar, K.; Yoriya, S.; Prakasam, H. E.; Varghese, (259) Huang, W.; Tang, X.; Wang, Y.; Koltypin, Y.; Gedanken, A. Chem.
O. K.; Mor, G. K.; Latempa, T. A.; Fitzgerald, A.; Grimes, C. A. J. Commun. 2000, 1415.
Phys. Chem. B 2006, 110, 16179. (260) Jokanovic, V.; Spasic, A. M.; Uskokovic, D. J. Colloid Interface
(215) Paulose, M.; Shankar, K.; Varghese, O. K.; Mor, G. K.; Hardin, B.; Sci. 2004, 278, 342.
Grimes, C. A. Nanotechnology 2006, 17, 1446. (261) Li, J.; Tang, Z.; Zhang, Z. Key Eng. Mater. 2005, 280-283, 651.
(216) Ruan, C. M.; Paulose, M.; Varghese, O. K.; Grimes, C. A. Sol. Energy (262) Meskin, P. E.; Ivanov, V. K.; Barantchikov, A. E.; Churagulov, B.
Mater. Sol. Cells 2006, 90, 1283. R.; Tretyakov, Y. D. Ultrason. Sonochem. 2006, 13, 47.
(217) Macak, J. M.; Tsuchiya, H.; Schmuki, P. Angew. Chem., Int. Ed. (263) Xia, H.; Wang, Q. Chem. Mater. 2002, 14, 2158.
2005, 44, 2100. (264) Yu, J. C.; Zhang, L.; Yu, J. Chem. Mater. 2002, 14, 4647.
(218) Macak, J. M.; Tsuchiya, H.; Taveira, L.; Aldabergerova, S.; Schmuki, (265) Yu, J. C.; Zhang, L.; Yu, J. New J. Chem. 2002, 26, 416.
P. Angew. Chem., Int. Ed. 2005, 44, 7463. (266) Yu, J. C.; Yu, J.; Zhang, L.; Ho, W. J. Photochem. Photobiol., A
(219) Macak, J. M.; Tsuchiya, H.; Ghicov, A.; Schmuki, P. Electrochem. 2002, 148, 263.
Commun. 2005, 7, 1133. (267) Yu, J. C.; Zhang, L.; Li, Q.; Kwong, K. W.; Xu, A. W.; Lin, J.
(220) Prida, V. M.; Hernandez-Velez, M.; Cervera, M.; Pirota, K.; Sanz, Langmuir 2003, 19, 7673.
R.; Navas, D.; Asenjo, A.; Aranda, P.; Ruiz-Hitzky, E.; Batallan, F.; (268) Yun, C. Y.; Chah, S.; Kang, S. K.; Yi, J. Korean J. Chem. Eng.
Vazquez, M.; Hernando, B.; Menendez, A.; Bordel, N.; Pereiro, R. 2004, 21, 1062.
J. Magn. Magn. Mater. 2005, 294, e69. (269) Zhu, Y.; Li, H.; Koltypin, Y.; Hacohen, Y. R.; Gedanken, A. Chem.
(221) Quan, X.; Yang, S.; Ruan, X.; Zhao, H. EnViron. Sci. Technol. 2005, Commun. 2001, 2616.
39, 3770. (270) Corradi, A. B.; Bondioli, F.; Focher, B.; Ferrari, A. M.; Grippo, C.;
(222) Tsuchiya, H.; Macak, J. M.; Taveira, L.; Balaur, E.; Ghicov, A.; Mariani, E.; Villa, C. J. Am. Ceram. Soc. 2005, 88, 2639.
Sirotna, K.; Schmuki, P. Electrochem. Commun. 2005, 7, 576. (271) Gressel-Michel, E.; Chaumont, D.; Stuerga, D. J. Colloid Interface
(223) Zhao, J.; Wang, X.; Chen, R.; Li, L. Solid State Commun. 2005, Sci. 2005, 285, 674.
134, 705. (272) Ma, G.; Zhao, X.; Zhu, J. Int. J. Mod. Phys. B 2005, 19, 2763.
(224) Cai, Q. Y.; Yang, L. X.; Yu, Y. Thin Solid Films 2006, 515, 1802. (273) Szabo, D. V.; Vollath, D.; Arnold, W. Ceram. Trans. 2001, 111,
(225) Lee, W. J.; Alhoshan, M.; Smyrl, W. H. J. Electrochem. Soc. 2006, 217.
153, B499.
(274) Uchida, S.; Tomiha, M.; Masaki, N.; Miyazawa, A.; Takizawa, H.
(226) Macak, J. M.; Aldabergerova, S.; Ghicov, A.; Schmuki, P. Phys. Sol. Energy Mater. Sol. Cells 2004, 81, 135.
Status Solidi A 2006, 203, R67.
(275) Wu, X.; Jiang, Q. Z.; Ma, Z. F.; Fu, M.; Shangguan, W. F. Solid
(227) Macak, J. M.; Tsuchiya, H.; Berger, S.; Bauer, S.; Fujimoto, S.;
State Commun. 2005, 136, 513.
Schmuki, P. Chem. Phys. Lett. 2006, 428, 421.
(228) Macak, J. M.; Schmuki, P. Electrochim. Acta 2006, 52, 1258. (276) Yamamoto, T.; Wada, Y.; Yin, H.; Sakata, T.; Mori, H.; Yanagida,
(229) Bauer, S.; Kleber, S.; Schmuki, P. Electrochem. Commun. 2006, 8, S. Chem. Lett. 2002, 964.
1321. (277) Antonelli, D. M.; Ying, J. Y. Angew. Chem., Int. Ed. 1995, 34, 2014.
(230) Seifried, S.; Winterer, M.; Hahn, H. Chem. Vap. Deposition 2000, (278) Bartl, M. H.; Puls, S. P.; Tang, J.; Lichtenegger, H. C.; Stucky, G.
6, 239. D. Angew. Chem., Int. Ed. 2004, 43, 3037.
(231) Ayllon, J. A.; Figueras, A.; Garelik, S.; Spirkova, L.; Durand, J.; (279) Bartl, M. H.; Boettcher, S. W.; Hu, E. L.; Stucky, G. D. J. Am. Chem.
Cot, L. J. Mater. Sci. Lett. 1999, 18, 1319. Soc. 2004, 126, 10826.
(232) Pradhan, S. K.; Reucroft, P. J.; Yang, F.; Dozier, A. J. Cryst. Growth (280) Bartl, M. H.; Boettcher, S. W.; Frindell, K. L.; Stucky, G. D. Acc.
2003, 256, 83. Chem. Res. 2005, 38, 263.
(233) Wu, J. J.; Yu, C. C. J. Phys. Chem. B 2004, 108, 3377. (281) Beyers, E.; Cool, P.; Vansant, E. F. J. Phys. Chem. B 2005, 109,
(234) Park, D. G.; Burlitch, J. M. Chem. Mater. 1992, 4, 500. 10081.
(235) Gurav, A.; Kodas, T.; Pluym, T.; Xiong, Y. Aerosol Sci. Technol. (282) Boissiere, C.; Grosso, D.; Amenitsch, H.; Gibaud, A.; Coupe, A.;
1993, 19, 411. Baccile, N.; Sanchez, C. Chem. Commun. 2003, 2798.
(236) Jang, H. D.; Kim, S. K. Mater. Res. Bull. 2001, 36, 627. (283) Cabrera, S.; El Haskouri, J.; Beltran-Porter, A.; Beltran-Porter, D.;
(237) Pratsinis, S. E.; Vemury, S.; Zhu, W. Polym. Mater. Sci. Eng. 1995, Marcos, M. D.; Amoros, P. Solid State Sci. 2000, 2, 513.
73, 31. (284) Crepaldi, E. L.; Soler-Illia, G. J. D. A.; Grosso, D.; Albouy, P. A.;
(238) Vemury, S.; Pratsinis, S. E. Appl. Phys. Lett. 1995, 66, 3275. Amenitsch, H.; Sanchez, C. Nanoporous Mater. III 2002, 141, 235.
(239) Vemury, S.; Pratsinis, S. E.; Kibbey, L. J. Mater. Res. 1997, 12, (285) Crepaldi, E. L.; Soler-Illia, G. J. D. A.; Grosso, D.; Cagnol, F.; Ribot,
1031. F.; Sanchez, C. J. Am. Chem. Soc. 2003, 125, 9770.
(240) Ishigaki, T.; Oh, S. M.; Park, D. W. Trans. Mater. Res. Soc. Jpn (286) Ding, H.; Sun, H.; Shan, Y. J. Photochem. Photobiol., A 2004, 169,
2004, 29, 3415. 101.
(241) Li, Y. L.; Ishigaki, T. Thin Solid Films 2002, 407, 79. (287) Frindell, K. L.; Bartl, M. H.; Popitsch, A.; Stucky, G. D. Angew.
(242) Li, Y. L.; Ishigaki, T. J. Phys. Chem. B 2004, 108, 15536. Chem., Int. Ed. 2002, 41, 960.
2952 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

(288) Grosso, D.; Soler-Illia, G. J. D. A.; Babonneau, F.; Sanchez, C.; (335) Holland, B. T.; Blanford, C. F.; Do, T.; Stein, A. Chem. Mater. 1999,
Albouy, P. A.; Brunet-Bruneau, A.; Balkenende, A. R. AdV. Mater. 11, 795.
2001, 13, 1085. (336) Huang, J. D.; Raikh, M.; Eratat, N.; Vardeny, Z. V.; Zakhidov, A.
(289) Grosso, D.; Soler-Illia, G. J. D. A.; Crepaldi, E. L.; Cagnol, F.; A.; Baughman, R. H. Synth. Met. 2001, 116, 505.
Sinturel, C.; Bourgeois, A.; Brunet-Bruneau, A.; Amenitsch, H.; (337) Ji, L.; Rong, J.; Yang, Z. Chem. Commun. 2003, 1080.
Albouy, P. A.; Sanchez, C. Chem. Mater. 2003, 15, 4562. (338) Johnson, N. P.; McComb, D. W.; Richel, A.; Treble, B. M.; De La
(290) Kavan, L.; Rathousky, J.; Grätzel, M.; Shklover, V.; Zukal, A. Rue, R. M. Synth. Met. 2001, 116, 469.
Microporous Mesoporous Mater. 2001, 44-45, 653. (339) Kavan, L.; Zukalova, M.; Kalbac, M.; Grätzel, M. J. Electrochem.
(291) Krtil, P.; Fattakhova, D.; Kavan, L.; Burnside, S.; Grätzel, M. Solid Soc. 2004, 151, A1301.
State Ionics 2000, 135, 101. (340) King, J. S.; Heineman, D.; Graugnard, E.; Summers, C. J. Appl. Surf.
(292) Liu, C.; Fu, L.; Economy, J. J. Mater. Chem. 2004, 14, 1187. Sci. 2005, 244, 511.
(293) Luo, H.; Wang, C.; Yan, Y. Chem. Mater. 2003, 15, 3841. (341) King, J. S.; Graugnard, E.; Summers, C. J. AdV. Mater. 2005, 17,
(294) Lyu, Y. Y.; Yi, S. H.; Shon, J. K.; Chang, S.; Pu, L. S.; Lee, S. Y.; 1010.
Yie, J. E.; Char, K.; Stucky, G. D.; Kim, J. M. J. Am. Chem. Soc. (342) Koenderink, A. F.; Bechger, L.; Lagendijk, A.; Vos, W. L. Phys.
2004, 126, 2310. Status Solidi A 2003, 197, 648.
(295) Ostomel, T. A.; Stucky, G. D. Chem. Commun. 2004, 1016. (343) Kuai, S. L.; Truong, V. V.; Hache, A.; Hu, X. F. J. Appl. Phys. 2004,
(296) Peng, T.; Zhao, D.; Dai, K.; Shi, W.; Hirao, K. J. Phys. Chem. B 96, 5982.
2005, 109, 4947. (344) Li, H. L.; Dong, W.; Bongard, H. J.; Marlow, F. J. Phys. Chem. B
(297) Perez, M. D.; Otal, E.; Bilmes, S. A.; Soler-Illia, G. J. A. A.; Crepaldi, 2005, 109, 9939.
E. L.; Grosso, D.; Sanchez, C. Langmuir 2004, 20, 6879. (345) Marlow, F.; Dong, W. ChemPhysChem. 2003, 4, 549.
(298) Putnam, R. L.; Nakagawa, N.; McGrath, K. M.; Yao, N.; Aksay, I. (346) Ni, P.; Cheng, B.; Zhang, D. Appl. Phys. Lett. 2002, 80, 1879.
A.; Gruner, S. M.; Navrotsky, A. Chem. Mater. 1997, 9, 2690. (347) Nishimura, S.; Abrams, N.; Lewis, B. A.; Halaoui, L. I.; Mallouk,
(299) Saito, Y.; Kambe, S.; Kitamura, T.; Wada, Y.; Yanagida, S. Sol. T. E.; Benkstein, K. D.; Van de Lagemaat, J.; Frank, A. J. J. Am.
Energy Mater. Sol. Cells 2004, 83, 1. Chem. Soc. 2003, 125, 6306.
(300) Smarsly, B.; Grosso, D.; Brezesinski, T.; Pinna, N.; Boissiere, C.; (348) Richel, A.; Johnson, N. P.; McComb, D. W. Appl. Phys. Lett. 2000,
Antonietti, M.; Sanchez, C. Chem. Mater. 2004, 16, 2948. 76, 1816.
(301) Soler-Illia, G. J. D. A.; Louis, A.; Sanchez, C. Chem. Mater. 2002, (349) Romanov, S. G.; Johnson, N. P.; Fokin, A. V.; Butko, V. Y.; Yates,
14, 750. H. M.; Pemble, M. E.; Torres, C. M. S. Appl. Phys. Lett. 1997, 70,
(302) Tang, J.; Wu, Y.; McFarland, E. W.; Stucky, G. D. Chem. Commun. 2091.
2004, 1670. (350) Schroden, R. C.; Al-Daous, M.; Stein, A. Chem. Mater. 2001, 13,
(303) Wang, X.; Yu, J. C.; Yip, H. Y.; Wu, L.; Wong, P. K.; Lai, S. Y. 2945.
Chem.sEur. J. 2005, 11, 2997. (351) Schroden, R. C.; Al-Daous, M.; Blanford, C. F.; Stein, A. Chem.
(304) Wang, Y.; Zhang, S.; Wu, X. Nanotechnology 2004, 15, 1162. Mater. 2002, 14, 3305.
(305) Yang, P.; Deng, T.; Zhao, D.; Feng, P.; Pine, D.; Chmelka, B. F.; (352) Somani, P. R.; Dionigi, C.; Murgia, M.; Palles, D.; Nozar, P.; Ruani,
Whitesides, G. M.; Stucky, G. D. Science 1998, 282, 2244. G. Sol. Energy Mater. Sol. Cells 2005, 87, 513.
(306) Yang, P.; Zhao, D.; Margolese, D. I.; Chmelka, B. F.; Stucky, G. D. (353) Subramania, G.; Biswas, R.; Constant, K.; Sigalas, M. M.; Ho, K.
Nature 1998, 396, 152. M. Phys. ReV. B 2001, 63, 235111/1.
(307) Yang, P.; Zhao, D.; Margolese, D. I.; Chmelka, B. F.; Stucky, G. D. (354) Wang, X. D.; Graugnard, E.; King, J. S.; Wang, Z. L.; Summers, C.
Chem. Mater. 1999, 11, 2813. J. Nano Lett. 2004, 4, 2223.
(308) Yi, D. K.; Kim, D. Y. Nano Lett. 2003, 3, 207. (355) Wang, X.; Neff, C.; Graugnard, E.; Ding, Y.; King, J. S.; Pranger,
(309) Yu, J. C.; Wang, X.; Wu, L.; Ho, W.; Zhang, L.; Zhou, G. AdV. L. A.; Tannenbaum, R.; Wang, Z. L.; Summers, C. J. AdV. Mater.
Funct. Mater. 2004, 14, 1178. 2005, 17, 2103.
(310) Yu, Y.; Yu, J. C.; Yu, J. G.; Kwok, Y. C.; Che, Y. K.; Zhao, J. C.; (356) Wijnhoven, J. E. G. J.; Vos, W. L. Science 1998, 281, 802.
Ding, L.; Ge, W. K.; Wong, P. K. Appl. Catal. A 2005, 289, 186. (357) Wijnhoven, J. E. G. J.; Bechger, L.; Vos, W. L. Chem. Mater. 2001,
(311) Zhang, Y.; Li, G.; Wu, Y.; Luo, Y.; Zhang, L. J. Phys. Chem. B 13, 4486.
2005, 109, 5478. (358) Yang, Z.; Niu, Z.; Lu, Y.; Hu, Z.; Han, C. C. Angew. Chem., Int.
(312) Zukalova, M.; Zukal, A.; Kavan, L.; Nazeeruddin, M. K.; Liska, P.; Ed. 2003, 42, 1943.
Grätzel, M. Nano Lett. 2005, 5, 1789. (359) Ma, R.; Sasaki, T.; Bando, Y. J. Am. Chem. Soc. 2004, 126, 10382.
(313) Frindell, K. L.; Bartl, M. H.; Robinson, M. R.; Bazan, G. C.; Popitsch, (360) Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. J. Am. Chem. Soc. 2004,
A.; Stucky, G. D. J. Solid State Chem. 2003, 172, 81. 126, 5851.
(314) Ayers, M. R.; Hunt, A. J. Mater. Lett. 1998, 34, 290. (361) Sasaki, T.; Watanabe, M.; Hashizume, H.; Yamada, H.; Nakazawa,
(315) Campbell, L. K.; Na, B. K.; Ko, E. I. Chem. Mater. 1992, 4, 1329. H. J. Am. Chem. Soc. 1996, 118, 8329.
(316) Dagan, G.; Tomkiewicz, M. J. Phys. Chem. 1993, 97, 12651. (362) Sasaki, T.; Watanabe, M. J. Phys. Chem. B 1997, 101, 10159.
(317) Dagan, G.; Tomkiewicz, M. J. Non-Cryst. Solids 1994, 175, 294. (363) Sasaki, T. Supramol. Sci. 1998, 5, 367.
(318) Kelly, S.; Pollak, F. H.; Tomkiewicz, M. J. Phys. Chem. B 1997, (364) Sasaki, T.; Watanabe, M. J. Am. Chem. Soc. 1998, 120, 4682.
101, 2730. (365) Sasaki, T.; Ebina, Y.; Watanabe, M.; Decher, G. Chem. Commun.
(319) Malinowska, B.; Walendziewski, J.; Robert, D.; Weber, J. V.; 2000, 2163.
Stolarski, M. Int. J. Photoenergy 2003, 5, 147. (366) Sasaki, T.; Ebina, Y.; Kitami, Y.; Watanabe, M.; Oikawa, T. J. Phys.
(320) Masson, O.; Rieux, V.; Guinebretiere, R.; Dauger, A. Nanostruct. Chem. B 2001, 105, 6116.
Mater. 1996, 7, 725. (367) Tanaka, T.; Ebina, Y.; Takada, K.; Kurashima, K.; Sasaki, T. Chem.
(321) Nishi, T.; Miao, L.; Tanemura, S.; Tanemura, M.; Toh, S.; Kaneko, Mater. 2003, 15, 3564.
K.; Tajiri, K. Trans. Mater. Res. Soc. Jpn 2005, 30, 553. (368) Wang, F.; Jiu, J.; Pei, L.; Nakagawa, K.; Isoda, S.; Adachi, M. Chem.
(322) Schneider, M.; Baiker, A. Catal. Today 1997, 35, 339. Lett. 2005, 34, 418.
(323) Suh, D. J.; Park, T. J. J. Mater. Sci. Lett. 1997, 16, 490. (369) Hamad, S.; Catlow, C. R. A.; Woodley, S. M.; Lago, S.; Mejias, J.
(324) Watson, J. M.; Cooper, A. T.; Flora, J. R. V. EnViron. Eng. Sci. A. J. Phys. Chem. B 2005, 109, 15741.
2005, 22, 666. (370) Swamy, V.; Kuznetsov, A.; Dubrovinsky, L. S.; Caruso, R. A.;
(325) Zhu, Z.; Lin, M.; Dagan, G.; Tomkiewicz, M. J. Phys. Chem. 1995, Shchukin, D. G.; Muddle, B. C. Phys. ReV. B 2005, 71, 184302/1.
99, 15950. (371) Barnard, A. S.; Zapol, P. Phys. ReV. B 2004, 70, 235403/1.
(326) Zhu, Z.; Tsung, L. Y.; Tomkiewicz, M. J. Phys. Chem. 1995, 99, (372) Barnard, A. S.; Zapol, P. J. Phys. Chem. B 2004, 108, 18435.
15945. (373) Barnard, A. S.; Zapol, P.; Curtiss, L. A. J. Chem. Theory Comput.
(327) Aronson, B. J.; Blanford, C. F.; Stein, A. Chem. Mater. 1997, 9, 2005, 1, 108.
2842. (374) Barnard, A. S.; Curtiss, L. A. Nano Lett. 2005, 5, 1261.
(328) Dong, W.; Bongard, H.; Tesche, B.; Marlow, F. AdV. Mater. 2002, (375) Barnard, A. S.; Zapol, P.; Curtiss, L. A. Surf. Sci. 2005, 582, 173.
14, 1457. (376) Enyashin, A. N.; Seifert, G. Phys. Status Solidi B 2005, 242, 1361.
(329) Dong, W.; Marlow, F. Phys. E 2003, 17, 431. (377) Chen, L. X.; Rajh, T.; Jager, W.; Nedeljkovic, J.; Thurnauer, M. C.
(330) Dong, W.; Bongard, H. J.; Marlow, F. Chem. Mater. 2003, 15, 568. J. Synchrotron Radiat. 1999, 6, 445.
(331) Gu, Z. Z.; Kubo, S.; Qian, W.; Einaga, Y.; Tryk, D. A.; Fujishima, (378) Qian, X.; Qin, D.; Song, Q.; Bai, Y.; Li, T.; Tang, X.; Wang, E.;
A.; Sato, O. Langmuir 2001, 17, 6751. Dong, S. Thin Solid Films 2001, 385, 152.
(332) Gu, Z. Z.; Fujishima, A.; Sato, O. Angew. Chem., Int. Ed. 2002, 41, (379) Hwu, Y.; Yao, Y. D.; Cheng, N. F.; Tung, C. Y.; Lin, H. M.
2067. Nanostruct. Mater. 1997, 9, 355.
(333) Halaoui, L. I.; Abrams, N. M.; Mallouk, T. E. J. Phys. Chem. B (380) Gribb, A. A.; Banfield, J. F. Am. Mineral. 1997, 82, 717.
2005, 109, 6334. (381) Ye, X.; Sha, J.; Jiao, Z.; Zhang, L. Nanostruct. Mater. 1998, 8, 919.
(334) Holland, B. T.; Blanford, C.; Stein, A. Science 1998, 281, 538. (382) Kominami, H.; Kohno, M.; Kera, Y. J. Mater. Chem. 2000, 10, 1151.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2953

(383) Ranade, M. R.; Navrotsky, A.; Zhang, H. Z.; Banfield, J. F.; Elder, (430) Bessekhouad, Y.; Robert, D.; Weber, J. V.; Chaoui, N. J. Photochem.
S. H.; Zaban, A.; Borse, P. H.; Kulkarni, S. K.; Doran, G. S.; Photobiol. A 2004, 167, 49.
Whitfield, H. J. Proc. Natl. Acad. Sci. 2002, 99, 6476. (431) Borgarello, E.; Kiwi, J.; Grätzel, M.; Pelizzetti, E.; Visca, M. J. Am.
(384) Li, W.; Ni, C.; Lin, H.; Huang, C. P.; Shah, S. I. J. Appl. Phys. Chem. Soc. 1982, 104, 2996.
2004, 96, 6663. (432) Bryan, J. D.; Heald, S. M.; Chambers, S. A.; Gamelin, D. R. J. Am.
(385) Lindgren, T.; Mwabora, J. M.; Avendano, E.; Jonsson, J.; Hoel, A.; Chem. Soc. 2004, 126, 11640.
Granqvist, C. G.; Lindquist, S. E. J. Phys. Chem. B 2003, 107, 5709. (433) Cao, Y.; Yang, W.; Zhang, W.; Liu, G.; Yue, P. New J. Chem. 2004,
(386) Barborini, E.; Kholmanov, I. N.; Piseri, P.; Ducati, C.; Bottani, C. 28, 218.
E.; Milani, P. Appl. Phys. Lett. 2002, 81, 3052. (434) Choi, W.; Termin, A.; Hoffmann, M. R. J. Phys. Chem. 1994, 98,
(387) Bersani, D.; Lottici, P. P.; Ding, X. Z. Appl. Phys. Lett. 1998, 72, 13669.
73. (435) Choi, W.; Termin, A.; Hoffmann, M. R. Angew. Chem. 1994, 106,
(388) Choi, H. C.; Jung, Y. M.; Kim, S. B. Vib. Spectrosc. 2005, 37, 33. 1148.
(389) Liang, L. H.; Shen, C. M.; Chen, X. P.; Liu, W. M.; Gao, H. J. J. (436) Choi, Y. J.; Banerjee, A.; Bandyopadhyay, A.; Bose, S. Ceram. Trans.
Phys.: Condens. Matter 2004, 16, 267. 2005, 159, 67.
(390) Ma, W.; Lu, Z.; Zhang, M. Appl. Phys. A 1998, A66, 621. (437) Coloma, F.; Marquez, F.; Rochester, C. H.; Anderson, J. A. Phys.
(391) Parker, J. C.; Siegel, R. W. J. Mater. Res. 1990, 5, 1246. Chem. Chem. Phys. 2000, 2, 5320.
(392) Parker, J. C.; Siegel, R. W. Appl. Phys. Lett. 1990, 57, 943. (438) Gracia, F.; Holgado, J. P.; Caballero, A.; Gonzalez-Elipe, A. R. J.
(393) Tanaka, A.; Onari, S.; Arai, T. Phys. ReV. B 1993, 47, 1237. Phys. Chem. B 2004, 108, 17466.
(394) Wang, Z.; Saxena, S. K. Solid State Commun. 2001, 118, 75. (439) Grätzel, M.; Howe, R. F. J. Phys. Chem. 1990, 94, 2566.
(395) Zhang, W. F.; He, Y. L.; Zhang, M. S.; Yin, Z.; Chen, Q. J. Phys. (440) Herrmann, J. M.; Disdier, J.; Pichat, P. Chem. Phys. Lett. 1984, 108,
D 2000, 33, 912. 618.
(396) Asahi, R.; Taga, Y.; Mannstadt, W.; Freeman, A. J. Phys. ReV. B (441) Karakitsou, K. E.; Verykios, X. E. J. Phys. Chem. 1993, 97, 1184.
2000, 61, 7459. (442) Li, F. B.; Li, X. Z.; Hou, M. F. Appl. Catal. B 2004, 48, 185.
(397) Sorantin, P. I.; Schwarz, K. Inorg. Chem. 1992, 31, 567. (443) Li, J.; Luo, S.; Yao, W.; Zhang, Z. Mater. Lett. 2003, 57, 3748.
(398) Wu, Z. Y.; Ouvrared, G.; Gressier, P.; Natoli, C. R. Phys. ReV. B (444) Li, W.; Wang, Y.; Lin, H.; Shah, S. I.; Huang, C. P.; Doren, D. J.;
1997, 55, 10382. Rykov, S. A.; Chen, J. G.; Barteau, M. A. Appl. Phys. Lett. 2003,
(399) Brydson, R.; Williams, B. G.; Engel, W.; Sauer, H.; Zeitler, E.; 83, 4143.
Thomas, J. M. Solid State Commun. 1987, 64, 609. (445) Luo, Z.; Gao, Q. J. Photochem. Photobiol., A 1992, 63, 367.
(400) Brydson, R.; Sauer, H.; Engel, W.; Thomas, J. M.; Zeitler, E.; Kosugi, (446) Martin, S. T.; Morrison, C. L.; Hoffmann, M. R. J. Phys. Chem.
N.; Kuroda, H. J. Phys.: Condens. Matter 1989, 1, 797. 1994, 98, 13695.
(401) Choi, H. C.; Ahn, H. J.; Jung, Y. M.; Lee, M. K.; Shin, H. J.; Kim, (447) Mu, W.; Herrmann, J. M.; Pichat, P. Catal. Lett. 1989, 3, 73.
S. B.; Sung, Y. E. Appl. Spectrosc. 2004, 58, 598. (448) Nagaveni, K.; Hegde, M. S.; Madras, G. J. Phys. Chem. B 2004,
(402) Finkelstein, L. D.; Kurmaev, E. Z.; Korotin, M. A.; Moewes, A.; 108, 20204.
Schneider, B.; Butorin, S. M.; Guo, J. H.; Nordgren, J.; Hartmann, (449) Palmisano, L.; Augugliaro, V.; Sclafani, A.; Schiavello, M. J. Phys.
D.; Neumann, M.; Ederer, D. L. Phys. ReV. B 1999, 60, 2212. Chem. 1988, 92, 6710.
(403) Luca, V.; Djajanti, S.; Howe, R. F. J. Phys. Chem. B 1998, 102, (450) Peill, N. J.; Bourne, L.; Hoffmann, M. R. J. Photochem. Photobiol.,
10650. A 1997, 108, 221.
(404) Sanjines, R.; Tang, H.; Berger, H.; Gozzo, F.; Margaritondo, G.; Levy, (451) Peng, S.; Li, Y.; Jiang, F.; Lu, G.; Li, S. Chem. Phys. Lett. 2004,
F. J. Appl. Phys. 1994, 75, 2945. 398, 235.
(405) Zimmermann, R.; Steiner, P.; Claessen, R.; Reinert, F.; Hufner, S. (452) Sakatani, Y.; Ando, H.; Okusako, K.; Koike, H.; Nunoshige, J.;
J. Electron. Spectrosc. Relat. Phenom. 1998, 96, 179. Takata, T.; Kondo, J. N.; Hara, M.; Domen, K. J. Mater. Res. 2004,
(406) Anpo, M.; Shima, T.; Kodama, S.; Kubokawa, Y. J. Phys. Chem. 19, 2100.
1987, 91, 4305. (453) Salmi, M.; Tkachenko, N.; Lamminmaeki, R. J.; Karvinen, S.;
(407) Kavan, L.; Stoto, T.; Grätzel, M.; Fitzmaurice, D.; Shklover, V. J. Vehmanen, V.; Lemmetyinen, H. J. Photochem. Photobiol., A 2005,
Phys. Chem. 1993, 97, 9493. 175, 8.
(408) Henglein, A. Chem. ReV. 1989, 89, 1861. (454) Soria, J.; Conesa, J. C.; Augugliaro, V.; Palmisano, L.; Schiavello,
(409) Serpone, N.; Lawless, D.; Khairutdinov, R.; Pelizzetti, E. J. Phys. M.; Sclafani, A. J. Phys. Chem. 1991, 95, 274.
Chem. 1995, 99, 16655. (455) Szabo, A.; Urda, A. Prog. Catal. 2002, 11, 73.
(410) Serpone, N.; Lawless, D.; Khairutdinov, R. J. Phys. Chem. 1995, (456) Szabo, A.; Urda, A. Prog. Catal. 2003, 12, 51.
99, 16646. (457) Wang, C. Y.; Bahnemann, D. W.; Dohrmann, J. K. Chem. Commun.
(411) Monticone, S.; Tufeu, R.; Kanaev, A. V.; Scolan, E.; Sanchez, C. 2000, 1539.
Appl. Surf. Sci. 2000, 162-163, 565. (458) Wang, C. Y.; Boettcher, C.; Bahnemann, D. W.; Dohrmann, J. K. J.
(412) Braginsky, L.; Shklover, V. Eur. Phys. J. D 1999, 9, 627. Nanopart. Res. 2004, 6, 119.
(413) Sato, H.; Ono, K.; Sasaki, T.; Yamagishi, A. J. Phys. Chem. B 2003, (459) Wang, W. Y.; Zhang, D. F.; Chen, X. L. J. Mater. Sci. 2003, 38,
107, 9824. 2049.
(414) Mora-Sero, I.; Bisquert, J. Nano Lett. 2003, 3, 945. (460) Wang, Y.; Cheng, H.; Hao, Y.; Ma, J.; Li, W.; Cai, S. J. Mater. Sci.
(415) Szczepankiewicz, S. H.; Colussi, A. J.; Hoffmann, M. R. J. Phys. 1999, 34, 3721.
Chem. B 2000, 104, 9842. (461) Wang, Y.; Hao, Y.; Cheng, H.; Ma, H.; Xu, B.; Li, W.; Cai, S. J.
(416) Howe, R. F.; Grätzel, M. J. Phys. Chem. 1985, 89, 4495. Mater. Sci. 1999, 34, 2773.
(417) Howe, R. F.; Grätzel, M. J. Phys. Chem. 1987, 91, 3906. (462) Wang, Y.; Cheng, H.; Hao, Y.; Ma, J.; Xu, B.; Li, W.; Cai, S. J.
(418) Hurum, D. C.; Agrios, A. G.; Crist, S. E.; Gray, K. A.; Rajh, T.; Mater. Sci. Lett. 1999, 18, 127.
Thurnauer, M. C. J. Electron. Spectrosc. Relat. Phenom. 2006, 150, (463) Wang, Y.; Cheng, H.; Hao, Y.; Ma, J.; Li, W.; Cai, S. Thin Solid
155. Films 1999, 349, 120.
(419) Hurum, D. C.; Gray, K. A.; Rajh, T.; Thurnauer, M. C. J. Phys. (464) Xu, X. H.; Wang, M.; Hou, Y.; Yao, W. F.; Wang, D.; Wang, H. J.
Chem. B 2005, 109, 977. Mater. Sci. Lett. 2002, 21, 1655.
(420) Colombo, D. P., Jr.; Bowman, R. M. J. Phys. Chem. 1995, 99, 11752. (465) Yamashita, H.; Ichihashi, Y.; Takeuchi, M.; Kishiguchi, S.; Anpo,
(421) Colombo, D. P., Jr.; Bowman, R. M. J. Phys. Chem. 1996, 100, M. J. Synchrotron Radiat. 1999, 6, 451.
18445. (466) Anpo, M.; Kishiguchi, S.; Ichihashi, Y.; Takeuchi, M.; Yamashita,
(422) Skinner, D. E.; Colombo, D. P., Jr.; Cavaleri, J. J.; Bowman, R. M. H.; Ikeue, K.; Morin, B.; Davidson, A.; Che, M. Res. Chem. Intermed.
J. Phys. Chem. 1995, 99, 7853. 2001, 27, 459.
(423) Bahnemann, D. W.; Hilgendorff, M.; Memming, R. J. Phys. Chem. (467) Anpo, M. Pure Appl. Chem. 2000, 72, 1787.
B 1997, 101, 4265. (468) Anpo, M. Pure Appl. Chem. 2000, 72, 1265.
(424) Szczepankiewicz, S. H.; Moss, J. A.; Hoffmann, M. R. J. Phys. Chem. (469) Anpo, M.; Takeuchi, M. Int. J. Photoenergy 2001, 3, 89.
B 2002, 106, 7654. (470) Anpo, M.; Takeuchi, M. J. Catal. 2003, 216, 505.
(425) Berger, T.; Sterrer, M.; Diwald, O.; Knoezinger, E.; Panayotov, D.; (471) Takeuchi, M.; Yamashita, H.; Matsuoka, M.; Anpo, M.; Hirao, T.;
Thompson, T. L.; Yates, J. T., Jr. J. Phys. Chem. B 2005, 109, 6061. Itoh, N.; Iwamoto, N. Catal. Lett. 2000, 67, 135.
(426) Burda, C.; Lou, Y.; Chen, X.; Samia, A. C. S.; Stout, J.; Gole, J. L. (472) Choi, Y.; Umebayashi, T.; Yoshikawa, M. J. Mater. Sci. 2004, 39,
Nano Lett. 2003, 3, 1049. 1837.
(427) Chen, X.; Burda, C. J. Phys. Chem. B 2004, 108, 15446. (473) Irie, H.; Watanabe, Y.; Hashimoto, K. Chem. Lett. 2003, 32, 772.
(428) Chen, X.; Lou, Y.; Samia, A. C. S.; Burda, C.; Gole, J. L. AdV. Funct. (474) Shen, M.; Wu, Z.; Huang, H.; Du, Y.; Zou, Z.; Yang, P. Mater. Lett.
Mater. 2005, 15, 41. 2006, 60, 693.
(429) Chen, X.; Lou, Y.; Samia, A. C.; Burda, C. Nano Lett. 2003, 3, 799. (475) Park, J. H.; Kim, S.; Bard, A. J. Nano Lett. 2006, 6, 24.
2954 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

(476) Khan, S. U. M.; Al-Shahry, M.; Ingler, W. B., Jr. Science 2002, 297, (521) Lindgren, T.; Lu, J.; Hoel, A.; Granqvist, C. G.; Torres, G. R.;
2243. Lindquist, S. E. Sol. Energy Mater. Sol. Cells 2004, 84, 145.
(477) Chen, X.; Lou, Y.; Burda, C. Int. J. Nanotechnol. 2004, 1, 105. (522) Nakano, Y.; Morikawa, T.; Ohwaki, T.; Taga, Y. Appl. Phys. Lett.
(478) Gole, J. L.; Stout, J. D.; Burda, C.; Lou, Y.; Chen, X. J. Phys. Chem. 2005, 87, 052111/1.
B 2004, 108, 1230. (523) Nakano, Y.; Morikawa, T.; Ohwaki, T.; Taga, Y. Appl. Phys. Lett.
(479) Liu, Y.; Chen, X.; Li, J.; Burda, C. Chemosphere 2005, 61, 11. 2005, 86, 132104/1.
(480) Prokes, S. M.; Gole, J. L.; Chen, X.; Burda, C.; Carlos, W. E. AdV. (524) Okato, T.; Sakano, T.; Obara, M. Phys. ReV. B 2005, 72, 115124/1.
Funct. Mater. 2005, 15, 161. (525) Sathish, M.; Viswanathan, B.; Viswanath, R. P.; Gopinath, C. S.
(481) Sakthivel, S.; Kisch, H. ChemPhysChem 2003, 4, 487. Chem. Mater. 2005, 17, 6349.
(482) Sakthivel, S.; Janczarek, M.; Kisch, H. J. Phys. Chem. B 2004, 108, (526) Sato, S.; Nakamura, R.; Abe, S. Appl. Catal. A 2005, 284, 131.
19384. (527) Tesfamichael, T.; Will, G.; Bell, J. Appl. Surf. Sci. 2005, 245, 172.
(483) Irie, H.; Watanabe, Y.; Hashimoto, K. J. Phys. Chem. B 2003, 107, (528) Tokudome, H.; Miyauchi, M. Chem. Lett. 2004, 33, 1108.
5483. (529) Torres, G. R.; Lindgren, T.; Lu, J.; Granqvist, C. G.; Lindquist, S.
(484) Chen, S.; Chen, L.; Gao, S.; Cao, G. Chem. Phys. Lett. 2005, 413, E. J. Phys. Chem. B 2004, 108, 5995.
404. (530) Yin, S.; Aita, Y.; Komatsu, M.; Wang, J.; Tang, Q.; Sato, T. J. Mater.
(485) Diwald, O.; Thompson, T. L.; Zubkov, T.; Goralski, E.; Walck, S. Chem. 2005, 15, 674.
D.; Yates, J. T., Jr. J. Phys. Chem. B 2004, 108, 6004. (531) Yin, S.; Yamaki, H.; Komatsu, M.; Zhang, Q.; Wang, J.; Tang, Q.;
(486) Nakamura, R.; Tanaka, T.; Nakato, Y. J. Phys. Chem. B 2004, 108, Saito, F.; Sato, T. Solid State Sci. 2005, 7, 1479.
10617. (532) Yin, S.; Ihara, K.; Komatsu, M.; Zhang, Q.; Saito, F.; Kyotani, T.;
(487) Huang, X. H.; Tang, Y. C.; Hu, C.; Yu, H. Q.; Chen, C. S. J. EnViron. Sato, T. Solid State Commun. 2006, 137, 132.
Sci. (Beijing, China) 2005, 17, 562. (533) Zhang, Q.; Wang, J.; Yin, S.; Sato, T.; Saito, F. J. Am. Ceram. Soc.
(488) Hong, Y. C.; Bang, C. U.; Shin, D. H.; Uhm, H. S. Chem. Phys. 2004, 87, 1161.
Lett. 2005, 413, 454. (534) Ho, W. K.; Yu, J. C.; Lee, S. C. J. Solid State Chem. 2006, 179,
(489) Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Science 1171.
2001, 293, 269. (535) Jang, J. S.; Kim, H. G.; Ji, S. M.; Bae, S. W.; Jung, J. H.; Shon, B.
(490) Chang, J. T.; Lai, Y. F.; He, J. L. Surf. Coat. Technol. 2005, 200, H.; Lee, J. S. J. Solid State Chem. 2006, 179, 1067.
1640. (536) Ghicov, A.; Macak, J. M.; Tsuchiya, H.; Kunze, J.; Haeublein, V.;
(491) Diwald, O.; Thompson, T. L.; Goralski, E. G.; Walck, S. D.; Yates, Frey, L.; Schmuki, P. Nano Lett. 2006, 6, 1080.
J. T., Jr. J. Phys. Chem. B 2004, 108, 52. (537) Xiong, Y. Y.; Ma, T.; Kong, L. H.; Chen, J. F.; Wu, X. Q.; Yu, H.
(492) Ohno, T. Water Sci. Technol. 2004, 49, 159. H.; Zhang, Z. X. J. Mater. Sci. Technol. 2006, 22, 353.
(493) Ohno, T.; Mitsui, T.; Matsumura, M. Chem. Lett. 2003, 32, 364. (538) Buzby, S.; Barakat, M. A.; Lin, H.; Ni, C.; Rykov, S. A.; Chen, J.
(494) Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; G.; Shah, S. I. J. Vac. Sci. Technol., B 2006, 24, 1210.
Matsumura, M. Appl. Catal. A 2004, 265, 115. (539) Yuan, J.; Chen, M. X.; Shi, J. W.; Shangguan, W. F. Int. J. Hydrogen
(495) Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai, K. Appl. Phys. Lett. Energy 2006, 31, 1326.
2002, 81, 454. (540) Venkatachalam, N.; Vinu, A.; Anandan, S.; Arabindoo, B.; Murug-
(496) Umebayashi, T.; Yamaki, T.; Tanaka, S.; Asai, K. Chem. Lett. 2003, esan, V. J. Nanosci. Nanotechnol. 2006, 6, 2499.
32, 330. (541) Yin, S.; Aita, Y.; Komatsu, M.; Sato, T. J. Eur. Ceram. Soc. 2006,
(497) Umebayashi, T.; Yamaki, T.; Yamamoto, S.; Miyashita, A.; Tanaka, 26, 2735.
S.; Sumita, T.; Asai, K. J. Appl. Phys. 2003, 93, 5156. (542) Wong, M. S.; Chu, W. C.; Sun, D. S.; Huang, H. S.; Chen, J. H.;
(498) Umebayashi, T.; Yamaki, T.; Sumita, T.; Yamamoto, S.; Miyashita, Tsai, P. J.; Lin, N. T.; Yu, M. S.; Hsu, S. F.; Wang, S. L.; Chang, H.
A.; Tanaka, S.; Asai, K. JAERI-ReV. 2003, 222. H. Appl. EnViron. Microbiol. 2006, 72, 6111.
(499) Umebayashi, T.; Yamaki, T.; Yamamoto, S.; Tanaka, S.; Asai, K. (543) Cong, Y.; Chen, F.; Zhang, J. L.; Anpo, M. Chem. Lett. 2006, 35,
Trans. Mater. Res. Soc. Jpn. 2003, 28, 461. 800.
(500) Hattori, A.; Yamamoto, M.; Tada, H.; Ito, S. Chem. Lett. 1998, 707. (544) In, S.; Orlov, A.; Garcia, F.; Tikhov, M.; Wright, D. S.; Lambert, R.
(501) Yu, J. G.; Yu, J. C.; Cheng, B.; Hark, S. K.; Iu, K. J. Solid State M. Chem. Commun. 2006, 4236.
Chem. 2003, 174, 372. (545) Macak, J. M.; Ghicov, A.; Hahn, R.; Tsuchiya, H.; Schmuki, P. J.
(502) Yu, J. C.; Yu, J. G.; Ho, W. K.; Jiang, Z. T.; Zhang, L. Z. Chem. Mater. Res. 2006, 21, 2824.
Mater. 2002, 14, 3808. (546) Serpone, N. J. Phys. Chem. B 2006, 110, 24287.
(503) Subbarao, S. N.; Yun, Y. H.; Kershaw, R.; Dwight, K.; Wold, A. (547) Livraghi, S.; Paganini, M. C.; Giamello, E.; Selloni, A.; Di Valentin,
Mater. Res. Bull. 1978, 13, 1461. C.; Pacchioni, G. J. Am. Chem. Soc. 2006, 128, 15666.
(504) Subbarao, S. N.; Yun, Y. H.; Kershaw, R.; Dwight, K.; Wold, A. (548) Klosek, S.; Raftery, D. J. Phys. Chem. B 2001, 105, 2815.
Inorg. Chem. 1979, 18, 488. (549) Matsumoto, Y.; Kurimoto, J.; Shimizu, T.; Sato, E. J. Electrochem.
(505) Li, D.; Haneda, H.; Labhsetwar, N. K.; Hishita, S.; Ohashi, N. Chem. Soc. 1981, 128, 1040.
Phys. Lett. 2005, 401, 579. (550) Meyer, G. J. Inorg. Chem. 2005, 44, 6852.
(506) Li, D.; Haneda, H.; Hishita, S.; Ohashi, N.; Labhsetwar, N. K. J. (551) Fitzmaurice, D.; Frei, H.; Rabani, J. J. Phys. Chem. 1995, 99, 9176.
Fluorine Chem. 2005, 126, 69. (552) Fujii, H.; Inata, K.; Ohtaki, M.; Eguchi, K.; Arai, H. J. Mater. Sci.
(507) Yamaki, T.; Umebayashi, T.; Sumita, T.; Yamamoto, S.; Maekawa, 2001, 36, 527.
M.; Kawasuso, A.; Itoh, H. Nucl. Instrum. Methods Phys. Res., Sect. (553) Hoyer, P.; Koenenkamp, R. Appl. Phys. Lett. 1995, 66, 349.
B 2003, 206, 254. (554) Matsumoto, H.; Matsunaga, T.; Sakata, T.; Mori, H.; Yoneyama, H.
(508) Luo, H.; Takata, T.; Lee, Y.; Zhao, J.; Domen, K.; Yan, Y. Chem. Langmuir 1995, 11, 4283.
Mater. 2004, 16, 846. (555) Qian, X.; Qin, D.; Bai, Y.; Li, T.; Tang, X.; Wang, E.; Dong, S. J.
(509) Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai, K. J. Phys. Chem. Solids Solid State Electrochem. 2001, 5, 562.
2002, 63, 1909. (556) Shen, Q.; Arae, D.; Toyoda, T. J. Photochem. Photobiol. A 2004,
(510) Wang, Y.; Doren, D. J. Solid State Commun. 2005, 136, 186. 164, 75.
(511) Wang, Y.; Doren, D. J. Solid State Commun. 2005, 136, 142. (557) Spanhel, L.; Weller, H.; Henglein, A. J. Am. Chem. Soc. 1987, 109,
(512) Di Valentin, C.; Pacchioni, G.; Selloni, A. Phys. ReV. B 2004, 70, 6632.
085116/1. (558) Vogel, R.; Hoyer, P.; Weller, H. J. Phys. Chem. 1994, 98, 3183.
(513) Di, Valentin, C.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Giamello, (559) Zaban, A.; Micic, O. I.; Gregg, B. A.; Nozik, A. J. Langmuir 1998,
E. J. Phys. Chem. B 2005, 109, 11414. 14, 3153.
(514) Di Valentin, C.; Pacchioni, G.; Selloni, A. Chem. Mater. 2005, 17, (560) Sant, P. A.; Kamat, P. V. Phys. Chem. Chem. Phys. 2002, 4, 198.
6656. (561) Robel, I.; Subramanian, V.; Kuno, M.; Kamat, P. V. J. Am. Chem.
(515) Enache, C. S.; Schoonman, J.; Van de Krol, R. J. Electroceram. 2004, Soc. 2006, 128, 2385.
13, 177. (562) Ohko, Y.; Tatsuma, T.; Fujii, T.; Naoi, K.; Niwa, C.; Kubota, Y.;
(516) Kuroda, Y.; Mori, T.; Yagi, K.; Makihata, N.; Kawahara, Y.; Nagao, Fujishima, A. Nat. Mater. 2003, 2, 29.
M.; Kittaka, S. Langmuir 2005, 21, 8026. (563) Naoi, K.; Ohko, Y.; Tatsuma, T. J. Am. Chem. Soc. 2004, 126, 3664.
(517) Lee, J. Y.; Park, J.; Cho, J. H. Appl. Phys. Lett. 2005, 87, 011904/1. (564) Naoi, K.; Ohko, Y.; Tatsuma, T. Chem. Commun. 2005, 1288.
(518) Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Chem. Mater. 2005, 17, (565) Kawahara, K.; Suzuki, K.; Ohko, Y.; Tatsuma, T. Phys. Chem. Chem.
2596. Phys. 2005, 7, 3851.
(519) Li, D.; Ohashi, N.; Hishita, S.; Kolodiazhnyi, T.; Haneda, H. J. Solid (566) Tian, Y.; Tatsuma, T. Chem. Commun. 2004, 1810.
State Chem. 2005, 178, 3293. (567) Tian, Y.; Tatsuma, T. J. Am. Chem. Soc. 2005, 127, 7632.
(520) Lin, Z.; Orlov, A.; Lambert, R. M.; Payne, M. C. J. Phys. Chem. B (568) Adachi, M.; Murata, Y.; Takao, J.; Jiu, J.; Sakamoto, M.; Wang, F.
2005, 109, 20948. J. Am. Chem. Soc. 2004, 126, 14943.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2955

(569) Adachi, M.; Murata, Y.; Okada, I.; Yoshikawa, S. J. Electrochem. (610) Kalyanasundaram, K.; Grätzel, M. Coord. Chem. ReV. 1998, 177,
Soc. 2003, 150, G488. 347.
(570) Argazzi, R.; Bignozzi, C. A.; Heimer, T. A.; Castellano, F. N.; Meyer, (611) Kamat, P. V. Electrochem. Nanomater. 2001, 229.
G. J. Inorg. Chem. 1994, 33, 5741. (612) Klein, C.; Nazeeruddin, M.; Di Censo, D.; Liska, P.; Grätzel, M.
(571) Argazzi, R.; Bignozzi, C. A.; Heimer, T. A.; Meyer, G. J. Inorg. Inorg. Chem. 2004, 43, 4216.
Chem. 1997, 36, 2. (613) Klein, C.; Nazeeruddin, M.; Liska, P.; Di Censo, D.; Hirata, N.;
(572) Argazzi, R.; Bignozzi, C. A.; Heimer, T. A.; Castellano, F. N.; Meyer, Palomares, E.; Durrant, J. R.; Grätzel, M. Inorg. Chem. 2005, 44,
G. J. J. Phys. Chem. B 1997, 101, 2591. 178.
(573) Argazzi, R.; Bignozzi, C. A.; Hasselmann, G. M.; Meyer, G. J. Inorg. (614) Kleverlaan, C.; Alebbi, M.; Argazzi, R.; Bignozzi, C. A.; Hasselmann,
Chem. 1998, 37, 4533. G. M.; Meyer, G. J. Inorg. Chem. 2000, 39, 1342.
(574) Argazzi, R.; Bignozzi, C. A.; Yang, M.; Hasselmann, G. M.; Meyer, (615) Kohle, O.; Ruile, S.; Grätzel, M. Inorg. Chem. 1996, 35, 4779.
G. J. Nano Lett. 2002, 2, 625. (616) Kovtyukhova, N. I.; Martin, B. R.; Mbindyo, J. K. N.; Smith, P. A.;
(575) Athanassov, Y.; Rotzinger, F. P.; Pechy, P.; Grätzel, M. J. Phys. Razavi, B.; Mayer, T. S.; Mallouk, T. E. J. Phys. Chem. B 2001,
Chem. B 1997, 101, 2558. 105, 8762.
(576) Bauer, C.; Boschloo, G.; Mukhtar, E.; Hagfeldt, A. J. Phys. Chem. (617) Kreller, D. I.; Kamat, P. V. J. Phys. Chem. 1991, 95, 4406.
B 2001, 105, 5585. (618) Kuciauskas, D.; Freund, M. S.; Gray, H. B.; Winkler, J. R.; Lewis,
(577) Bauer, C.; Boschloo, G.; Mukhtar, E.; Hagfeldt, A. J. Phys. Chem. N. S. J. Phys. Chem. B 2001, 105, 392.
B 2002, 106, 12693. (619) Kuciauskas, D.; Monat, J. E.; Villahermosa, R.; Gray, H. B.; Lewis,
(578) Benkoe, G.; Skrman, B.; Wallenberg, R.; Hagfeldt, A.; Sundstroem, N. S.; McCusker, J. K. J. Phys. Chem. B 2002, 106, 9347.
V.; Yartsev, A. P. J. Phys. Chem. B 2003, 107, 1370. (620) Liska, P.; Vlachopoulos, N.; Nazeeruddin, M. K.; Comte, P.; Grätzel,
(579) Burfeindt, B.; Hannappel, T.; Storck, W.; Willig, F. J. Phys. Chem. M. J. Am. Chem. Soc. 1988, 110, 3686.
1996, 100, 16463. (621) Liu, F.; Meyer, G. J. Inorg. Chem. 2003, 42, 7351.
(580) Dabestani, R.; Bard, A. J.; Campion, A.; Fox, M. A.; Mallouk, T. (622) Liu, F.; Meyer, G. J. Inorg. Chem. 2005, 44, 9305.
E.; Webber, S. E.; White, J. M. J. Phys. Chem. 1988, 92, 1872. (623) Martini, I.; Hodak, J.; Hartland, G. V.; Kamat, P. V. J. Chem. Phys.
(581) Dai, Q.; Rabani, J. J. Photochem. Photobiol. A 2002, 148, 17. 1997, 107, 8064.
(582) Desilvestro, J.; Grätzel, M.; Kavan, L.; Moser, J.; Augustynski, J. J. (624) Moss, J. A.; Yang, J. C.; Stipkala, J. M.; Wen, X.; Bignozzi, C. A.;
Am. Chem. Soc. 1985, 107, 2988. Meyer, G. J.; Meyer, T. J. Inorg. Chem. 2004, 43, 1784.
(583) Ferrere, S.; Gregg, B. A. J. Am. Chem. Soc. 1998, 120, 843. (625) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.;
(584) Ferrere, S. Chem. Mater. 2000, 12, 1083. Mueller, E.; Liska, P.; Vlachopoulos, N.; Grätzel, M. J. Am. Chem.
(585) Fessenden, R. W.; Kamat, P. V. J. Phys. Chem. 1995, 99, 12902. Soc. 1993, 115, 6382.
(586) Garcia, C. G.; Murakami Iha, N. Y.; Argazzi, R.; Bignozzi, C. A. J. (626) Nazeeruddin, M.; Pechy, P.; Grätzel, M. Chem. Commun. 1997, 1705.
Photochem. Photobiol., A 1998, 115, 239. (627) Nazeeruddin, M.; Grätzel, M. Comp. Coord. Chem. II 2004, 9, 719.
(587) Garcia, C. G.; Iha, N. Y. M. Int. J. Photoenergy 2001, 3, 131. (628) Nazeeruddin, M.; Humphry-Baker, R.; Officer, D. L.; Campbell, W.
(588) Garcia, C. G.; Nakano, A. K.; Kleverlaan, C. J.; Murakami Iha, N. M.; Burrell, A. K.; Grätzel, M. Langmuir 2004, 20, 6514.
Y. J. Photochem. Photobiol., A 2002, 151, 165. (629) Nazeeruddin, M.; Klein, C.; Liska, P.; Grätzel, M. Coord. Chem.
(589) Garcia, C. G.; Kleverlaan, C. J.; Bignozzi, C. A.; Murakami Iha, N. ReV. 2005, 249, 1460.
Y. J. Photochem. Photobiol., A 2002, 147, 143. (630) Nazeeruddin, M. K.; Muller, E.; Humphry-Baker, R.; Vlachopoulos,
(590) Geary, E. A. M.; Yellowlees, L. J.; Jack, L. A.; Oswald, I. D. H.; N.; Grätzel, M. Dalton Trans.: Inorg. Chem. 1997, 4571.
Parsons, S.; Hirata, N.; Durrant, J. R.; Robertson, N. Inorg. Chem. (631) Nazeeruddin, M. K.; Pechy, P.; Renouard, T.; Zakeeruddin, S. M.;
2005, 44, 242. Humphry-Baker, R.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.;
(591) Gillaizeau-Gauthier, I.; Odobel, F.; Alebbi, M.; Argazzi, R.; Costa, Shklover, V.; Spiccia, L.; Deacon, G. B.; Bignozzi, C. A.; Grätzel,
E.; Bignozzi, C. A.; Qu, P.; Meyer, G. J. Inorg. Chem. 2001, 40, M. J. Am. Chem. Soc. 2001, 123, 1613.
6073. (632) Nazeeruddin, M. K.; De Angelis, F.; Fantacci, S.; Selloni, A.;
(592) Gregg, B. A.; Pichot, F.; Ferrere, S.; Fields, C. L. J. Phys. Chem. B Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Grätzel, M. J. Am. Chem.
2001, 105, 1422. Soc. 2005, 127, 16835.
(593) Gregg, B. A. J. Phys. Chem. B 2003, 107, 4688. (633) Pechy, P.; Rotzinger, F. P.; Nazeeruddin, M. K.; Kohle, O.;
(594) Gregg, B. A.; Chen, S. G.; Ferrere, S. J. Phys. Chem. B 2003, 107, Zakeeruddin, S. M.; Humphry-Baker, R.; Grätzel, M. Chem. Com-
3019. mun. 1995, 65.
(595) Hannappel, T.; Burfeindt, B.; Storck, W.; Willig, F. J. Phys. Chem. (634) Pichot, F.; Ferrere, S.; Pitts, R. J.; Gregg, B. A. J. Electrochem. Soc.
B 1997, 101, 6799. 1999, 146, 4324.
(596) Haque, S. A.; Palomares, E.; Upadhyaya, H. M.; Otley, L.; Potter, (635) Pichot, F.; Gregg, B. A. J. Phys. Chem. B 2000, 104, 6.
R. J.; Holmes, A. B.; Durrant, J. R. Chem. Commun. 2003, 3008. (636) Ruile, S.; Kohle, O.; Pechy, P.; Grätzel, M. Inorg. Chim. Acta 1997,
(597) Haque, S. A.; Handa, S.; Peter, K.; Palomares, E.; Thelakkat, M.; 261, 129.
Durrant, J. R. Angew. Chem., Int. Ed. 2005, 44, 5740. (637) Saupe, G. B.; Mallouk, T. E.; Kim, W.; Schmehl, R. H. J. Phys.
(598) Haque, S. A.; Palomares, E.; Cho, B. M.; Green, A. N. M.; Hirata, Chem. B 1997, 101, 2508.
N.; Klug, D. R.; Durrant, J. R. J. Am. Chem. Soc. 2005, 127, 3456. (638) Sauve, G.; Cass, M. E.; Doig, S. J.; Lauermann, I.; Pomykal, K.;
(599) Hara, K.; Sugihara, H.; Tachibana, Y.; Islam, A.; Yanagida, M.; Lewis, N. S. J. Phys. Chem. B 2000, 104, 3488.
Sayama, K.; Arakawa, H.; Fujihashi, G.; Horiguchi, T.; Kinoshita, (639) Schlichthoerl, G.; Huang, S. Y.; Sprague, J.; Frank, A. J. J. Phys.
T. Langmuir 2001, 17, 5992. Chem. B 1997, 101, 8141.
(600) Hara, K.; Tachibana, Y.; Ohga, Y.; Shinpo, A.; Suga, S.; Sayama, (640) Schlichthoerl, G.; Park, N. G.; Frank, A. J. J. Phys. Chem. B 1999,
K.; Sugihara, H.; Arakawa, H. Sol. Energy Mater. Sol. Cells 2003, 103, 782.
77, 89. (641) Schmidt-Mende, L.; Kroeze, J. E.; Durrant, J. R.; Nazeeruddin, M.;
(601) Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Ohga, Y.; Shinpo, A.; Grätzel, M. Nano Lett. 2005, 5, 1315.
Suga, S.; Sayama, K.; Sugihara, H.; Arakawa, H. J. Phys. Chem. B (642) Schmidt-Mende, L.; Zakeeruddin, S. M.; Grätzel, M. Appl. Phys.
2003, 107, 597. Lett. 2005, 86, 013504/1.
(602) Hara, K.; Miyamoto, K.; Abe, Y.; Yanagida, M. J. Phys. Chem. B (643) Schmidt-Mende, L.; Campbell, W. M.; Wang, Q.; Jolley, K. W.;
2005, 109, 23776. Officer, D. L.; Nazeeruddin, M.; Grätzel, M. Chem. Phys. Chem.
(603) Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Yoshihara, T.; Murai, M.; 2005, 6, 1253.
Kurashige, M.; Ito, S.; Shinpo, A.; Suga, S.; Arakawa, H. AdV. Funct. (644) Shklover, V.; Haibach, T.; Bolliger, B.; Hochstrasser, M.; Erbudak,
Mater. 2005, 15, 246. M.; Nissen, H.; Zakeeruddin, S. M.; Nazeeruddin, M.; Grätzel, M.
(604) Heimer, T. A.; Bignozzi, C. A.; Meyer, G. J. J. Phys. Chem. 1993, J. Solid State Chem. 1997, 132, 60.
97, 11987. (645) Sprintschnik, G.; Sprintschnik, H. W.; Kirsch, P. P.; Whitten, D. G.
(605) Heimer, T. A.; D’Arcangelis, S. T.; Farzad, F.; Stipkala, J. M.; Meyer, J. Am. Chem. Soc. 1976, 98, 2337.
G. J. Inorg. Chem. 1996, 35, 5319. (646) Sprintschnik, G.; Sprintschnik, H. W.; Kirsch, P. P.; Whitten, D. G.
(606) Heimer, T. A.; Heilweil, E. J. J. Phys. Chem. B 1997, 101, 10990. J. Am. Chem. Soc. 1977, 99, 4947.
(607) Hermann, R.; Grätzel, M.; Nissen, H. U.; Shklover, V.; Nazeeruddin, (647) Tennakone, K.; Kumara, G. R. R. A.; Kottegoda, I. R. M.; Perera,
M. K.; Zakeeruddin, S. M.; Barbe, C.; Kay, A.; Haibach, T.; Steurer, V. P. S.; Weerasundara, P. S. R. S. J. Photochem. Photobiol., A 1998,
W. Chem. Mater. 1997, 9, 430. 117, 137.
(608) Jasieniak, J.; Johnston, M.; Waclawik, E. R. J. Phys. Chem. B 2004, (648) Tennakone, K.; Kumara, G. R. R. A.; Kottegoda, I. R. M.; Perera,
108, 12962. V. P. S. Chem. Commun. 1999, 15.
(609) Kalyanasundaram, K.; Nazeeruddin, M. K.; Grätzel, M.; Viscardi, (649) Tennakone, K.; Bandaranayake, P. K. M.; Jayaweera, P. V. V.;
G.; Savarino, P.; Barni, E. Inorg. Chim. Acta 1992, 198-200, 831. Konno, A.; Kumara, G. R. R. A. Phys. E 2002, 14, 190.
2956 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

(650) Vinodgopal, K.; Hua, X.; Dahlgren, R. L.; Lappin, A. G.; Patterson, (690) Weng, Y. X.; Wang, Y. Q.; Asbury, J. B.; Ghosh, H. N.; Lian, T. J.
L. K.; Kamat, P. V. J. Phys. Chem. 1995, 99, 10883. Phys. Chem. B 2000, 104, 93.
(651) Vlachopoulos, N.; Liska, P.; Augustynski, J.; Grätzel, M. J. Am. (691) Anderson, N. A.; Lian, T. Coord. Chem. ReV. 2004, 248, 1231.
Chem. Soc. 1988, 110, 1216. (692) Gregg, B. A. Coord. Chem. ReV. 2004, 248, 1215.
(652) Wang, P.; Zakeeruddin, S. M.; Humphry-Baker, R.; Moser, J. E.; (693) Katoh, R.; Furube, A.; Barzykin, A. V.; Arakawa, H.; Tachiya, M.
Grätzel, M. AdV. Mater. 2003, 15, 2101. Coord. Chem. ReV. 2004, 248, 1195.
(653) Wang, P.; Zakeeruddin, S. M.; Moser, J. E.; Nazeeruddin, M. K.; (694) Nelson, J.; Chandler, R. E. Coord. Chem. ReV. 2004, 248, 1181.
Sekiguchi, T.; Grätzel, M. Nat. Mater. 2003, 2, 402. (695) Cao, F.; Oskam, G.; Searson, P. C. J. Phys. Chem. 1996, 100, 17021.
(654) Wang, P.; Zakeeruddin, S. M.; Moser, J. E.; Humphry-Baker, R.; (696) deJongh, P. E.; Vanmaekelbergh, D. Phys. ReV. Lett. 1996, 77, 3427.
Comte, P.; Aranyos, V.; Hagfeldt, A.; Nazeeruddin, M. K.; Grätzel, (697) deJongh, P. E.; Vanmaekelbergh, D. J. Phys. Chem. B 1997, 101,
M. AdV. Mater. 2004, 16, 1806. 2716.
(655) Wang, P.; Zakeeruddin, S. M.; Humphry-Baker, R.; Grätzel, M. (698) Dloczik, L.; Ileperuma, O.; Lauermann, I.; Peter, L. M.; Ponomarev,
Chem. Mater. 2004, 16, 2694. E. A.; Redmond, G.; Shaw, N. J.; Uhlendorf, I. J. Phys. Chem. B
(656) Wang, P.; Dai, Q.; Zakeeruddin, S. M.; Forsyth, M.; MacFarlane, 1997, 101, 10281.
D. R.; Grätzel, M. J. Am. Chem. Soc. 2004, 126, 13590. (699) Kambili, A.; Walker, A. B.; Qiu, F. L.; Fisher, A. C.; Savin, A. D.;
(657) Wang, P.; Zakeeruddin, S. M.; Grätzel, M. J. Fluorine Chem. 2004, Peter, L. M. Physica E 2002, 14, 203.
125, 1241. (700) Van de Lagemaat, J.; Park, N. G.; Frank, A. J. J. Phys. Chem. B
(658) Wang, P.; Zakeeruddin, S. M.; Moser, J. E.; Humphry-Baker, R.; 2000, 104, 2044.
Grätzel, M. J. Am. Chem. Soc. 2004, 126, 7164. (701) Van de Lagemaat, J.; Frank, A. J. J. Phys. Chem. B 2000, 104, 4292.
(659) Wang, P.; Wenger, B.; Humphry-Baker, R.; Moser, J. E.; Teuscher, (702) Vanmaekelbergh, D.; De Jongh, P. E. Phys. ReV. B 2000, 61, 4699.
J.; Kantlehner, W.; Mezger, J.; Stoyanov, E. V.; Zakeeruddin, S. M.; (703) Duffy, N. W.; Peter, L. M.; Rajapakse, R. M. G.; Wijayantha, K. G.
Grätzel, M. J. Am. Chem. Soc. 2005, 127, 6850. U. J. Phys. Chem. B 2000, 104, 8916.
(660) Wang, Q.; Campbell, W. M.; Bonfantani, E. E.; Jolley, K. W.; Officer, (704) Fisher, A. C.; Peter, L. M.; Ponomarev, E. A.; Walker, A. B.;
D. L.; Walsh, P. J.; Gordon, K.; Humphry-Baker, R.; Nazeeruddin, Wijayantha, K. G. U. J. Phys. Chem. B 2000, 104, 949.
M. K.; Grätzel, M. J. Phys. Chem. B 2005, 109, 15397. (705) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F.; Ferriols,
(661) Wang, Q.; Moser, J. E.; Grätzel, M. J. Phys. Chem. B 2005, 109, N. S.; Bogdanoff, P.; Pereira, E. C. J. Phys. Chem. B 2000, 104,
14945. 2287.
(662) Wang, Z. S.; Kawauchi, H.; Kashima, T.; Arakawa, H. Coord. Chem. (706) Dittrich, T.; Weidmann, J.; Timoshenko, V. Y.; Petrov, A. A.; Koch,
ReV. 2004, 248, 1381. F.; Lisachenko, M. G.; Lebedev, E. Mater. Sci. Eng. B 2000, 69,
(663) Wang, Z. S.; Sayama, K.; Sugihara, H. J. Phys. Chem. B 2005, 109, 489.
22449. (707) Fabregat-Santiago, F.; Garcia-Belmonte, G.; Bisquert, J.; Zaban, A.;
(664) Wang, Z. S.; Yamaguchi, T.; Sugihara, H.; Arakawa, H. Langmuir Salvador, P. J. Phys. Chem. B 2002, 106, 334.
2005, 21, 4272. (708) Kern, R.; Sastrawan, R.; Ferber, J.; Stangl, R.; Luther, J. Electrochim.
(665) Yan, S. G.; Hupp, J. T. J. Phys. Chem. 1996, 100, 6867. Acta 2002, 47, 4213.
(666) Yanagida, M.; Islam, A.; Tachibana, Y.; Fujihashi, G.; Katoh, R.; (709) Kron, G.; Egerter, T.; Werner, J. H.; Rau, U. J. Phys. Chem. B 2003,
Sugihara, H.; Arakawa, H. New J. Chem. 2002, 26, 963. 107, 3556.
(667) Yanagida, M.; Yamaguchi, T.; Kurashige, M.; Hara, K.; Katoh, R.; (710) Schwarzburg, K.; Willig, F. J. Phys. Chem. B 2003, 107, 3552.
Sugihara, H.; Arakawa, H. Inorg. Chem. 2003, 42, 7921. (711) Benkstein, K. D.; Kopidakis, N.; van de Lagemaat, J.; Frank, A. J.
(668) Yang, M.; Thompson, D. W.; Meyer, G. J. Inorg. Chem. 2000, 39, J. Phys. Chem. B 2003, 107, 7759.
3738. (712) Dittrich, T.; Lebedev, E. A.; Weidmann, J. Phys. Status Solidi A 1998,
(669) Yang, M.; Thompson, D. W.; Meyer, G. J. Inorg. Chem. 2002, 41, 165, R5.
1254. (713) Dittrich, T. Phys. Status Solidi A 2000, 182, 447.
(670) Zaban, A.; Chen, S. G.; Chappel, S.; Gregg, B. A. Chem. Commun. (714) Kopidakis, N.; Schiff, E. A.; Park, N. G.; van de Lagemaat, J.; Frank,
2000, 2231. A. J. J. Phys. Chem. B 2000, 104, 3930.
(671) Zafer, C.; Karapire, C.; Serdar Sariciftci, N.; Icli, S. Sol. Energy (715) Kopidakis, N.; Benkstein, K. D.; van de Lagemaat, J.; Frank, A. J.
Mater. Sol. Cells 2005, 88, 11. J. Phys. Chem. B 2003, 107, 11307.
(672) Zakeeruddin, S. M.; Nazeeruddin, M. K.; Pechy, P.; Rotzinger, F. (716) Solbrand, A.; Lindstrom, H.; Rensmo, H.; Hagfeldt, A.; Lindquist,
P.; Humphry-Baker, R.; Kalyanasundaram, K.; Grätzel, M.; Shklover, S. E.; Sodergren, S. J. Phys. Chem. B 1997, 101, 2514.
V.; Haibach, T. Inorg. Chem. 1997, 36, 5937. (717) Solbrand, A.; Henningsson, A.; Sodergren, S.; Lindstrom, H.;
(673) Zakeeruddin, S. M.; Nazeeruddin, M.; Humphry-Baker, R.; Pechy, Hagfeldt, A.; Lindquist, S. E. J. Phys. Chem. B 1999, 103, 1078.
P.; Quagliotto, P.; Barolo, C.; Viscardi, G.; Grätzel, M. Langmuir (718) van de Lagemaat, J.; Frank, A. J. J. Phys. Chem. B 2001, 105, 11194.
2002, 18, 952. (719) Dittrich, T.; Duzhko, V.; Koch, F.; Kytin, V.; Rappich, J. Phys. ReV.
(674) Qu, P.; Meyer, G. J. Langmuir 2001, 17, 6720. B 2002, 65.
(675) Blackbourn, R. L.; Johnson, C. S.; Hupp, J. T. J. Am. Chem. Soc. (720) Peter, L. M.; Walker, A. B.; Boschloo, G.; Hagfeldt, A. J. Phys.
1991, 113, 1060. Chem. B 2006, 110, 13694.
(676) Frank, A. J.; Kopidakis, N.; Van de Lagemaat, J. Coord. Chem. ReV. (721) Cahen, D.; Hodes, G.; Grätzel, M.; Guillemoles, J. F.; Riess, I. J.
2004, 248, 1165. Phys. Chem. B 2000, 104, 2053.
(677) Stipkala, J. M.; Heimer, T. A.; Kelly, C. A.; Livi, K. J. T.; Meyer, (722) Franco, G.; Gehring, J.; Peter, L. M.; Ponomarev, E. A.; Uhlendorf,
G. J. Chem. Mater. 1997, 9, 2341. I. J. Phys. Chem. B 1999, 103, 692.
(678) Benkö, G.; Kallioinen, J.; Korppi-Tommola, J. E. I.; Yartsev, A. P.; (723) Bisquert, J.; Vikhrenko, V. S. J. Phys. Chem. B 2004, 108, 2313.
Sundstrom, V. J. Am. Chem. Soc. 2002, 124, 489. (724) Kopidakis, N.; Neale, N. R.; Zhu, K.; van de Lagemaat, J.; Frank,
(679) Kelly, C. A.; Farzad, F.; Thompson, D. W.; Stipkala, J. M.; Meyer, A. J. Appl. Phys. Lett. 2005, 87, 202106/1.
G. J. Langmuir 1999, 15, 7047. (725) Gupta, K. K.; Tripathi, V. S.; Ram, H.; Raj, H. Colourage 2002, 49,
(680) Tachibana, Y.; Rubtsov, I. V.; Montanari, I.; Yoshihara, K.; Klug, 35.
D. R.; Durrant, J. R. J. Photochem. Photobiol., A 2001, 142, 215. (726) Hwang, D. K.; Moon, J. H.; Shul, Y. G.; Jung, K. T.; Kim, D. H.;
(681) Tachibana, Y.; Moser, J. E.; Grätzel, M.; Klug, D. R.; Durrant, J. R. Lee, D. W. J. Sol-Gel Sci. Technol. 2003, 26, 783.
J. Phys. Chem. 1996, 100, 20056. (727) Kim, J. W.; Shim, J. W.; Bae, J. H.; Han, S. H.; Kim, H. K.; Chang,
(682) Tachibana, Y.; Haque, S. A.; Mercer, I. P.; Durrant, J. R.; Klug, D. I. S.; Kang, H. H.; Suh, K. D. Colloid Polym. Sci. 2002, 280, 584.
R. J. Phys. Chem. B 2000, 104, 1198. (728) Mahltig, B.; Boettcher, H.; Rauch, K.; Dieckmann, U.; Nitsche, R.;
(683) Tachibana, Y.; Haque, S. A.; Mercer, I. P.; Moser, J. E.; Klug, D. Fritz, T. Thin Solid Films 2005, 485, 108.
R.; Durrant, J. R. J. Phys. Chem. B 2001, 105, 7424. (729) Popov, A. P.; Priezzhev, A. V.; Lademann, J.; Myllylae, R. J. Phys.
(684) Rehm, J. M.; McLendon, G. L.; Nagasawa, Y.; Yoshihara, K.; Moser, D 2005, 38, 2564.
J.; Grätzel, M. J. Phys. Chem. 1996, 100, 9577. (730) Zhang, Y.; Zhang, L.; Mo, C.; Li, Y.; Yao, L.; Cai, W. J. Mater.
(685) Paulsson, H.; Kloo, L.; Hagfeldt, A.; Boschloo, G. J. Electroanal. Sci. Technol. (Shenyang, China) 2000, 16, 277.
Chem. 2006, 586, 56. (731) Meng, S.; Zhang, Z. Y.; Kaxiras, E. Phys. ReV. Lett. 2006, 97.
(686) Schwarzburg, K.; Ernstorfer, R.; Felber, S.; Willig, F. Coord. Chem. (732) Sirghi, L.; Aoki, T.; Hatanaka, Y. Surf. ReV. Lett. 2003, 10, 345.
ReV. 2004, 248, 1259. (733) Takata, Y.; Hidaka, S.; Cao, J. M.; Tanaka, K.; Masuda, M.; Ito, T.;
(687) Wenger, B.; Grätzel, M.; Moser, J. E. J. Am. Chem. Soc. 2005, 127, Watanabe, T.; Shimohigoshi, M. Therm. Sci. Eng. 2000, 8, 33.
12150. (734) Takata, Y.; Hidaka, S.; Masuda, M.; Ito, T. Int. J. Energy Res. 2003,
(688) Thompson, D. W.; Kelly, C. A.; Farzad, F.; Meyer, G. J. Langmuir 27, 111.
1999, 15, 650. (735) Bozzi, A.; Yuranova, T.; Kiwi, J. J. Photochem. Photobiol., A 2005,
(689) Vrachnou, E.; Vlachopoulos, N.; Grätzel, M. Chem. Commun. 1987, 172, 27.
868. (736) Cassar, L. MRS Bull. 2004, 29, 328.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2957

(737) Fujishima, A.; Rao, T. N.; Tryk, D. A. Int. Glass ReV. 2002, 128. (786) Canle Lopez, M.; Fernandez, M. I.; Rodriguez, S.; Santaballa, J. A.;
(738) Guan, K. Surf. Coat. Technol. 2005, 191, 155. Steenken, S.; Vulliet, E. Chem. Phys. Chem. 2005, 6, 2064.
(739) Meilert, K. T.; Laub, D.; Kiwi, J. J. Mol. Catal. A 2005, 237, 101. (787) Canle, L.; Santaballa, J. A.; Vulliet, E. J. Photochem. Photobiol., A
(740) Nakajima, A.; Hashimoto, K.; Watanabe, T.; Takai, K.; Yamauchi, 2005, 175, 192.
G.; Fujishima, A. Langmuir 2000, 16, 7044. (788) Chatterjee, D.; Dasgupta, S. J. Photochem. Photobiol., C 2005, 6,
(741) Ohko, Y.; Saitoh, S.; Tatsuma, T.; Fujishima, A. J. Electrochem. 186.
Soc. 2001, 148, B24. (789) Fu, J. f.; Ji, M.; An, D. n. J. EnViron. Sci. (Beijing, China) 2005, 17,
(742) Sam, E. D.; Urgen, M.; Tepehan, F. Z.; Gunay, V. Key Eng. Mater. 942.
2004, 264-268, 407. (790) Hidaka, H.; Horikoshi, S.; Oyama, T.; Watanabe, N.; Serpone, N.
(743) Smith, S. Chem. ReV. (Deddington, U.K.) 2002, 12, 31. Photo/Electrochem. Photobiol. EnViron., Energy Fuel 2003, 101.
(744) Zhang, X. T.; Sato, O.; Taguchi, M.; Einaga, Y.; Murakami, T.; (791) Hidaka, H.; Koike, T.; Kurihara, T.; Serpone, N. New J. Chem. 2004,
Fujishima, A. Chem. Mater. 2005, 17, 696. 28, 1100.
(745) Miyazaki, H.; Hyodo, T.; Shimizu, Y.; Egashira, M. Sens. Actuators, (792) Hidaka, H.; Watanabe, N.; Horikoshi, S. Trends Air Pollut. Res. 2005,
B 2005, B108, 467. 157.
(746) Ruiz, A. M.; Cornet, A.; Morante, J. R. Sens. Actuators, B 2005, (793) Kikuchi, Y.; Sunada, K.; Iyoda, T.; Hashimoto, K.; Fujishima, A. J.
B111-B112, 7. Photochem. Photobiol., A 1997, 106, 51.
(747) Ruiz, A. M.; Cornet, A.; Shimanoe, K.; Morante, J. R.; Yamazoe, (794) Lhomme, L.; Brosillon, S.; Wolbert, D.; Dussaud, J. Appl. Catal. B
N. Sens. Actuators, B 2005, B109, 7. 2005, 61, 227.
(748) Tai, W. P. Nanomaterials 2005, 115. (795) Li, L.; Zhu, W.; Chen, L.; Zhang, P.; Chen, Z. J. Photochem.
(749) Tokudome, H.; Yamada, Y.; Sonezaki, S.; Ishikawa, H.; Bekki, M.; Photobiol., A 2005, 175, 172.
Kanehira, K.; Miyauchi, M. Appl. Phys. Lett. 2005, 87, 213901/1. (796) Li, L.; Wu, Q. Y.; Guo, Y. H.; Hu, C. W. Microporous Mesoporous
(750) Yadav, B. C.; Shukla, R. K.; Bali, L. M. Indian J. Pure Appl. Phys. Mater. 2005, 87, 1.
2005, 43, 51. (797) Li, S. X.; Zheng, F. Y.; Liu, X. L.; Wu, F.; Deng, N. S.; Yang, J. H.
(751) Si, S. H.; Fung, Y. S.; Zhu, D. R. Sens. Actuators, B 2005, B108, Chemosphere 2005, 61, 589.
165. (798) Liau, L. C.-K.; Chang, H.; Yang, T. C.-K.; Huang, C. L. J. Chem.
(752) Granqvist, C. G. AdV. Mater. 2003, 15, 1789. Eng. Jpn. 2005, 38, 813.
(753) Abe, R.; Sayama, K.; Domen, K.; Arakawa, H. Chem. Phys. Lett. (799) Mahmoodi, N. M.; Arami, M.; Limaee, N. Y.; Tabrizi, N. S. Chem.
2001, 344, 339. Eng. J. 2005, 112, 191.
(754) Abe, R.; Sayama, K.; Arakawa, H. Chem. Phys. Lett. 2003, 371, (800) Mahmoodi, N. M.; Arami, M.; Limaee, N. Y.; Tabrizi, N. S. J. Colloid
360. Interface Sci. 2006, 295, 159.
(755) Abe, R.; Sayama, K.; Sugihara, H. J. Phys. Chem. B 2005, 109, (801) Mohanty, S.; Rao, N. N.; Khare, P.; Kaul, S. N. Water Res. 2005,
16052. 39, 5064.
(756) Abe, R.; Sayama, K.; Sugihara, H. J. Sol. Energy Eng. 2005, 127, (802) Mozia, S.; Tomaszewska, M.; Kosowska, B.; Grzmil, B.; Morawski,
413. A. W.; Kalucki, K. Appl. Catal. B 2005, 55, 195.
(757) Galinska, A.; Walendziewski, J. Energy Fuels 2005, 19, 1143. (803) Mozia, S.; Tomaszewska, M.; Morawski, A. W. Desalination 2005,
185, 449.
(758) Hartig, K. J.; Getoff, N. AdV. Hydrogen Energy 1984, 4, 1085.
(804) Nguyen, V. N. H.; Amal, R.; Beydoun, D. Chem. Eng. Sci. 2003,
(759) Khan, S. U. M.; Akikusa, J. J. Electrochem. Soc. 1998, 145, 89.
58, 4429.
(760) Kiwi, J. Chem. Phys. Lett. 1981, 83, 594.
(805) Nguyen, V. N. H.; Beydoun, D.; Amal, R. J. Photochem. Photobiol.,
(761) Kudo, A. Catal. SurV. Asia 2003, 7, 31. A 2005, 171, 113.
(762) Matsumoto, Y.; Unal, U.; Tanaka, N.; Kudo, A.; Kato, H. J. Solid (806) Parga, J. R.; Shukla, S. S.; Cocke, D. L. Res. J. Chem. EnViron.
State Chem. 2004, 177, 4205. 2005, 9, 60.
(763) Matsuoka, M.; Kitano, M.; Takeuchi, M.; Anpo, M.; Thomas, J. M. (807) Rahman, M. A.; Muneer, M. Desalination 2005, 181, 161.
Top. Catal. 2005, 35, 305. (808) Raja, P.; Bandara, J.; Giordano, P.; Kiwi, J. Ind. Eng. Chem. Res.
(764) Matsuoka, M.; Kitano, M.; Takeuchi, M.; Anpo, M.; Thomas, J. M. 2005, 44, 8959.
Mater. Sci. Forum 2005, 486-487, 81. (809) Rengaraj, S.; Li, X. Z. J. Mol. Catal. A 2006, 243, 60.
(765) Park, J. H.; Bard, A. J. Electrochem. Solid-State Lett. 2005, 8, G371. (810) Sahoo, C.; Gupta, A. K.; Pal, A. Desalination 2005, 181, 91.
(766) Pujadas, M.; Salvador, P. J. Electrochem. Soc. 1989, 136, 716. (811) Sonawane, R. S.; Dongare, M. K. J. Mol. Catal. A 2006, 243, 68.
(767) Salvador, P. New J. Chem. 1988, 12, 35. (812) Tanimura, T.; Yoshida, A.; Yamazaki, S. Appl. Catal. B 2005, 61,
(768) Schwarz, A.; Hartig, K. J.; Getoff, N. AdV. Hydrogen Energy 1986, 346.
5, 643. (813) Thevenet, F.; Guaitella, O.; Herrmann, J. M.; Rousseau, A.; Guillard,
(769) Tafalla, D.; Salvador, P. J. Electroanal. Chem. Interfacial Electro- C. Appl. Catal. B 2005, 61, 58.
chem. 1989, 270, 285. (814) Thiruvenkatachari, R.; Kwon, T. O.; Moon, I. S. Sep. Sci. Technol.
(770) Takabayashi, S.; Nakamura, R.; Nakato, Y. J. Photochem. Photobiol., 2005, 40, 2871.
A 2004, 166, 107. (815) Wang, C. C.; Zhang, Z.; Ying, J. Y. Nanostruct. Mater. 1997, 9,
(771) Biancardo, M.; Argazzi, R.; Bignozzi, C. A. Inorg. Chem. 2005, 44, 583.
9619. (816) Wang, J.; Wen, F. Y.; Zhang, Z. H.; Zhang, X. D.; Pan, Z. J.; Zhang,
(772) Chen, M. J.; Shen, H. Acta Metall. Sin. (Engl. Lett.) 2005, 18, 275. L.; Wang, L.; Xu, L.; Kang, P. L.; Zhang, P. J. EnViron. Sci. (Beijing,
(773) Freestone, N. Chem. Ind. (London, U.K.) 2005, 27. China) 2005, 17, 727.
(774) He, T.; Ma, Y.; Cao, Y.; Liu, H.; Yang, W.; Yao, J. J. Colloid (817) Yang, J.; Chen, C.; Ji, H.; Ma, W.; Zhao, J. J. Phys. Chem. B 2005,
Interface Sci. 2004, 279, 117. 109, 21900.
(775) Iuchi, K.; Ohko, Y.; Tatsuma, T.; Fujishima, A. Chem. Mater. 2004, (818) Yang, S. G.; Quan, X.; Li, X. Y.; Fang, N.; Zhang, N.; Zhao, H. M.
16, 1165. J. EnViron. Sci. (Beijing, China) 2005, 17, 290.
(776) Okumu, J.; Dahmen, C.; Sprafke, A. N.; Luysberg, M.; von Plessen, (819) Yang, S. Y.; Chen, Y. X.; Lou, L. P.; Wu, X. N. J. EnViron. Sci.
G.; Wuttig, M. J. Appl. Phys. 2005, 97, 094305/1. (Beijing, China) 2005, 17, 761.
(777) Palgrave, R. G.; Parkin, I. P. J. Mater. Chem. 2004, 14, 2864. (820) Zhang, A. P.; Sun, Y. P. World J. Gastroenterol. 2004, 10, 3191.
(778) Akbal, F. EnViron. Prog. 2005, 24, 317. (821) Zhang, W.; Wang, X. X.; Fu, X. Z. Chin. Chem. Lett. 2005, 16,
(779) Akurati, K. K.; Vital, A.; Hany, R.; Bommer, B.; Graule, T.; Winterer, 1275.
M. Int. J. Photoenergy 2005, 7, 153. (822) Zhu, J.; Zhang, J.; Chen, F.; Iino, K.; Anpo, M. Top. Catal. 2005,
(780) Augugliaro, V.; Coluccia, S.; Garcia-Lopez, E.; Loddo, V.; Marci, 35, 261.
G.; Martra, G.; Palmisano, L.; Schiavello, M. Top. Catal. 2005, 35, (823) Huang, N. P.; Xu, M. H.; Yuan, C. W.; Yu, R. R. J. Photochem.
237. Photobiol., A 1997, 108, 229.
(781) Augugliaro, V.; Garcia-Lopez, E.; Loddo, V.; Malato-Rodriguez, S.; (824) Ivankovic, S.; Gotic, M.; Jurin, M.; Music, S. J. Sol-Gel Sci. Technol.
Maldonado, I.; Marci, G.; Molinari, R.; Palmisano, L. Sol. Energy 2003, 27, 225.
2005, 79, 402. (825) Sakai, H.; Baba, R.; Hashimoto, K.; Kubota, Y.; Fujishima, A. Chem.
(782) Blake, D. M.; Maness, P. C.; Huang, Z.; Wolfrum, E. J.; Huang, J.; Lett. 1995, 185.
Jacoby, W. A. Sep. Purif. Methods 1999, 28, 1. (826) Zhang, Z.; Wang, C. C.; Zakaria, R.; Ying, J. Y. J. Phys. Chem. B
(783) Bosc, F.; Edwards, D.; Keller, N.; Keller, V.; Ayral, A. Thin Solid 1998, 102, 10871.
Films 2006, 495, 272. (827) Egerton, T. A.; Kosa, S. A. M.; Christensen, P. A. Phys. Chem. Chem.
(784) Boujday, S.; Wunsch, F.; Portes, P.; Bocquet, J. F.; Colbeau-Justin, Phys. 2006, 8, 398.
C. Sol. Energy Mater. Sol. Cells 2004, 83, 421. (828) Kemp, T. J.; McIntyre, R. A. Polym. Degrad. Stab. 2005, 91, 165.
(785) Cai, R.; Hashimoto, K.; Kubota, Y.; Fujishima, A. Chem. Lett. 1992, (829) Kim, D. H.; Woo, S. I.; Moon, S. H.; Kim, H. D.; Kim, B. Y.; Cho,
427. J. H.; Joh, Y. G.; Kim, E. C. Solid State Commun. 2005, 136, 554.
2958 Chemical Reviews, 2007, Vol. 107, No. 7 Chen and Mao

(830) Kim, S.; Hwang, S. J.; Choi, W. J. Phys. Chem. B 2005, 109, 24260. (874) Rothenberger, G.; Comte, P.; Grätzel, M. Sol. Energy Mater. Sol.
(831) Park, M. S.; Kwon, S. K.; Min, B. I. Phys. ReV. B 2002, 65, 161201/ Cells 1999, 58, 321.
1. (875) Usami, A. Sol. Energy Mater. Sol. Cells 1999, 59, 163.
(832) Ranjit, K. T.; Viswanathan, B. J. Photochem. Photobiol., A 1997, (876) Usami, A. Sol. Energy Mater. Sol. Cells 2000, 62, 239.
108, 79. (877) Usami, A. Chem. Phys. Lett. 1997, 277, 105.
(833) Wei, H.; Wu, Y.; Lun, N.; Zhao, F. J. Mater. Sci. 2004, 39, 1305. (878) Usami, A. Sol. Energy Mater. Sol. Cells 2000, 64, 73.
(834) Xia, C.; Zhou, Y.; Li, X.; Zeng, J.; Xu, R. Rare Met. (Beijing, China) (879) Tocci, M. D.; Bloemer, M. J.; Scalora, M.; Bowden, C. M.; Dowling,
2005, 24, 358. J. P. NATO ASI Ser. E: Appl. Sci. 1996, 324, 237.
(835) Xu, J. C.; Lu, M.; Guo, X. Y.; Li, H. L. J. Mol. Catal. A 2005, 226, (880) Vlasov, Y.; Petit, S.; Klein, G.; Honerlage, B.; Hirlimann, C. Phys.
123. ReV. E 1999, 60, 1030.
(836) Xu, K. J. Nat. Gas Chem. 2005, 14, 168. (881) McFarland, E. W.; Tang, J. Nature 2003, 421, 616.
(837) Chen, X. Q.; Yang, J. Y.; Zhang, J. S. J. Cent. South UniV. Technol. (882) Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K. Renewable
2004, 11, 161. Sustainable Energy ReV. 2007, 11, 401.
(838) Gandhe, A. R.; Naik, S. P.; Fernandes, J. B. Microporous Mesoporous (883) Hameed, A.; Gondal, M. A. J. Mol. Catal. A 2004, 219, 109.
Mater. 2005, 87, 103. (884) Yamakata, A.; Ishibashi, T. A.; Onishi, H. J. Mol. Catal. A 2003,
(839) Irie, H.; Washizuka, S.; Watanabe, Y.; Kako, T.; Hashimoto, K. J. 199, 85.
Electrochem. Soc. 2005, 152, E351. (885) Yamakata, A.; Ishibashi, T. A.; Onishi, H. Int. J. Photoenergy 2003,
(840) Morawski, A.; Janus, W., M. Pol. J. Chem. Technol. 2005, 7, 81. 5, 7.
(841) Mrowetz, M.; Balcerski, W.; Colussi, A. J.; Hoffmann, M. R. J. Phys. (886) Gondal, M. A.; Hameed, A.; Yamani, Z. H.; Suwaiyan, A. Chem.
Chem. B 2004, 108, 17269. Phys. Lett. 2004, 385, 111.
(842) Sato, S.; White, J. M. Chem. Phys. Lett. 1980, 72, 83. (887) Hameed, A.; Gondal, M. A.; Yamani, Z. H.; Yahya, A. H. J. Mol.
(843) Tachikawa, T.; Tojo, S.; Kawai, K.; Endo, M.; Fujitsuka, M.; Ohno, Catal. A 2005, 227, 241.
T.; Nishijima, K.; Miyamoto, Z.; Majima, T. J. Phys. Chem. B 2004, (888) Sayama, K.; Arakawa, H. Faraday Trans. 1997, 93, 1647.
108, 19299. (889) Moon, S. C.; Mametsuka, H.; Tabata, S.; Suzuki, E. Catal. Today
(844) Wang, H.; Lewis, J. P. J. Phys.: Condens. Matter 2005, 17, L209. 2000, 58, 125.
(845) Wong, M. S.; Chou, H. P.; Yang, T. S. Thin Solid Films 2005, 494, (890) Lee, S. G.; Lee, S.; Lee, H. I. Appl. Catal. A 2001, 207, 173.
244. (891) Li, Y.; Lu, G.; Li, S. Chemosphere 2003, 52, 843.
(846) Yang, M. C.; Yang, T. S.; Wong, M. S. Thin Solid Films 2004, 469- (892) Fujihara, K.; Ohno, T.; Matsumura, M. Faraday Trans. 1998, 94,
470, 1. 3705.
(847) Yang, S.; Gao, L. J. Am. Ceram. Soc. 2004, 87, 1803. (893) Jin, Z. L.; Lu, G. X. Energy Fuels 2005, 19, 1126.
(848) Sano, T.; Negishi, N.; Koike, K.; Takeuchi, K.; Matsuzawa, S. J. (894) Jing, D.; Zhang, Y.; Guo, L. Chem. Phys. Lett. 2005, 415, 74.
Mater. Chem. 2004, 14, 380. (895) Fujishima, A. Science 2003, 301, 1673.
(849) Takeshita, K.; Yamakata, A.; Ishibashi, T. A.; Onishi, H.; Nishijima, (896) Hagglund, C.; Grätzel, M.; Kasemo, B. Science 2003, 301, 1673.
K.; Ohno, T. J. Photochem. Photobiol., A 2006, 177, 269. (897) Lackner, K. S. Science 2003, 301, 1673.
(850) Nazeeruddin, M. K.; Zakeeruddin, S. M.; Humphry-Baker, R.; (898) Duonghong, D.; Borgarello, E.; Grätzel, M. J. Am. Chem. Soc. 1981,
Jirousek, M.; Liska, P.; Vlachopoulos, N.; Shklover, V.; Fischer, C. 103, 4685.
H.; Grätzel, M. Inorg. Chem. 1999, 38, 6298. (899) Akikusa, J.; Khan, S. U. M. Int. J. Hydrogen Energy 2002, 27, 863.
(851) Ohsaki, Y.; Masaki, N.; Kitamura, T.; Wada, Y.; Okamoto, T.; (900) Aliev, A. E.; Shin, H. W. Solid State Ionics 2002, 154-155, 425.
Sekino, T.; Niihara, K.; Yanagida, S. Phys. Chem. Chem. Phys. 2005, (901) Bach, U.; Corr, D.; Lupo, D.; Pichot, F.; Ryan, M. AdV. Mater. 2002,
7, 4157. 14, 845.
(852) Han, H.; Zan, L.; Zhong, J.; Zhao, X. J. Mater. Sci. 2005, 40, 4921. (902) Bonhote, P.; Gogniat, E.; Campus, F.; Walder, L.; Grätzel, M.
(853) Han, H.; Zhao, X.; Liu, J. J. Electrochem. Soc. 2005, 152, A164. Displays 1999, 20, 137.
(854) Kang, T. S.; Moon, S. H.; Kim, K. J. J. Electrochem. Soc. 2002, (903) Bonhote, P.; Gogniat, E.; Grätzel, M.; Ashrit, P. V. Thin Solid Films
149, E155. 1999, 350, 269.
(855) Hagfeldt, A.; Grätzel, M.; Nogueria, A. F.; Furtado, L. F. O.; Formiga, (904) Campus, F.; Bonhote, P.; Grätzel, M.; Heinen, S.; Walder, L. Sol.
A. L. B.; Nakamura, M.; Araki, K.; Toma, H. E.; Panigrahi, S.; Pal, Energy Mater. Sol. Cells 1999, 56, 281.
T. Chemtracts 2004, 17, 175. (905) Choi, S. Y.; Mamak, M.; Coombs, N.; Chopra, N.; Ozin, G. A. Nano
(856) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F. J. Solid State Lett. 2004, 4, 1231.
Electrochem. 1999, 3, 337. (906) Cinnsealach, R.; Boschloo, G.; Rao, S. N.; Fitzmaurice, D. Sol.
(857) Hagfeldt, A.; Lindquist, S. E.; Grätzel, M. Sol. Energy Mater. Sol. Energy Mater. Sol. Cells 1998, 55, 215.
Cells 1994, 32, 245. (907) Cinnsealach, R.; Boschloo, G.; Nagaraja Rao, S.; Fitzmaurice, D.
(858) Zaban, A.; Meier, A.; Gregg, B. A. J. Phys. Chem. B 1997, 101, Sol. Energy Mater. Sol. Cells 1999, 57, 107.
7985. (908) Cummins, D.; Boschloo, G.; Ryan, M.; Corr, D.; Rao, S. N.;
(859) Kavan, L.; Grätzel, M.; Gilbert, S. E.; Klemenz, C.; Scheel, H. J. J. Fitzmaurice, D. J. Phys. Chem. B 2000, 104, 11449.
Am. Chem. Soc. 1996, 118, 6716. (909) Deb, S. K.; Lee, S. H.; Tracy, C. E.; Pitts, J. R.; Gregg, B. A.; Branz,
(860) Chen, S. G.; Chappel, S.; Diamant, Y.; Zaban, A. Chem. Mater. 2001, H. M. Electrochim. Acta 2001, 46, 2125.
13, 4629. (910) Garcia-Canadas, J.; Peter, L. M.; Upul Wijayantha, K. G. Electro-
(861) Diamant, Y.; Chen, S. G.; Melamed, O.; Zaban, A. J. Phys. Chem. chem. Commun. 2003, 5, 199.
B 2003, 107, 1977. (911) Garcia-Canadas, J.; Fabregat-Santiago, F.; Kapla, J.; Bisquert, J.;
(862) Diamant, Y.; Chappel, S.; Chen, S. G.; Melamed, O.; Zaban, A. Garcia-Belmonte, G.; Mora-Sero, I.; Edwards, M. O. M. Electrochim.
Coord. Chem. ReV. 2004, 248, 1271. Acta 2004, 49, 745.
(863) Kay, A.; Grätzel, M. Chem. Mater. 2002, 14, 2930. (912) Hagfeldt, A.; Vlachopoulos, N.; Grätzel, M. J. Electrochem. Soc.
(864) Nasr, C.; Kamat, P. V.; Hotchandani, S. J. Phys. Chem. B 1998, 1994, 141, L82.
102, 10047. (913) Moeller, M.; Asaftei, S.; Corr, D.; Ryan, M.; Walder, L. AdV. Mater.
(865) Palomares, E.; Clifford, J. N.; Haque, S. A.; Lutz, T.; Durrant, J. R. 2004, 16, 1558.
J. Am. Chem. Soc. 2003, 125, 475. (914) Ottaviani, M.; Panero, S.; Morzilli, S.; Scrosati, B.; Lazzari, M. Solid
(866) Park, N. G.; Kang, M. G.; Kim, K. M.; Ryu, K. S.; Chang, S. H.; State Ionics 1986, 20, 197.
Kim, D. K.; van de Lagemaat, J.; Benkstein, K. D.; Frank, A. J. (915) Turhan, I.; Tepehan, F. Z.; Tepehan, G. G. J. Mater. Sci. 2005, 40,
Langmuir 2004, 20, 4246. 1359.
(867) Chappel, S.; Chen, S. G.; Zaban, A. Langmuir 2002, 18, 3336. (916) Vayssieres, L.; Alfredsson, Y.; Siegbahn, H. Electrochem. Solid-
(868) Fabregat-Santiago, F.; Garcia-Canadas, J.; Palomares, E.; Clifford, State Lett. 1999, 2, 648.
J. N.; Haque, S. A.; Durrant, J. R.; Garcia-Belmonte, G.; Bisquert, (917) Cusack, L.; Marguerettaz, X.; Rao, S. N.; Wenger, J.; Fitzmaurice,
J. J. Appl. Phys. 2004, 96, 6903. D. Chem. Mater. 1997, 9, 1765.
(869) O’Regan, B. C.; Scully, S.; Mayer, A. C.; Palomares, E.; Durrant, J. (918) Marguerettaz, X.; O’Neill, R.; Fitzmaurice, D. J. Am. Chem. Soc.
J. Phys. Chem. B 2005, 109, 4616. 1994, 116, 2629.
(870) Palomares, E.; Clifford, J. N.; Haque, S. A.; Lutz, T.; Durrant, J. R. (919) Marguerettaz, X.; Redmond, G.; Rao, S. N.; Fitzmaurice, D. Chem.s
Chem. Commun. 2002, 1464. Eur. J. 1996, 2, 420.
(871) Kumara, G. R. A.; Okuya, M.; Murakami, K.; Kaneko, S.; Jayaweera, (920) Hu, X.; Skadtchenko, B. O.; Trudeau, M.; Antonelli, D. M. J. Am.
V. V.; Tennakone, K. J. Photochem. Photobiol., A 2004, 164, 183. Chem. Soc. 2006, 128, 11740.
(872) Chappel, S.; Grinis, L.; Ofir, A.; Zaban, A. J. Phys. Chem. B 2005, (921) Benkstein, K. D.; Semancik, S. Sens. Actuators, B 2006, 113, 445.
109, 1643. (922) Birkefeld, L. D.; Azad, A. M.; Akbar, S. A. J. Am. Ceram. Soc.
(873) Ferber, J.; Luther, J. Sol. Energy Mater. Sol. Cells 1998, 54, 265. 1992, 75, 2964.
Titanium Dioxide Nanomaterials Chemical Reviews, 2007, Vol. 107, No. 7 2959

(923) Carney, C. M.; Yoo, S.; Akbar, S. A. Sens. Actuators, B 2005, B108, (942) Li, Y. X.; Galatsis, K.; Wlodarski, W.; Passacantando, M.; Santucci,
29. S.; Siciliano, P.; Catalano, M. Sens. Actuators, B 2001, B77, 27.
(924) Carotta, M. C.; Ferroni, M.; Gnani, D.; Guidi, V.; Merli, M.; (943) Montesperelli, G.; Pumo, A.; Traversa, E.; Gusmano, G.; Bearzotti,
Martinelli, G.; Casale, M. C.; Notaro, M. Sens. Actuators, B 1999, A.; Montenero, A.; Gnappi, G. Sens. Actuators, B 1995, B25, 705.
B58, 310. (944) Rothschild, A.; Edelman, F.; Komem, Y.; Cosandey, F. Sens.
(925) Chow, L. L. W.; Yuen, M. M. F.; Chan, P. C. H.; Cheung, A. T. Actuators, B 2000, B67, 282.
Sens. Actuators, B 2001, B76, 310. (945) Savage, N. O.; Akbar, S. A.; Dutta, P. K. Sens. Actuators, B 2001,
(926) Comini, E.; Faglia, G.; Sberveglieri, G.; Li, Y. X.; Wlodarski, W.; B72, 239.
Ghantasala, M. K. Sens. Actuators, B 2000, B64, 169. (946) Sberveglieri, G.; Comini, E.; Faglia, G.; Atashbar, M. Z.; Wlodarski,
(927) Comini, E.; Guidi, V.; Frigeri, C.; Ricco, I.; Sberveglieri, G. Sens. W. Sens. Actuators, B 2000, B66, 139.
Actuators, B 2001, B77, 16. (947) Schierbaum, K. D.; Kirner, U. K.; Geiger, J. F.; Goepel, W. Sens.
(928) Demarne, V.; Balkanova, S.; Grisel, A.; Rosenfeld, D.; Levy, F. Sens. Actuators, B 1991, B4, 87.
Actuators, B 1993, 14, 497. (948) Sharma, R. K.; Bhatnagar, M. C.; Sharma, G. L. Sens. Actuators, B
(929) Devi, G. S.; Hyodo, T.; Shimizu, Y.; Egashira, M. Sens. Actuators, 1998, B46, 194.
B 2002, B87, 122. (949) Sharma, R. K.; Bhatnagar, M. C. Sens. Actuators, B 1999, B56, 215.
(930) Ferroni, M.; Carotta, M. C.; Guidi, V.; Martinelli, G.; Ronconi, F.; (950) Sheng, J.; Yoshida, N.; Karasawa, J.; Fukami, T. Sens. Actuators, B
Richard, O.; Van, Dyck, D.; Van Landuyt, J. Sens. Actuators, B 2000, 1997, B41, 131.
B68, 140. (951) Shimizu, Y. Chem. Sens. 2002, 18, 42.
(931) Ferroni, M.; Carotta, M. C.; Guidi, V.; Martinelli, G.; Ronconi, F.; (952) Shimizu, Y.; Kuwano, N.; Hyodo, T.; Egashira, M. Sens. Actuators,
Sacerdoti, M.; Traversa, E. Sens. Actuators, B 2001, B77, 163. B 2002, B83, 195.
(932) Gao, L.; Li, Q.; Song, Z.; Wang, J. Sens. Actuators, B 2000, B71, (953) Sotter, E.; Vilanova, X.; Llobet, E.; Stankova, M.; Correig, X. J.
179. Optoelectron. AdV. Mater. 2005, 7, 1395.
(933) Garzella, C.; Comini, E.; Tempesti, E.; Frigeri, C.; Sberveglieri, G. (954) Takeuchi, T. Sens. Actuators 1988, 14, 109.
Sens. Actuators, B 2000, B68, 189. (955) Tan, O. K.; Cao, W.; Zhu, W.; Chai, J. W.; Pan, J. S. Sens. Actuators,
(934) Garzella, C.; Bontempi, E.; Depero, L. E.; Vomiero, A.; la Mea, G.; B 2003, B93, 396.
Sberveglieri, G. Sens. Actuators, B 2003, B93, 495. (956) Tang, H.; Prasad, K.; Sanjines, R.; Levy, F. Sens. Actuators, B 1995,
(935) Hasegawa, S.; Sasaki, Y.; Matsuhara, S. Sens. Actuators, B 1993, B26, 71.
14, 509. (957) Trinchi, A.; Li, Y. X.; Wlodarski, W.; Kaciulis, S.; Pandolfi, L.;
(936) Islam, M. R.; Kumazawa, N.; Takeuchi, M. Sens. Actuators, B 1998, Viticoli, S.; Comini, E.; Sberveglieri, G. Sens. Actuators, B 2003,
B46, 114. B95, 145.
(937) Jun, Y. K.; Kim, H. S.; Lee, J. H.; Hong, S. H. Sens. Actuators, B (958) Wu, M. T.; Yao, X.; Yuan, Z. H.; Sun, H. T.; Wu, W. C.; Chen, Q.
2005, B107, 264. H.; Xu, G. Y. Sens. Actuators, B 1993, 14, 491.
(938) Kirner, U.; Schierbaum, K. D.; Goepel, W.; Leibold, B.; Nicoloso, (959) Xu, Y.; Yao, K.; Zhou, X.; Cao, Q. Sens. Actuators, B 1993, 14,
N.; Weppner, W.; Fischer, D.; Chu, W. F. Sens. Actuators, B 1990, 492.
B1, 103. (960) Yamada, Y.; Seno, Y.; Masuoka, Y.; Nakamura, T.; Yamashita, K.
(939) Kobayashi, H.; Kishimoto, K.; Nakato, Y. Surf. Sci. 1994, 306, 393. Sens. Actuators, B 2000, B66, 164.
(940) Kobayashi, H.; Kishimoto, K.; Nakato, Y.; Tsubomura, H. Sens. (961) Zheng, L.; Xu, M.; Xu, T. Sens. Actuators, B 2000, B66, 28.
Actuators, B 1993, 13, 125.
(941) Li, M.; Chen, Y. Sens. Actuators, B 1996, B32, 83. CR0500535

You might also like