Vous êtes sur la page 1sur 21

Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Contents lists available at SciVerse ScienceDirect

Journal of Photochemistry and Photobiology C:


Photochemistry Reviews
journal homepage: www.elsevier.com/locate/jphotochemrev

Invited review

TiO2 photocatalysis: Design and applications


Kazuya Nakata a,b, , Akira Fujishima a,b
a

Photocatalyst Group, Kanagawa Academy of Science and Technology, KSP Building East 412, 3-2-1 Sakado, Takatsu-ku, Kawasaki, Kanagawa 213-0012, Japan
Research Institute for Science and Technology, Energy and Environment Photocatalyst Research Division, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601,
Japan
b

a r t i c l e

i n f o

Article history:
Received 5 March 2012
Received in revised form 30 May 2012
Accepted 1 June 2012
Available online 9 June 2012
Keywords:
TiO2 photocatalysis
Structural design
Dimensionality
Wettability pattern
Printing

a b s t r a c t
TiO2 photocatalysis is widely used in a variety of applications and products in the environmental and
energy elds, including self-cleaning surfaces, air and water purication systems, sterilization, hydrogen evolution, and photoelectrochemical conversion. The development of new materials, however, is
strongly required to provide enhanced performances with respect to the photocatalytic properties and
to nd new uses for TiO2 photocatalysis. In this review, recent developments in the area of TiO2 photocatalysis research, in terms of new materials from a structural design perspective, have been summarized.
The dimensionality associated with the structure of a TiO2 material can affect its properties and functions, including its photocatalytic performance, and also more specically its surface area, adsorption,
reectance, adhesion, and carrier transportation properties. We provide a brief introduction to the current situation in TiO2 photocatalysis, and describe structurally controlled TiO2 photocatalysts which can
be classied into zero-, one-, two-, and three-dimensional structures. Furthermore, novel applications of
TiO2 surfaces for the fabrication of wettability patterns and for printing are discussed.
2012 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Structural design and properties of TiO2 photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
TiO2 spheres (zero-dimensional) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
TiO2 bers and tubes (one-dimensional) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
TiO2 nanosheets (two-dimensional) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
TiO2 interconnected architecture (three-dimensional) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
New applications of TiO2 photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Wettability patterning using TiO2 photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Photocatalytic offset printing plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author at: Kanagawa Academy of Science and Technology, 412 KSP East, 3-2-1 Sakado, Takatsu-ku, Kanagawa 213-0012, Japan.
Tel.: +81 44 819 2040; fax: +81 44 819 2070.
E-mail address: pg-nakata@newkast.or.jp (K. Nakata).
1389-5567/$20.00 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jphotochemrev.2012.06.001

170
172
172
174
176
177
179
179
182
185
185
185

170

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Kazuya Nakata was born in 1977 in Sapporo, Japan. He


received his BSc (2000) from Shizuoka University, and
his MSc (2002) and PhD (2005) in science from Tokyo
Metropolitan University under the supervision of Professor Masahiro Yamashita. In 2005, he joined Tohoku
University as a research fellow of the Japan Society for
the Promotion of Science (JSPS), and then joined the
Massachusetts Institute of Technology in 2006 as a JSPS
research fellow. He has been a full-time researcher in the
photocatalyst group at the Kanagawa Academy of Science and Technology (KAST) since December 2007. In
September 2010, he also joined in the Organic Solar Cell
Assessment Project at KAST. He has also been a visiting
associate professor at the Tokyo University of Science since January 2011. In 2012,
he received a Sano award for young scientists from the electrochemical society of
Japan.
Akira Fujishima was born in 1942 in Tokyo, Japan. He
received his BSc (1966) from the Yokohama National University, and his MSc (1968) and PhD (1971) in engineering
from The University of Tokyo. He became a lecturer at
Kanagawa University in 1971 and then a lecturer at The
University of Tokyo in 1975. After serving as an associate professor (1978) and professor (1986), he became
a professor at The University of Tokyo Graduate School of
Engineering in 1995. He was appointed as the Chairman of
the Kanagawa Academy of Science and Technology (KAST)
and Director of the Functional Materials Research Laboratory of the Central Japan Railway Company in 2003. He
was appointed as professor emeritus of The University of
Tokyo and later became a special university professor emeritus of The University of
Tokyo in 2005. He served as the chairman for the Chemical Society of Japan from
2006 to 2007 and has been the Director of the China Research Center at the Japanese
Science and Technology Agency (JST) since 2008. He has been the President of Tokyo
University of Science since 2010. He received the Asahi Prize in 1983, the Chemical
Society of Japan Award in 2000, the Purple Ribbon Medal (Shijuhosho) in 2003, and
in 2004 he received the Japan Prize and the Japan Academy Prize and was named an
Honorable Citizen of Kawasaki City. He also received the Imperial Invention Award
and Kanagawa Culture Award in 2006. He received a Cultural Contributor in 2010.

1. Introduction
The development of photocatalysis has been the focus of considerable attention in recent years with photocatalysis being used in a
variety of products across a broad range of research areas, including
especially environmental and energy-related elds (Fig. 1) [14].
Following on from the water splitting breakthrough reported by
Fujishima and Honda in 1972 [5], the photocatalytic properties
of certain materials have been used to convert solar energy into

Fig. 1. Applications of TiO2 photocatalysis.

Fig. 2. Schematic illustration of the formation of photogenerated charge carriers


(hole and electron) upon absorption of ultraviolet (UV) light.

chemical energy to oxidize or reduce materials to obtain useful


materials including hydrogen [58] and hydrocarbons [9], and to
remove pollutants and bacteria [1018] on wall surfaces and in
air and water [14,1934]. Of the many different photocatalysts,
TiO2 has been the most widely studied and used in many applications because of its strong oxidizing abilities [23,3539] for the
decomposition of organic pollutants [25,26], superhydrophilicity
[40], chemical stability, long durability, nontoxicity, low cost, and
transparency to visible light.
The photocatalytic properties of TiO2 are derived from the
formation of photogenerated charge carriers (hole and electron) which occurs upon the absorption of ultraviolet (UV) light
corresponding to the band gap (Fig. 2) [1,3,19,41,42]. The photogenerated holes in the valence band diffuse to the TiO2 surface and react
with adsorbed water molecules, forming hydroxyl radicals ( OH)
(Fig. 3) [3]. The photogenerated holes and the hydroxyl radicals
oxidize nearby organic molecules on the TiO2 surface. Meanwhile,
electrons in the conduction band typically participate in reduction
processes, which are typically react with molecular oxygen in the
air to produce superoxide radical anions (O2 ).
In addition, TiO2 surfaces become superhydrophilic with a contact angle of less than 5 under UV-light irradiation (Fig. 4) [40].
The superhydrophilicity is originated from chemical conformation
changes of a surface [42]. The majority of the holes are subsequently
consumed by reacting directly with adsorbed organic species or
adsorbed water, producing OH radicals as described above. However, a small proportion of the holes is trapped at lattice oxygen
sites and may react with TiO2 itself, which weakens the bonds
between the lattice titanium and oxygen ions. Water molecules can
then interrupt these bonds, forming new hydroxyl groups (Fig. 5).
The singly coordinated OH groups produced by UV-light irradiation
are thermodynamically less stable and have high surface energy,
which leads to the formation of a superhydrophilic surface.
The construction of TiO2 nano- or micro-structures with
interesting morphologies and properties has recently attracted
considerable attention [43] and many TiO2 nanostructural materials, such as spheres [4465], nanorods [6674], bers [7589],
tubes [28,90108], sheets [109142], and interconnected architectures [143162], have been fabricated. Nanostructured TiO2
materials are widely used not only in photocatalysis, but also in
dye-sensitized solar cells (DSSCs) [163165], lithium-ion batteries
[166,167], and electrochromic displays [168].
It is well known that there are many factors which can exert
signicant inuence on photocatalytic performance, including the
size, specic surface area, pore volume, pore structure, crystalline
phase, and the exposed surface facets. Thus, the development of
performance improvements by adjusting these factors remains the
focus of photocatalysis research. Structural dimensionality is also a

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

171

Fig. 3. Processes occurring on bare TiO2 particles after UV excitation [3].

Fig. 4. (a) Hydrophobic surface before ultraviolet irradiation. (b) A highly


hydrophilic surface formed under ultraviolet irradiation. (c) Exposure of a hydrophobic TiO2 -coated glass to water vapor. The formation of fog (small water droplets)
hindered the view of the text on paper placed behind the glass. (d) Creation of an
antifogging surface by ultraviolet irradiation. The high hydrophilicity of the surface
prevents the formation of water droplets, making the text clearly visible [184].

of organic pollutants [55]. One-dimensional bers or tubes have


advantages with regard to less recombination because of the short
distance for charge carrier diffusion [28], light-scattering properties [169], and fabrication of self-standing nonwoven mats [170].
Zero- and one-dimensional structures have been well developed
and will be discussed in greater detail in the following sections.
Two-dimensional nanosheets have smooth surfaces and high adhesion [118,131], whereas three-dimensional monoliths may have
high carrier mobility as a result of their interconnecting structure and be used in environmental decontamination. Choosing TiO2
materials with the appropriate dimensionalities enables us to take
full advantage of the unique properties offered by TiO2 materials.
In this review, recent research in the eld of TiO2 photocatalysis
has been reviewed from the perspective of both structural design
and novel applications. We will initially introduce TiO2 photocatalysts possessing spheres as a zero-dimensional structure, bers
and tubes as one-dimensional structures, nanosheets as a twodimensional structures, and interconnected architectures as three
dimensional structures. We will then proceed to discuss the fabrication of wettability patterns and their application for the offset
printing plate as a new application of TiO2 photocatalysis. Finally, a
conclusion of the reviewed research will be provided together with
a brief future perspective.

factor which can affect the photocatalytic performance and also has
a signicant impact on the properties of TiO2 materials (Fig. 6). For
example, a sphere with zero dimensionality has a high specic surface area, resulting in a higher rate of photocatalytic decomposition

Fig. 5. Mechanism of photo-induced hydrophilicity [1].

Fig. 6. Schematic illustration of structural dimensionality of materials with


expected properties.

172

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 7. SEM and TEM (inset) images of TiO2 microspheres (a and b) and the as-prepared hierarchical TiO2 microspheres (c and d) and after calcination at 400 (e and f) and
600 C (g and h) [64].

2. Structural design and properties of TiO2 photocatalysts


2.1. TiO2 spheres (zero-dimensional)
Nano- or micro-structured TiO2 spheres are the most widely
studied and used in TiO2 -related materials. Useful and interesting
properties derived from their unique structures have been reported
in a great many publications [51,52,56,57,59,60,6365]. These TiO2
spheres usually possess a high specic surface area and a high pore
volume and pore size, with these properties increasing the size of
the accessible surface area and the rate of mass transfer for organic

pollutant adsorption. Overall, these increases result in better photocatalytic performance because the photocatalytic reactions are
based on chemical reactions on the surface of the photocatalyst
[52,56,57,6365]. Furthermore, these structural features increase
the light-harvesting capabilities of these materials because they
enhance light use by allowing as much light as possible to access
the interior [51,52]. Their light-harvesting capability makes them
good candidates for use not only in photocatalysis, but also in DSSCs
[51].
Spheres of this type are typically prepared from a titanium alkoxide such as titanium tetraisopropoxide or titanium

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

173

Fig. 8. (a) Multiple diffractions and reections on the hollow spheres; SEM image of (b) the PS particles, (c) the titania-coated PS particles and (d) the TiO2 hollow spheres;
(e) TEM image of the TiO2 hollow spheres [51].

tetrabutoxide, in the presence of a polymer to provide a porous


structure and with or without the addition of an acid to accelerate the reaction [55]. The TiO2 spheres obtained can be further
treated according to the hydrothermal methods to produce porous
or hierarchical structures. Zheng et al. reported the synthesis of
hierarchical TiO2 microspheres covered with nanotubular and dendritic structures, with diameters in the range of 46 mm (Fig. 7)
[64]. They prepared TiO2 spheres from titanium tetrabutoxide
and sulfuric acid, obtaining the TiO2 spheres as a white powder.
The resulting TiO2 spheres were then further treated according to
the hydrothermal method in an aqueous NaOH solution at 150 C
for 24 h. The TiO2 spheres obtained following periods of washing
and annealing had different structures, with the structure being

dependent on the annealing temperature. Annealing at high temperatures decreased the specic surface area of the TiO2 spheres.
The TiO2 spheres formed at higher annealing temperatures exhibited superior photocatalytic activities and the TiO2 spheres formed
following an annealing process at 400 C in particular, showed
the highest photocatalytic activity for decomposition of organic
molecules because of their high specic surface area and highly
crystalline form. The formation of hollow structures represents
another strategy for obtaining high photocatalytic performance
[52,54,58,61] because a hollow structure not only has a high specic
surface area but also enables multiple diffractions and reections
of light (Fig. 8) [51]. Kondo et al. reported the preparation of TiO2
hollow spheres by the hydrolysis of titanium tetraisopropoxide in

174

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 10. Comparison of photocatalytic activities of the TiO2 spheres with solid,
sphere-in-sphere, and hollow structure. Inset shows a schematic illustration of
multireections within the sphere-in-sphere structure [52].

Fig. 9. Schematic illustration of the fabrication procedures for hollow TiO2 shells
(polymer sphere, PEI, and TiO2 nanosheets are indicated) [61].

the presence of cationic polystyrene (PS) spheres, followed by calcination [51]. During the synthetic process, the cationic PS spheres,
which were used as a template, were initially coated with the titanium species. Following the calcination process, the PS spheres
were removed and the titanium species converted to TiO2 which
resulted in the formation of the TiO2 hollow spheres. The resulting
TiO2 hollow spheres possessed a large specic surface area and the
potential for multiple diffractions and reections of light, providing
the sphere with properties that are advantageous for photocatalytic
decomposition. These advantages were clearly demonstrated by
the superior photocatalytic decomposition rate of 2-propanol using
the TiO2 hollow spheres compared with the rate achieved using
commercially available TiO2 particles (P25). A similar method of
preparation has also been reported involving the use of titanium
nanosheets coated onto PS or poly(methyl methacrylate) spheres,
followed by calcination with UV irradiation (Fig. 9). This process
produced ultrathin TiO2 shells which were derived from the TiO2
nanosheet, with the thickness of the TiO2 shell reported as being
approximately 5 nm [61]. Other methods of preparation have also
been developed, including the use of soft templates of O2 bubbles
[54] and template-free processes [52]. Li et al. reported the preparation of hollow TiO2 spheres with a sphere-in-sphere structure
using a template-free process and the resulting spheres were found
to have high photocatalytic activity, likely because of multiple scattering and reections of light within the TiO2 spheres would extend
the light path length (Fig. 10) [52].
Anatase TiO2 crystals with exposed high-energy (001) facets
have high photocatalytic activity. With this in mind, it was envisaged that TiO2 spheres containing such facets would exhibit greatly
enhanced levels of photocatalytic activity (Fig. 11) [53,56]. Hierarchical [44,47,49,63,65] and hollow [57] structures with (001)
facets were typically prepared according to the hydro- or solvothermal methods using titanium alkoxides. The introduction of
both a hierarchical structure and high-energy (001) facets may
provide high photocatalytic activity. Liu et al. reported the preparation of mesoporous coreshell spheres composed of small TiO2
nanocrystals with exposed step-like (001) and (010) facets using
a combination of electrospray and hydrothermal treatment methods [55]. The electrospray technique uses a high-voltage electric

eld to obtain micro- or nano-sized spheres and in this particular


case, the coreshell TiO2 spheres were produced by electrospray
using titanium alkoxide and polyvinylpyrrolidone (PVP). Subsequent hydrothermal treatment provided a hierarchical structure
with tunable pore size, pore volume, specic surface area, and
percentage of specied crystal facets, because the PVP formed a
network and the amorphous TiO2 particles lled the pores within
the network after the electrospray of the coreshell TiO2 spheres.
The PVP was subsequently removed from the TiO2 spheres (Fig. 12)
and the resulting TiO2 was crystallized during the hydrothermal
process (Fig. 13). The crystal morphology depended on several conditions including the amount of PVP used and the conditions of the
hydrothermal method.
2.2. TiO2 bers and tubes (one-dimensional)
TiO2 materials with one-dimensional structures, such as bers
and tubes, possess unique properties and advantages for photocatalytic reactions. In bers and tubes, the higher surface-to-volume
ratio enables a reduction in the holeelectron recombination rate
and a high interfacial charge carrier transfer rate, with both of
these effects being favorable for photocatalytic reactions [171]. Furthermore, self-standing nonwoven mats can be obtained from the
assembly of bers, representing an application that is only available to one-dimensional materials. TiO2 bers are already used in
an extensive range of applications, including photocatalysis, gas
sensing [172174], DSSCs [175178] and batteries [25,179]. There
are several methods available for preparing TiO2 bers and a useful tool in the fabrication of TiO2 bers is electrospinning, which
is a simple and effective method of producing nanobers through
a process using a high-voltage electric eld [180], with the setup
being very similar to that used in the electrospray method (Fig. 14).
The morphologies of the bers can be tailored by adjusting the
process parameters, including the solution concentration, molecular weight of the polymer, choice of solvent, applied electric-eld
strength, and deposition distance. Nonwoven mats assembled from
bers possess properties such as low resistance to mass transport
and large volumetric surface area because of their open structure
with the pores between the bers.
TiO2 bers are typically prepared from the electrospinning of a
mixture of titanium alkoxide and a polymer, such as PVP, with a
subsequent calcination step employed to remove the polymer and
crystallize the TiO2 [80]. The average diameter of the TiO2 bers can
be controlled by varying a number of processing parameters such
as the ratio of polymer to titanium tetraisopropoxide, the strength
of the electric eld, and the feeding rate of the precursor in the

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

175

Fig. 11. (a) TEM image of the ower-like TiO2 nanostructures. (b) Schematic diagram of the ower-like TiO2 nanostructures. (c) HRTEM image of a truncated tetragonal
pyramidal TiO2 nanocrystal. (d) The corresponding fast-Fourier transform (FFT) pattern [56].

electrospinning process. Electrospun TiO2 bers generally have


large surface area-to-volume ratios and are therefore expected to
provide efcient photocatalytic performances for the degradation
of organic molecules such as methylene blue and methyl orange
[181]. We would like to point out that brous photocatalysts are
attractive candidates for practical applications in environmental
cleanup. In a practical photocatalytic process, the separation of
particle-type photocatalysts from the solution following the reaction can be very difcult and the tendency of the particles to
agglomerate into larger particles effectively reduces the photocatalytic activity during the cycled use. In contrast, the bers can be
used as a mat, eliminating the problems associated with separation.
The introduction of a porous structure to increase the specic
surface area represents one way of improving the photocatalytic
performance of TiO2 . A co-jetting method has been reported for the
construction of hollow TiO2 bers [76,81,82,87,89]. In the co-jetting
process, a multi-channel nozzle is used and a solution containing
the TiO2 precursor and the polymer is sprayed from one nozzle,
whereas an oil is sprayed from another nozzle (Fig. 15). Following the co-jet electrospinning process, multi-component bers are
obtained composed of amorphous TiO2 on the outside of the ber
and oil on the inside of the ber. Subsequent calcinations effectively
crystallize the TiO2 and remove the oil from the ber, resulting in
formation of hollow TiO2 bers. Zhao et al. reported that by using
a multi-channel nozzle, multi-channel hollow TiO2 bers could
be obtained and they reported the construction of multi-channel

hollow TiO2 bers with zero to three channels (Fig. 16) [182].
The photocatalytic performance improved as the number of channels increased because the specic surface area increased with
the increasing numbers of channels, and multiple reections of
incident light could be expected with an increased number of
channels.
Anodization is a powerful tool for the fabrication of TiO2 tubes.
Anodization is typically conducted in an electrolyte containing a
hydrogen uoride (HF)-based aqueous solution. A titanium foil
electrode and counter electrode are soaked in the electrolyte, with
the two being connected to a power supply to apply a constant
voltage (Fig. 17) [28,90108,183]. Following the anodization process, TiO2 nanotube arrays are formed on the foil surface and
have straight channels against the foil. This structural characteristic provides superior photocatalytic decomposition of organic
pollutants because it enables the facile diffusion of organic pollutants into the TiO2 nanotubes. Thins walls are another structural
characteristic of TiO2 nanotubes and these structural features
reduce the recombination of holes and electrons generated by
photo-absorption because the half-thickness of the nanotube wall
is signicantly less than the carrier diffusion length in TiO2 .
The TiO2 nanotube arrays consequently have the potential to
achieve higher photocatalytic activity for the decomposition of
organic pollutants. In the case of the TiO2 nanotubes prepared
according to the procedure using an HF-based aqueous solution,
the length of the TiO2 nanotube was typically several hundred

176

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 12. Cross-sectional SEM images of cut TiO2 spheres obtained after hydrothermal treatment with (A) 0 and (B) 2.0 mL of acetic acid; (C) shows an FE-SEM image of the
sample obtained after hydrothermal treatment with 2.0 mL of acetic acid [55].

nanometers (Fig. 18) [28,98]. The morphology and especially


the length of the nanotubes, affects the photocatalytic performance of TiO2 nanotube arrays. Increasing the length of the
TiO2 nanotubes therefore improves the photocatalytic performance because of the increased specic surface area. Liu et al.
prepared 17 m TiO2 nanotube arrays by a combination of extending the anodization time and the use of a mixed electrolyte
containing ammonium uoride, formamide, and water (Fig. 19)
[92]. An improvement in the photocatalytic performance for the
decomposition of organic molecules (acetaldehyde in this case)
was observed which was attributed to the larger surface area.
They also demonstrated that the long TiO2 nanotubes exhibited
a greater photocatalytic performance than that of a P25 nanoparticulate lm of almost the same thickness. The highly ordered
TiO2 nanotube array contained vertically straight channels with
open mouths against the substrate that allowed easy access of the
acetaldehyde to the TiO2 surface. In contrast, the P25 TiO2 nanoparticulate lm has a disordered structure as a result of randomly
packed TiO2 particles, which suppressed the diffusion of organic
molecules.
2.3. TiO2 nanosheets (two-dimensional)
A nanosheet is a nanosized ake-shaped material with a at
surface and high aspect ratio. Furthermore, nanosheets have
an extremely small thickness of 110 nm and a lateral size

ranging from the submicrometer level to several tens of micrometers. The shape results in low turbidity, excellent adhesion
to substrates, and high smoothness [131]. Nanosheets show
photocatalytic properties including the photocatalytic decomposition of organic molecules and superhydrophilicity, under
UV irradiation [130]. The combination of their photocatalytic
properties and highly smooth surface has led to the potential application of nanosheet lms as self-cleaning coatings
[131].
TiO2 nanosheets are typically prepared according to an alkaline
hydrothermal process using TiO2 powder as a precursor [131] or
from protonic titanate hydrates [130], followed by either a calcination process or hydro/solvothermal reactions.
The photocatalytic decomposition of organic molecules by TiO2
nanosheets has been investigated. Shibata et al. reported the
fabrication of TiO2 nanostructured lms from TiO2 nanosheets
and went on to describe their photocatalytic performance for
the decomposition of 2-propanol and methylene blue dye [130].
Slow photocatalytic decomposition of 2-propanol was observed
under UV irradiation, with CO2 being generated as a byproduct. Given that the valence band edge of the TiO2 nanosheet is
located 0.5 eV deeper than that of bulk anatase TiO2 [126], it
was envisaged that the TiO2 nanosheet would have enough oxidizing power to decompose 2-propanol. Electronhole pairs are
generally generated in photocatalysts under irradiation with light
with a photon energy corresponding to the band gap, and they

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

177

Fig. 13. High-magnication TEM images of the TiO2 nanocrystals in the TiO2 spheres after hydrothermal treatment with (A) 0 and (B) 2.0 mL of acetic acid. (C) Schematic
illustration of crystal facets for the TiO2 nanocrystals [55].

subsequently diffuse from inside the bulk to the surface. Given


that each TiO2 nanosheet is electronically insulated, however,
very few of the electronhole pairs produced inside an assembly of TiO2 nanosheets can diffuse to the surface and the TiO2
nanosheet consequently provides only slow decomposition of 2propanol.
One possible way to improve the photocatalytic performance
for the decomposition of organic molecules is to use high-energy
facets as described in the section of zero-dimensional TiO2 materials. Han et al. prepared TiO2 nanosheets (thickness 8 nm) with
(001) facets and examined their photocatalytic performance in the
decomposition of methyl orange (Fig. 20) [115]. In their experiments, commercially available P25 TiO2 nanoparticles were used
as a reference. The TiO2 nanosheets with (001) facets enabled the
decomposition of methyl orange at a higher rate than that achieved
using P25 TiO2 particles.
TiO2 nanosheets also exhibit photo-induced hydrophilicity
under UV irradiation. As stated above, a TiO2 nanosheet not only
possesses photocatalytic properties, but also has low turbidity,
excellent adhesion, and high smoothness and one potential application of TiO2 nanosheets is consequently in self-cleaning surfaces.
Shichi et al. reported the production of a self-cleaning glass using
TiO2 nanosheets [131]. The fabrication procedure typically involved
the application of a dense blocking layer such as SiO2 onto sodalime glass to avoid diffusion of Na+ ions from the glass during the
heating process, because the presence of Na+ ions greatly reduced
the photocatalytic activity. The glass was then coated with a TiO2
nanosheet, followed by a calcinations process. The TiO2 nanosheet
had low turbidity, strong adhesion to glass, and good hardness, as
demonstrated by the pencil test, which is generally evaluated by a
lm hardness tester using pencils [118]. Furthermore, following

the dip-coating of the self-cleaning glass in a solution containing methylene blue, barely any methylene blue was observed to
be attached to the self-cleaning glass, indicating high antifouling properties. In contrast to this result, glass coated with TiO2
nanoparticles experienced considerable attachment of methylene
blue. The difference between these results can be attributed to
the very smooth surface of the self-cleaning glass prepared using
a TiO2 nanosheet, which reduces the attachment of methylene
blue molecules, whereas the glass prepared using TiO2 particles
has a relatively rough surface, which promotes attachment of the
methylene blue molecules. Materials with this combination of low
adhesion and photocatalytic properties (oxidative decomposition
of organic molecules and photo-induced hydrophilicity) are good
candidates for application as novel self-cleaning coatings.
2.4. TiO2 interconnected architecture (three-dimensional)
Of the various morphological structures, three-dimensional
interconnected structures have the potential for producing a new
class of materials with novel applications. Structures of this type are
important for practical applications because the three-dimensional
hierarchical structures with pores have potentially large surfaceto-volume ratios, which provide a signicant advantage in the form
of efcient diffusion pathways for guest species, such as organic
pollutants, into the framework and ultimately support efcient
purication, separation, and storage. Furthermore, an interconnected structure is potentially superior not only for carrier mobility,
but also from a practical point of view. For example, at present,
almost all photocatalytic puriers use TiO2 particles coated onto
porous structured ceramics. The TiO2 particles may be stripped
from the ceramic, however, leading to the release of small dust

178

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 14. Schematic illustration of the setup for electrospinning.

particles and degradation of the photocatalytic properties. To avoid


this stripping process, an interlayer such as SiO2 is typically placed
between the TiO2 particles and a substrate is used to bind the TiO2
particles.
Konishi et al. prepared macroporous TiO2 monoliths for use
in chromatographic applications (Fig. 21) [148,149]. The materials
were prepared according to a solgel process using poly(ethylene

oxide) as a phase separator and N-methylformamide as a proton


scavenger. Heat treatment allowed for the formation of an interconnected TiO2 framework with mesoporous structures. These TiO2
monoliths showed excellent chromatographic separation properties for the purication of phosphorus-containing compounds.
Yu et al. reported sponge-like macro/mesoporous TiO2 prepared
from titanium tetrabutoxide, followed by hydrothermal treatment,
without the use of templates or additives [162]. The structural
parameters of the material, such as pore size, pore volume and specic surface area, and crystallinity, could be controlled by adjusting
the hydrothermal reaction time. Furthermore, the photocatalytic
performance of the material for the decomposition of acetone was
superior to that of P25.
The use of photonic crystal structures, which are inverse opals
that have a spatially periodic structure with lattice parameters
of the order of the wavelength of light, represents one way for
improving the photocatalytic performance. These photonic crystals possess high surface areas and photonic band gaps that allow
for control of spontaneous emission and localization of photons. A number of researchers have produced periodic structures
of TiO2 photonic crystals (Fig. 22) [152,154,158,159], with air
spheres lling the uniform voids, using colloidal templates. The
use of TiO2 photonic crystals for the photocatalytic degradation of
1,2-dichlorobenzene has been demonstrated [154]. The research
involved the preparation of a thin TiO2 lm of photonic structure by a solgel technique, using PS spheres as a template, with
alternating air and TiO2 spaces with a periodicity of approximately
150 nm. The band gap of anatase TiO2 is 3.2 eV and so a photonic
band gap in the mid-to-high UV wavelength range (280380 nm)
was required to increase the quantum efciency of the photocatalyst. The photocatalytic performance of the TiO2 photonic
crystal was examined by measuring the decomposition of 1,2dichlorobenzene under UV irradiation, with the results indicating a
higher photonic efciency than that of the commercially available
P25.

Fig. 15. (A) Schematic illustration of the setup for electrospinning nanobers having a core/sheath structure. (B) TEM image of two as-spun hollow bers after the oily cores
had been extracted with octane. (C) TEM image of TiO2 hollow bers that were obtained by calcining the composite nanotubes in air at 500 C. (D) SEM image of a uniaxially
aligned array of TiO2 hollow bers that were collected across the gap between a pair of electrodes [81].

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

179

Fig. 16. SEM images of TiO2 bers with (a) 0, (b) 1, (c) 2 and (d) 3 channel [89].

3. New applications of TiO2 photocatalysts


3.1. Wettability patterning using TiO2 photocatalysts
Multiple opportunities still exist for the practical application
of TiO2 photocatalysis to a variety of different elds, especially in

Fig. 17. Schematic illustration of the setup for anodization of Ti.

the environmental and energy elds. TiO2 surfaces exhibit a strong


oxidizing ability for the decomposition of organic molecules, and
superhydrophilicity [1,3,19,21,184,185], and these properties can
be used for generating wettability patterns. Wettability patterns
have been used in many elds such as offset printing [186188] and
in printed-circuit boards, and have potential applications for uid
microchips [189191,192]. For these applications, the wettability
patterns are typically fabricated according to a photolithographic
technique using a photosensitive polymer and light illumination
[193,194], to provide the desired hydrophobic(super)hydrophilic
wettability patterns. In this process, a photomask is typically used
to transfer its geometric pattern to a photosensitive polymer on the
substrate using light. The illuminated area on the photosensitive
polymer is chemically changed and alkaline chemical treatments
then engrave the desired exposure pattern onto the substrate
underneath the photosensitive polymer.
TiO2 materials with at or rough surfaces can become
hydrophobic or superhydrophobic, respectively, when coated
with hydrophobic compounds such as octadodecylphosphonic
acid (ODP), octadecyltrimethoxysilane (ODS), or uoroalkylsilane [26,195198]. These surfaces can be changed to
become superhydrophilic through the decomposition of the

180

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 18. Cross-sectional and top view images of TiO2 nanotube arrays prepared in different electrolytes of (a) 0.5 vol % HF, (b) 0.5 vol % HF and 1 M H3 PO4 and (c) 0.5 vol% HF
and 4 M H3 PO4 [28].

Fig. 19. Cross-sectional and top view images of TiO2 nanotube arrays prepared by anodizing Ti foil at 20 V in 0.5 wt% HF aqueous solution for 20 min (a, b) and in formamidebased electrolyte for 6 h (ce) [92].

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

181

Fig. 20. (A) XRD pattern and (B) low-magnication TEM image of TiO2 nanosheets. (C) High-magnication TEM image of an individual TiO2 nanosheet; the inset shows the
corresponding SAED pattern. (D) High resolution TEM image from the vertical nanosheets [115].

hydrophobic compounds, as a result of their photocatalytic


decomposition and superhydrophilicity under UV irradiation. A
superhydrophobicsuperhydrophilic pattern can therefore be
prepared by selective light illumination through a patterned photomask [195198]. Increases to the surface roughness can be used
to reduce the time required for the photocatalytic decomposition
of hydrophobic compounds. For example, Zhang et al. reported
that the formation of a rough TiO2 surface which was prepared
using CF4 plasma [195]. TiO2 nanotubes can also be used because of
its high surface roughness and high performance for the oxidative
decomposition for organic compounds [199].
A new process for producing superhydrophobicsuperhydrophilic
patterns on TiO2 surfaces has been reported [200] in which the
wettability patterns were fabricated by coating a self-assembled
monolayer (SAM) of a hydrophobic compound, namely ODS,
and selectively irradiating with UV light to decompose the ODS
molecules (Fig. 23). In this case, an ink-jet printer was used to
deposit a water-based ink pattern as a photoresist. Following UV
irradiation, the SAM on the area that was not protected by the
water-based ink was photocatalytically decomposed becoming
superhydrophilic, and the SAM covered by the water-based ink
remained superhydrophobic, even after UV irradiation, leading to
the formation of a superhydrophobicsuperhydrophilic pattern. It
is worthy of note that to obtain a superhydrophobic surface after
coating with the SAM, the titanium substrate was initially oxidized
to rutile TiO2 and Ti2 O by calcination at high temperature, with

the substrate being subsequently cooled to room temperature.


Interestingly, a fragile white layer was generated which could be
separated from the substrate spontaneously after cooling, with the
remaining substrate possessing a relatively rough surface, which
was suitable for forming a superhydrophobic surface following
SAM coating (Fig. 24). The superhydrophobicsuperhydrophilic
pattern could be erased by irradiation with UV light of the entire
surface by photocatalytic decomposition of the SAM (Fig. 25). A
different superhydrophobicsuperhydrophilic pattern could then
be formed using the above process. This process offers a renewable,
resource-saving, and environmentally friendly method for the
formation of wettability patterns.
Wettability pattern formation without the use of SAMs has
also been reported [201] in which composite lms of TiO2
and polydimethylsiloxane (PDMS) were prepared by a solgel
method, cured with UV irradiation, and then treated in hot
water to crystallize the TiO2 in the lm. The surface of the
TiO2 PDMS composite was initially hydrophobic, but became
superhydrophilic following UV irradiation. The presence of
anatase TiO2 contributed to the photoinduced superhydrophilicity. Hydrophobicsuperhydrophilic patterns could therefore be
obtained by selective UV illumination through a photomask
(Fig. 26). The wettability patterns could be erased by UV irradiation and thermal treatment, and new wettability patterns could
be constructed. This TiO2 PDMS composite exhibited rewritable
wettability without the need for organic chemicals.

182

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 21. (ae) SEM images of dried TiO2 gels prepared with varied molar ratios of water/TiO2 . (f) Digital picture of monolithic TiO2 gels prepared in Teon tubes and a coin.
(g) Schematic illustration of coarsening of phase-separated domains. Bicontinuous structure is obtained by inducing the phase separation in parallel to the gelation [148].

3.2. Photocatalytic offset printing plates


Offset printing, which is a method that is widely used to print
newspapers, is a good candidate for the application of wettability
patterns. The technique of offset printing requires a master printing plate with a hydrophobichydrophilic wettability pattern on an

Fig. 22. SEM image of photonic crystals of air spheres in TiO2 [158].

anodized aluminum substrate, which is called a PS plate. A wettability pattern of a PS plate is prepared by a photolithography technique
using photosensitive polymer, illuminated through a photomask
and subsequently developed to remove the illuminated (or protected) area of the photosensitive polymer [187]. It is worthy of
note that the surfaces of the photosensitive polymer are hydrophobic and the exposed areas of the anodized aluminum surface are
hydrophilic (in the case of a positive photoresist). In the printing
process, the PS plate is rst moistened with water, which is selectively attracted to the hydrophilic area and an oil-based printing
ink is then selectively attached onto the hydrophobic area because
the presence of water on the hydrophilic areas prevents the adhesion of the oil-based ink to these areas. As a result, an ink pattern
is formed on the PS plate which is transferred to a rubber blanket
and subsequently transferred onto the paper.
From a perspective of saving resources, and reducing the
nancial and environmental impacts, there are some drawbacks
associated with this process, which needs to be overcome. Firstly,
a photomask, which is generally made using a plastic substrate,
has to be prepared prior to printing and subsequently discarded
following the printing process. Alkaline chemical wastes are also
generated during the development process. Furthermore, the plate
life is short because the printing plate is generally made using an
aluminum substrate with a soft surface, whose surface can be galled
during the printing process. Lastly, the plate has to be disposed of
following the printing process.
One solution to overcome the problems described above is to use
wettability patterns prepared by TiO2 photocatalysis as described

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

183

Fig. 23. Schematic representation of the wettability pattern fabrication procedure [200].

Fig. 24. Photographs of the delamination of the white layer from the remaining substrate after sintering at 1000 C [200].

Fig. 25. (a) Photographs of the water-based ink patterns and (b) the wettability patterns after the deposition of water in the rst process. (c) Photographs of the water-based
ink patterns and (d) the wettability patterns after the deposition of water in the secondary process. Scale bar is 5 mm. [200].

184

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

Fig. 27. Schematic diagram of patterning and reusing processes [187].

Fig. 26. (a) Schematic representation of the preparation procedure of the


wettability patterns on the TiO2 /PDMS composite lm, and photographs of
hydrophobicsuperhydrophilic patterns formed by UV irradiation through photomasks after (b) rst and (c) secondary processing [201].

above. In practice, a superhydrophilichydrophobic pattern on a


TiO2 -coated aluminum substrate can be prepared and used for
offset printing [188]. On this substrate, the wettability patterns
consisted of a hydrophobic area based on organic compounds that
could be eliminated by photocatalytic decomposition using TiO2
under UV irradiation and a superhydrophilic area based on TiO2 .
Suda et al. reported that printing paper was successively obtained
when using the above substrate [188]. Interestingly, the wettability patterns could be cleaned by UV irradiation as a consequence of
the elimination of organic compounds by TiO2 photocatalysis, and

Fig. 28. Photographs of posters (309 mm 570 mm) printed using the superhydrophobicsuperhydrophilic patterns formed on the pristine plate (left) and reused plate
(right) [187].

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

repeatedly resurfaced with new superhydrophilichydrophobic


patterns. This renewal method of wettability patterns should lead
to a reduction in the number of printing plates being disposed of.
The ink-jet technique provides water-based ink patterns
as a photomask for offset printing. Nishimoto et al. reported
the fabrication of superhydrophobicsuperhydrophilic patterns
on a TiO2 surface coated with organic compounds (octadodecylphosphonic acid) and the subsequent application of the
wettability patterns in offset printing [187]. They used a waterbased ink containing organic complex that absorbed UV-light.
Selective UV-light irradiation through a photomask made of waterbased ink patterns created using the ink-jet technique formed
superhydrophobicsuperhydrophilic patterns in combination with
TiO2 photocatalysis for the decomposition of organic compounds
(Fig. 27). Furthermore, they used a SAM in this process for a making
of hydrophobic area instead of photoresist. This eliminates scope
for the development of a process that uses a strong alkaline solution, which could potentially reduce the environmental impact of
the process. The renewability of the substrate was demonstrated
by the elimination of the SAM and reconstruction of a wettability
pattern in not only black-white images, but also in color images
(Fig. 28).
An aluminum substrate can be galled during the printing process and the development of a substrate with strong durability is
therefore required. Nakata et al. reported that a titanium-based
offset printing plate can also be used as a renewable printing
plate [186]. They prepared superhydrophilicsuperhydrophobic
patterns based on the patterned SAM and TiO2 surface of a titanium
substrate. The TiO2 surface was easily prepared according to the
simple calcination of titanium substrate that offered strong adhesion of TiO2 with a titanium substrate. They successfully achieved
the printing of 1000 sheets on high-grade paper with resolutions of
133 lpi and 150 lpi. The titanium substrate based printing plate also
showed renewability as described above. The titanium substrate
was highly durable compared to the corresponding aluminum
substrates. Thus, it is envisaged that the combination of the renewability of the superhydrophilicsuperhydrophobic patterns and the
high durability of the titanium substrate would allow for a long
plate-life for offset printing.

4. Conclusions
An overview of recent signicant publications in the eld of
TiO2 photocatalysis, especially from the perspective of the design
and new applications of TiO2 materials, has been provided. The
review initially highlights the structural design of TiO2 materials
regarding their dimensional classication. In the zero-dimensional
structure of spheres, TiO2 materials have a high specic surface
area which and are typically produced according to the hydrothermal and/or electrospray methods to obtain porous structure. The
introduction of a hollow and high energy facet into the structure
should also enhance the photocatalytic performance. The onedimensional structures of ber or tubes may show lower levels
of recombination and can be obtained according to the electrospinning or anodization methods. The electrospinning method in
particular allows for the fabrication of self-standing nonwoven
mats which are useful in a number of different applications. As twodimensional structures, TiO2 nanosheets which have a at surface,
high aspect ratio, low turbidity, and excellent adhesion for substrates, can be effectively applied in self-cleaning glass. We also
discussed the three-dimensional architecture of monolithic TiO2
materials which have an interconnected structure. This unique
structure can provide potential advantages for purication, separation and storage. Overall, the properties of TiO2 materials are
strongly dependent on the dimensionality of their structures and

185

it is therefore very important that the dimensionality is considered


prior to the application of TiO2 materials.
Several novel applications of TiO2 materials involving the formation of superhydrophilic wettability patterns and their use
in offset printing have also been discussed. The combination of
oxidative decomposition ability for organic molecules and superhydrophilicity enables the formation of wettability patterns. The
wettability pattern made from TiO2 has the unique advantage of
being renewability compared to typical aluminum based systems.
In this particular case, a combination of TiO2 photocatalysis, coating of SAMs and inkjet patterning allows for the fabrication of a
superhydrophilic-superhydrophobic pattern. Patterns of this type
exhibit renewability through the decomposition of SAM by TiO2
photocatalysis. In a separate case, renewable wettability patterns
were exhibited by composite TiO2 -PDMS lms, which provided a
renewable wettability pattern without the use of a SAM. Lastly, the
application of wettability patterns for an offset printing method
was discussed. Wettability patterns made through TiO2 photocatalysis can be used for an offset printing plate and the plates
themselves exhibited renewability. Thus, this technology could
offer a renewable offset printing plate, which could allow for potential resource-saving.
TiO2 is the most widely studied photocatalyst and it is used
in numerous applications because of its compatibility with modern technology. New materials and applications involving TiO2 can
improve our lives in areas such as energy production and environmental protection.
Acknowledgments
This work was supported by a Grant-in-Aid for Scientic
Research (B) and for Challenging Exploratory Research (No.
21654043) from the Ministry of Education, Culture, Sports, Science
and Technology (MEXT), Japan, a Kurata Research Grant, and the
Nippon Sheet Glass Foundation for Materials Science and Engineering.
References
[1] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, J. Photochem. Photobiol. C 1 (2000) 121.
[2] A. Fujishima, T.N. Rao, D.A. Tryk, TiO2 photocatalysts and diamond electrodes,
Electrochim. Acta 45 (2000) 46834690.
[3] A. Fujishima, X. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface
phenomena, Surf. Sci. Rep. 63 (2008) 515582.
[4] K. Hashimoto, H. Irie, A. Fujishima, TiO2 photocatalysis: a historical overview
and future prospects, Jpn. J. Appl. Phys. 44 (2005) 82698285.
[5] A. Fujishima, K. Honda, Electrochemical photolysis of water at a semiconductor electrode, Nature 238 (1972) 3738.
[6] A. Kudo, Y. Miseki, Heterogeneous photocatalyst materials for water splitting,
Chem. Soc. Rev. 38 (2009) 253278.
[7] K. Maeda, Photocatalytic water splitting using semiconductor particles: history and recent developments, J. Photochem. Photobiol. C 12 (2011) 237268.
[8] A. Ryu, Recent progress on photocatalytic and photoelectrochemical water
splitting under visible light irradiation, J. Photochem. Photobiol. C 11 (2010)
179209.
[9] T. Inoue, A. Fujishima, S. Konishi, K. Honda, Photoelectrocatalytic reduction
of carbon dioxide in aqueous suspensions of semiconductor powders, Nature
277 (1979) 637638.
[10] L. Caballero, K.A. Whitehead, N.S. Allen, J. Verran, Inactivation of
Escherichia coli on immobilized TiO2 using uorescent light, J. Photochem.
Photobiol. A 202 (2009) 9298.
[11] R. Cai, K. Hashimoto, K. Itoh, Y. Kubota, A. Fujishima, Photokilling of malignant
cells with ultrane TiO2 powder, Bull. Chem. Soc. Japan 4 (1991) 12681273.
[12] T. Matsunaga, R. Tomoda, T. Nakajima, H. Wake, Photoelectrochemical sterilization of microbial cells by semiconductor powders, FEMS Microbiol. Lett.
29 (1985) 211214.
[13] C. McCullagh, J. Robertson, D. Bahnemann, P. Robertson, The application of TiO2 photocatalysis for disinfection of water contaminated with
pathogenic micro-organisms: a review, Res. Chem. Intermed. 33 (2007)
359375.
[14] J.R. Peller, R.L. Whitman, S. Grifth, P. Harris, C. Peller, J. Scalzitti, TiO2 as
a photocatalyst for control of the aquatic invasive alga, Cladophora, under
natural and articial light, J. Photochem. Photobiol. A 186 (2007) 212217.

186

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

[15] K. Sunada, Y. Kikuchi, K. Hashimoto, A. Fujishima, Bactericidal and detoxication effects of TiO2 thin lm photocatalysts, Environ. Sci. Technol. 32 (1998)
726728.
[16] K. Sunada, T. Watanabe, K. Hashimoto, Studies on photokilling of bacteria on
TiO2 thin lm, J. Photochem. Photobiol. A 156 (2003) 227233.
[17] K. Sunada, T. Watanabe, K. Hashimoto, Bactericidal activity of copperdeposited TiO2 thin lm under weak UV light illumination, Environ. Sci.
Technol. 37 (2003) 47854789.
[18] E.J. Wolfrum, J. Huang, D.M. Blake, P.-C. Maness, Z. Huang, J. Fiest, W.A. Jacoby,
Photocatalytic oxidation of bacteria, bacterial and fungal spores, and model
biolm components to carbon dioxide on titanium dioxide-coated surfaces,
Environ. Sci. Technol. 36 (2002) 34123419.
[19] A. Fujishima, K. Hashimoto, T. Watanabe, TiO2 Photocatalysis: Fundamentals
and Applications, BKC, Tokyo, 1999.
[20] E.N.S. Pelizzetti (Ed.), Homogeneous and Heterogeneous Photocatalysis, D.
Reidel Publishing Company, Dordrecht, 1986.
[21] N.E.P. Serpone (Ed.), Photocatalysis-Fundamentals and Applications, John
Wiley & Sons, New York, 1989.
[22] D.F.H.A.-E. Ollis (Ed.), Photocatalytic Purication and Treatment of Water and
Air, Elsevier, Amsterdam, 1993.
[23] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Environmental applications of semiconductor photocatalysis, Chem. Rev. 95 (1995) 6996.
[24] A. Fujishima, D.A. Tryk, M.S. In: A.J. Bard, S. Licht (Eds.), Encyclopedia of Electrochemistry, Vol. 6: Semiconductor Electrodes and Photoelectrochemistry,
Weinheim, 2002.
[25] A. Fujishima, X. Zhang, Titanium dioxide photocatalysis: present situation and
future approaches, C. R. Chim. 9 (2006) 750760.
[26] A. Fujishima, X. Zhang, D.A. Tryk, Heterogeneous photocatalysis: from water
photolysis to applications in environmental cleanup, Int. J. Hydrogen Energ.
32 (2007) 26642672.
[27] K. Nakata, B. Liu, Y. Goto, T. Ochiai, M. Sakai, H. Sakai, T. Murakami, M. Abe,
A. Fujishima, Visible light responsive electrospun TiO2 bers embedded with
WO3 nanoparticles, Chem. Lett. 40 (2011) 11611162.
[28] K. Nakata, B. Liu, Y. Ishikawa, M. Sakai, H. Saito, T. Ochiai, H. Sakai, T. Murakami,
M. Abe, K. Takagi, A. Fujishima, Fabrication and photocatalytic properties
of TiO2 nanotube arrays modied with phosphate, Chem. Lett. 40 (2011)
11071109.
[29] K. Nakata, N. Watanabe, Y. Yuda, D.A. Tryk, T. Ochiai, T. Murakami, Y. Koide, A.
Fujishima, Electrospun bers composed of Al2 O3 -TiO2 nanocrystals, J. Ceram.
Soc. Japan 117 (2009) 12031207.
[30] P.V. Kamat, Photochemistry on nonreactive and reactive (semiconductor) surfaces, Chem. Rev. 93 (1993) 267300.
[31] A. Heller, Chemistry and applications of photocatalytic oxidation of thin
organic lms, Acc. Chem. Res. 28 (1995) 503508.
[32] A. Mills, S.L. Hunte, An overview of semiconductor photocatalysis, J. Photochem. Photobiol. A 108 (1997) 135.
[33] J. Peral, X. Domnech, D.F. Ollis, Heterogeneous photocatalysis for purication, decontamination and deodorization of air, J. Chem. Technol. Biotechnol.
70 (1997) 117140.
[34] D.A. Tryk, A. Fujishima, K. Honda, Recent topics in photoelectrochemistry:
achievements and future prospects, Electrochim. Acta 45 (2000) 23632376.
[35] Y. Nosaka, T. Daimon, A.Y. Nosaka, Y. Murakami, Singlet oxygen formation in
photocatalytic TiO2 aqueous suspension, PCCP 6 (2004) 29172918.
[36] Y. Nosaka, S. Komori, K. Yawata, T. Hirakawa, A.Y. Nosaka, Photocatalytic [radical dot]OH radical formation in TiO2 aqueous suspension studied by several
detection methods, PCCP 5 (2003) 47314735.

[37] A. Janczyk,
E. Krakowska, G. Stochel, W. Macyk, Singlet oxygen photogeneration at surface modied titanium dioxide, J. Am. Chem. Soc. 128 (2006)
1557415575.
[38] Y. Nosaka, M. Nakamura, T. Hirakawa, Behavior of superoxide radicals formed
on TiO2 powder photocatalysts studied by a chemiluminescent probe method,
PCCP 4 (2002) 10881092.
[39] A.Y. Nosaka, E. Kojima, T. Fujiwara, H. Yagi, H. Akutsu, Y. Nosaka, Photoinduced
changes of adsorbed water on a TiO2 photocatalytic lm as studied by 1H NMR
spectroscopy, J. Phys. Chem. B 107 (2003) 1204212044.
[40] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M.
Shimohigoshi, T. Watanabe, Light-induced amphiphilic surfaces, Nature 388
(1997) 431432.
[41] H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, A. Heller, Local detection of photoelectrochemically produced H2 O2 with a Wired horseradish peroxidase
microsensor, J. Phys. Chem. 99 (1995) 1189611900.
[42] K. Ikeda, H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, Photocatalytic reactions
involving radical chain reactions using microelectrodes, J. Phys. Chem. B 101
(1997) 26172620.
[43] X. Chen, S.S. Mao, Synthesis of titanium dioxide (TiO2 ) nanomaterials, J.
Nanosci. Nanotechnol. 6 (2006) 906925.
[44] H. Bai, Z. Liu, D.D. Sun, Hierarchically multifunctional TiO2 nano-thorn membrane for water purication, Chem. Commun. 46 (2010) 65426544.
[45] J.S. Chen, C. Chen, J. Liu, R. Xu, S.Z. Qiao, X.W. Lou, Ellipsoidal hollow nanostructures assembled from anatase TiO2 nanosheets as a magnetically separable
photocatalyst, Chem. Commun. 47 (2011) 26312633.
[46] J.S. Chen, D. Luan, C.M. Li, F.Y.C. Boey, S. Qiao, X.W. Lou, TiO2 and SnO2 @TiO2
hollow spheres assembled from anatase TiO2 nanosheets with enhanced
lithium storage properties, Chem. Commun. 46 (2010) 82528254.
[47] J.S. Chen, Y.L. Tan, C.M. Li, Y.L. Cheah, D. Luan, S. Madhavi, F.Y.C.
Boey, L.A. Archer, X.W. Lou, Constructing hierarchical spheres from large

[48]
[49]

[50]

[51]

[52]

[53]

[54]
[55]

[56]

[57]

[58]
[59]

[60]

[61]

[62]

[63]

[64]

[65]

[66]

[67]
[68]

[69]

[70]

[71]
[72]

[73]

[74]

[75]

ultrathin anatase TiO2 nanosheets with nearly 100% exposed (001) facets for
fast reversible lithium storage, J. Am. Chem. Soc. 132 (2010) 61246130.
Y. Dai, C.M. Cobley, J. Zeng, Y. Sun, Y. Xia, Synthesis of anatase TiO2 nanocrystals with exposed {001} facets, Nano Lett. 9 (2009) 24552459.
W.Q. Fang, J.Z. Zhou, J. Liu, Z.G. Chen, C. Yang, C.H. Sun, G.R. Qian, J. Zou,
S.Z. Qiao, H.G. Yang, Hierarchical structures of single-crystalline anatase TiO2
nanosheets dominated by {001} facets, Chem. Eur. J. 17 (2011) 14231427.
W. Ho, J.C. Yu, S. Lee, Synthesis of hierarchical nanoporous F-doped TiO2
spheres with visible light photocatalytic activity, Chem. Commun. 14 (2006)
11151117.
Y. Kondo, H. Yoshikawa, K. Awaga, M. Murayama, T. Mori, K. Sunada, S.
Bandow, S. Iijima, Preparation, photocatalytic activities, and dye-sensitized
solar-cell performance of submicron-scale TiO2 hollow spheres, Langmuir 24
(2008) 547550.
H. Li, Z. Bian, J. Zhu, D. Zhang, G. Li, Y. Huo, H. Li, Y. Lu, Mesoporous titania
spheres with tunable chamber stucture and enhanced photocatalytic activity,
J. Am. Chem. Soc. 129 (2007) 84068407.
J. Li, D. Xu, Tetragonal faceted-nanorods of anatase TiO2 single crystals
with a large percentage of active {100} facets, Chem. Commun. 46 (2010)
23012303.
X. Li, Y. Xiong, Z. Li, Y. Xie, Large-scale fabrication of TiO2 hierarchical hollow
spheres, Inorg. Chem. 45 (2006) 34933495.
B. Liu, K. Nakata, M. Sakai, H. Saito, T. Ochiai, T. Murakami, K. Takagi, A.
Fujishima, Mesoporous TiO2 core shell spheres composed of nanocrystals
with exposed high-energy facets: facile synthesis and formation mechanism,
Langmuir 27 (2011) 85008508.
M. Liu, L. Piao, W. Lu, S. Ju, L. Zhao, C. Zhou, H. Li, W. Wang, Flower-like TiO2
nanostructures with exposed {001} facets: facile synthesis and enhanced
photocatalysis, Nanoscale 2 (2010) 11151117.
S. Liu, J. Yu, M. Jaroniec, Tunable photocatalytic selectivity of hollow TiO2
microspheres composed of anatase polyhedra with exposed {001} facets, J.
Am. Chem. Soc. 132 (2010) 1191411916.
S. Liu, J. Yu, S. Mann, Spontaneous construction of photoactive hollow TiO2
microspheres and chains, Nanotechnology 20 (2009) 325606.
X. L, F. Huang, X. Mou, Y. Wang, F. Xu, A general preparation strategy for
hybrid TiO2 hierarchical spheres and their enhanced solar energy utilization
efciency, Adv. Mater. 22 (2010) 37193722.
C. Wang, L. Yin, L. Zhang, Y. Qi, N. Lun, N. Liu, Large scale synthesis and
gas-sensing properties of anatase TiO2 three-dimensional hierarchical nanostructures, Langmuir 26 (2010) 1284112848.
L. Wang, T. Sasaki, Y. Ebina, K. Kurashima, M. Watanabe, Fabrication of
controllable ultrathin hollow shells by layer-by-layer assembly of exfoliated titania nanosheets on polymer templates, Chem. Mater. 14 (2002)
48274832.
P. Wen, H. Itoh, W. Tang, Q. Feng, Single nanocrystals of anatase-type TiO2
prepared from layered titanate nanosheets: formation mechanism and characterization of surface properties, Langmuir 23 (2007) 1178211790.
Q. Xiang, J. Yu, M. Jaroniec, Tunable photocatalytic selectivity of TiO2 lms
consisted of ower-like microspheres with exposed {001} facets, Chem. Commun. 47 (2011) 45324534.
Z. Zheng, B. Huang, X. Qin, X. Zhang, Y. Dai, Strategic synthesis of hierarchical
TiO2 microspheres with enhanced photocatalytic activity, Chem. Eur. J. 16
(2010) 1126611270.
Z. Zheng, B. Huang, X. Qin, X. Zhang, Y. Dai, M. Jiang, P. Wang, M.-H.
Whangbo, Highly efcient photocatalyst: TiO2 microspheres produced from
TiO2 nanosheets with a high percentage of reactive {001} facets, Chem. Eur.
J. 15 (2009) 1257612579.
P.D. Cozzoli, A. Kornowski, H. Weller, Low-temperature synthesis of soluble
and processable organic-capped anatase TiO2 nanorods, J. Am. Chem. Soc. 125
(2003) 1453914548.
X. Feng, J. Zhai, L. Jiang, The fabrication and switchable superhydrophobicity
of TiO2 nanorod lms, Angew. Chem. Int. Ed. 44 (2005) 51155118.
J. Joo, S.G. Kwon, T. Yu, M. Cho, J. Lee, J. Yoon, T. Hyeon, Large-scale synthesis of
TiO2 nanorods via nonhydrolytic sol-gel ester elimination reaction and their
application to photocatalytic inactivation of E. coli, J. Phys. Chem. B 109 (2005)
1529715302.
J.-N. Nian, H. Teng, Hydrothermal synthesis of single-crystalline anatase TiO2
nanorods with nanotubes as the precursor, J. Phys. Chem. B 110 (2006)
41934198.
Y. Wang, L. Zhang, K. Deng, X. Chen, Z. Zou, Low temperature synthesis and
photocatalytic activity of rutile TiO2 nanorod superstructures, J. Phys. Chem.
C 111 (2007) 27092714.
J.-J. Wu, C.-C. Yu, Aligned TiO2 nanorods and nanowalls, J. Phys. Chem. B 108
(2004) 33773379.
J.-M. Wu, T.-W. Zhang, Y.-W. Zeng, S. Hayakawa, K. Tsuru, A. Osaka, Largescale preparation of ordered titania nanorods with enhanced photocatalytic
activity, Langmuir 21 (2005) 69957002.
H. Xu, F. Jia, Z. Ai, L. Zhang, A general soft interface platform for the growth
and assembly of hierarchical rutile TiO2 nanorods spheres, Cryst. Growth Des.
7 (2007) 12161219.
H.J. Yun, H. Lee, J.B. Joo, W. Kim, J. Yi, Inuence of aspect ratio of TiO2 nanorods
on the photocatalytic decomposition of formic acid, J. Phys. Chem. C 113
(2009) 30503055.
H. An, B. Zhu, J. Li, J. Zhou, S. Wang, S. Zhang, S. Wu, W. Huang, Synthesis and
characterization of thermally stable nanotubular TiO2 and its photocatalytic
activity, J. Phys. Chem. C 112 (2008) 1877218775.

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189
[76] Y. Cheng, W. Huang, Y. Zhang, L. Zhu, Y. Liu, X. Fan, X. Cao, Preparation of TiO2
hollow nanobers by electrospining combined with solgel process, Cryst.
Eng. Comm. 12 (2010) 22562260.
[77] S. Chuangchote, J. Jitputti, T. Sagawa, S. Yoshikawa, Photocatalytic activity for
hydrogen evolution of electrospun TiO2 nanobers, ACS Appl. Mater. Interfaces 1 (2009) 11401143.
[78] H. Imai, M. Matsuta, K. Shimizu, H. Hirashima, N. Negishi, Preparation of TiO2
bers with well-organized structures, J. Mater. Chem. 10 (2000) 20052006.
[79] A. Kumar, R. Jose, K. Fujihara, J. Wang, S. Ramakrishna, Structural and
optical properties of electrospun TiO2 nanobers, Chem. Mater. 19 (2007)
65366542.
[80] D. Li, Y. Xia, Fabrication of titania nanobers by electrospinning, Nano Lett. 3
(2003) 555560.
[81] D. Li, Y. Xia, Direct fabrication of composite and ceramic hollow nanobers
by electrospinning, Nano Lett. 4 (2004) 933938.
[82] J.T. McCann, D. Li, Y. Xia, Electrospinning of nanobers with core-sheath,
hollow, or porous structures, J. Mater. Chem. 15 (2005) 735738.
[83] T. Peng, A. Hasegawa, J. Qiu, K. Hirao, Fabrication of titania tubules with
high surface area and well-developed mesostructural walls by surfactantmediated templating method, Chem. Mater. 15 (2003) 20112016.
[84] Y. Srivastava, I. Loscertales, M. Marquez, T. Thorsen, Electrospinning of hollow and core/sheath nanobers using a microuidic manifold, Microuid.
Nanouid. 4 (2007) 245250.
[85] C. Sun, N. Wang, S. Zhou, X. Hu, S. Zhou, P. Chen, Preparation of self-supporting
hierarchical nanostructured anatase/rutile composite TiO2 lm, Chem. Commun. (2008) 32933295.
[86] Y. Wu, M. Long, W. Cai, S. Dai, C. Chen, D. Wu, J. Bai, Preparation of photocatalytic anatase nanowire lms byin situoxidation of titanium plate,
Nanotechnology 20 (2009) 185703.
[87] S. Zhan, D. Chen, X. Jiao, C. Tao, Long TiO2 hollow bers with mesoporous
walls: sol-gel combined electrospun fabrication and photocatalytic properties, J. Phys. Chem. B 110 (2006) 1119911204.
[88] X. Zhang, T. Zhang, J. Ng, D.D. Sun, High-performance multifunctional TiO2
nanowire ultraltration membrane with a hierarchical layer structure for
water treatment, Adv. Funct. Mater. 19 (2009) 37313736.
[89] T. Zhao, Z. Liu, K. Nakata, S. Nishimoto, T. Murakami, Y. Zhao, L. Jiang, A.
Fujishima, Multichannel TiO2 hollow bers with enhanced photocatalytic
activity, J. Mater. Chem. 20 (2010) 50955099.
[90] S.P. Albu, A. Ghicov, J.M. Macak, R. Hahn, P. Schmuki, Self-organized,
free-standing TiO2 nanotube membrane for ow-through photocatalytic
applications, Nano Lett. 7 (2007) 12861289.
[91] Z. Liu, X. Zhang, S. Nishimoto, M. Jin, D.A. Tryk, T. Murakami, A. Fujishima,
Highly ordered TiO2 nanotube arrays with controllable length for photoelectrocatalytic degradation of phenol, J. Phys. Chem. C 112 (2007) 253259.
[92] Z. Liu, X. Zhang, S. Nishimoto, T. Murakami, A. Fujishima, Efcient photocatalytic degradation of gaseous acetaldehyde by highly ordered TiO2 nanotube
arrays, Environ. Sci. Technol. 42 (2008) 85478551.
[93] J.M. Macak, M. Zlamal, J. Krysa, P. Schmuki, Self-organized TiO2 nanotube
layers as highly efcient photocatalysts, Small 3 (2007) 300304.
[94] G.K. Mor, O.K. Varghese, M. Paulose, C.A. Grimes, Transparent highly ordered
TiO2 nanotube arrays via anodization of titanium thin lms, Adv. Funct. Mater.
15 (2005) 12911296.
[95] G.K. Mor, O.K. Varghese, M. Paulose, K. Shankar, C.A. Grimes, A review on
highly ordered, vertically oriented TiO2 nanotube arrays: fabrication, material properties, and solar energy applications, Sol. Energy Mater. Sol. Cells 90
(2006) 20112075.
[96] M. Paulose, H.E. Prakasam, O.K. Varghese, L. Peng, K.C. Popat, G.K. Mor, T.A.
Desai, C.A. Grimes, TiO2 nanotube arrays of 1000 m length by anodization of
titanium foil: phenol red diffusion, J. Phys. Chem. C 111 (2007) 1499214997.
[97] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor,
T.A. Latempa, A. Fitzgerald, C.A. Grimes, Anodic growth of highly ordered
TiO2 nanotube arrays to 134 m in length, J. Phys. Chem. B 110 (2006)
1617916184.
[98] X. Quan, S. Yang, X. Ruan, H. Zhao, Preparation of titania nanotubes and their
environmental applications as electrode, Environ. Sci. Technol. 39 (2005)
37703775.
[99] C. Ruan, M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, Fabrication of highly
ordered TiO2 nanotube arrays using an organic electrolyte, J. Phys. Chem. B
109 (2005) 1575415759.
[100] K. Shankar, J.I. Basham, N.K. Allam, O.K. Varghese, G.K. Mor, X. Feng, M.
Paulose, J.A. Seabold, K.-S. Choi, C.A. Grimes, Recent advances in the use of
TiO2 nanotube and nanowire arrays for oxidative photoelectrochemistry, J.
Phys. Chem. C 113 (2009) 63276359.
[101] K. Shankar, G.K. Mor, H.E. Prakasam, S. Yoriya, M. Paulose, O.K. Varghese, C.A.
Grimes, Highly-ordered TiO2 nanotube arrays up to 220 m in length: use
in water photoelectrolysis and dye-sensitized solar cells, Nanotechnology 18
(2007) 065707.
[102] D. Wang, T. Hu, L. Hu, B. Yu, Y. Xia, F. Zhou, W. Liu, Microstructured arrays of
TiO2 nanotubes for improved photo-electrocatalysis and mechanical stability,
Adv. Funct. Mater. 19 (2009) 19301938.
[103] J. Wang, Z. Lin, Freestanding TiO2 nanotube arrays with ultrahigh aspect ratio
via electrochemical anodization, Chem. Mater. 20 (2008) 12571261.
[104] S. Yoriya, C.A. Grimes, Self-assembled TiO2 nanotube arrays by anodization of
titanium in diethylene glycol: approach to extended pore widening, Langmuir
26 (2009) 417420.

187

[105] S. Yoriya, M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, Fabrication of vertically oriented TiO2 nanotube arrays using dimethyl sulfoxide electrolytes, J.
Phys. Chem. C 111 (2007) 1377013776.
[106] J. Yu, G. Dai, B. Cheng, Effect of crystallization methods on morphology and
photocatalytic activity of anodized TiO2 nanotube array lms, J. Phys. Chem.
C 114 (2010) 1937819385.
[107] Z. Zhang, G. Yuan, Y. Shi, L. Fang, H. Liang, L. Ding, Jin, Photoelectrocatalytic
activity of highly ordered TiO2 nanotube arrays electrode for azo dye degradation, Environ. Sci. Technol. 41 (2007) 62596263.
[108] H.-F. Zhuang, C.-J. Lin, Y.-K. Lai, L. Sun, J. Li, Some critical structure factors
of titanium oxide nanotube array in its photocatalytic activity, Environ. Sci.
Technol. 41 (2007) 47354740.
[109] Y. Aoyama, Y. Oaki, R. Ise, H. Imai, Mesocrystal nanosheet of rutile TiO2
and its reaction selectivity as a photocatalyst, Cryst. Eng. Comm. 14 (2012)
14051411.
[110] D.V. Bavykin, J.M. Friedrich, F.C. Walsh, Protonated titanates and TiO2 nanostructured materials: synthesis, properties, and applications, Adv. Mater. 18
(2006) 28072824.
[111] Y. Chen, G. Tian, Z. Ren, C. Tian, K. Pan, W. Zhou, H. Fu, Solvothermal synthesis, characterization, and formation mechanism of a single-layer anatase
TiO2 nanosheet with a porous structure, Eur. J. Inorg. Chem. 2011 (2011)
754760.
[112] J.-H. Choy, H.-C. Lee, H. Jung, H. Kim, H. Boo, Exfoliation and restacking route to
anatase-layered titanate nanohybrid with enhanced photocatalytic activity,
Chem. Mater. 14 (2002) 24862491.
[113] S. Feng, J. Yang, H. Zhu, M. Liu, J. Zhang, J. Wu, J. Wan, Synthesis of single
crystalline anatase TiO2 (001) tetragonal nanosheet-array lms on uorinedoped tin oxide substrate, J. Am. Ceram. Soc. 94 (2011) 310315.
[114] X. Gan, X. Gao, J. Qiu, P. He, X. Li, X. Xiao, TiO2 nanorod-derived synthesis
of upstanding hexagonal kassite nanosheet arrays: an intermediate route
to novel nanoporous TiO2 nanosheet arrays, Cryst. Growth Des. 12 (2011)
289296.
[115] X. Han, Q. Kuang, M. Jin, Z. Xie, L. Zheng, Synthesis of titania nanosheets with a
high percentage of exposed (001) facets and related photocatalytic properties,
J. Am. Chem. Soc. 131 (2009) 31523153.
[116] E. Hosono, H. Matsuda, I. Honma, M. Ichihara, H. Zhou, Synthesis of a perpendicular TiO2 nanosheet lm with the superhydrophilic property without UV
irradiation, Langmuir 23 (2007) 74477450.
[117] K.-i. Katsumata, C.E.J. Cordonier, T. Shichi, A. Fujishima, Photocatalytic activity
of NaNbO3 thin lms, J. Am. Chem. Soc. 131 (2009) 38563857.
[118] K.-i. Katsumata, S. Okazaki, C.E.J. Cordonier, T. Shichi, T. Sasaki, A. Fujishima,
Preparation and characterization of self-cleaning glass for vehicle with niobia
nanosheets, ACS Appl. Mater. Interfaces 2 (2010) 12361241.
[119] L. Kavan, M. Kalbc, M. Zukalov, I. Exnar, V. Lorenzen, R. Nesper, M. Graetzel,
Lithium storage in nanostructured TiO2 made by hydrothermal growth, Chem.
Mater. 16 (2004) 477485.
[120] G. Liu, H.G. Yang, X. Wang, L. Cheng, H. Lu, L. Wang, G.Q. Lu, H.-M. Cheng,
Enhanced photoactivity of oxygen-decient anatase TiO2 sheets with dominant {001} facets, J. Phys. Chem. C 113 (2009) 2178421788.
[121] G. Liu, H.G. Yang, X. Wang, L. Cheng, J. Pan, G.Q. Lu, H.-M. Cheng, Visible light
responsive nitrogen doped anatase TiO2 sheets with dominant {001} facets
derived from TiN, J. Am. Chem. Soc. 131 (2009) 1286812869.
[122] H. Liu, L. Gao, Preparation and properties of nanocrystalline alpha-Fe2 O3 sensitized TiO2 nanosheets as a visible light photocatalyst, J. Am. Ceram. Soc.
89 (2006) 370373.
[123] K.K. Manga, Y. Zhou, Y. Yan, K.P. Loh, Multilayer hybrid lms consisting of
alternating graphene and titania nanosheets with ultrafast electron transfer
and photoconversion properties, Adv. Funct. Mater. 19 (2009) 36383643.
[124] A. Matsuda, T. Matoda, T. Kogure, K. Tadanaga, T. Minami, M. Tatsumisago,
Formation and characterization of titania nanosheet-precipitated coatings via
solgel process with hot water treatment under vibration, Chem. Mater. 17
(2005) 749757.
[125] C.-W. Peng, T.-Y. Ke, L. Brohan, M. Richard-Plouet, J.-C. Huang, E. Puzenat, H.T. Chiu, C.-Y. Lee, (101)-Exposed anatase TiO2 nanosheets, Chem. Mater. 20
(2008) 24262428.
[126] N. Sakai, Y. Ebina, K. Takada, T. Sasaki, Electronic band structure of titania
semiconductor nanosheets revealed by electrochemical and photoelectrochemical studies, J. Am. Chem. Soc. 126 (2004) 58515858.
[127] N. Sakai, K. Fukuda, T. Shibata, Y. Ebina, K. Takada, T. Sasaki, Photoinduced
hydrophilic conversion properties of titania nanosheets, J. Phys. Chem. B 110
(2006) 61986203.
[128] T. Sasaki, M. Watanabe, Semiconductor nanosheet crystallites of quasi-TiO2
and their optical properties, J. Phys. Chem. B 101 (1997) 1015910161.
[129] T. Sasaki, M. Watanabe, H. Hashizume, H. Yamada, H. Nakazawa,
Macromolecule-like aspects for a colloidal suspension of an exfoliated
titanate. Pairwise association of nanosheets and dynamic reassembling process initiated from it, J. Am. Chem. Soc. 118 (1996) 83298335.
[130] T. Shibata, N. Sakai, K. Fukuda, Y. Ebina, T. Sasaki, Photocatalytic properties
of titania nanostructured lms fabricated from titania nanosheets, PCCP 9
(2007) 24132420.
[131] T. Shichi, K.-i. Katsumata, Development of photocatalytic self-cleaning glasses
utilizing metal oxide nanosheets, Hyomen Gijutsu 61 (2010) 3035.
[132] T. Tachikawa, T. Yui, M. Fujitsuka, K. Takagi, T. Majima, Photocatalytic electron transfer in hybrid titania nanosheets studied by nanosecond laser ash
photolysis, Chem. Lett. 34 (2005) 15221523.

188

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189

[133] T. Tanaka, Y. Ebina, K. Takada, K. Kurashima, T. Sasaki, Oversized titania


nanosheet crystallites derived from ux-grown layered titanate single crystals, Chem. Mater. 15 (2003) 35643568.
[134] Y. Umemura, E. Shinohara, A. Koura, T. Nishioka, T. Sasaki, Photocatalytic
decomposition of an alkylammonium cation in a langmuirblodgett lm of
a titania nanosheet, Langmuir 22 (2006) 38703877.
[135] M. Wei, Y. Konishi, H. Arakawa, Synthesis and characterization of nanosheetshaped titanium dioxide, J. Mater. Chem. 42 (2006) 529533.
[136] C.Z. Wen, J.Z. Zhou, H.B. Jiang, Q.H. Hu, S.Z. Qiao, H.G. Yang, Synthesis of microsized titanium dioxide nanosheets wholly exposed with high-energy {001}
and {100} facets, Chem. Commun. 47 (2011) 44004402.
[137] G. Xiang, T. Li, J. Zhuang, X. Wang, Large-scale synthesis of metastable TiO2 (B)
nanosheets with atomic thickness and their photocatalytic properties, Chem.
Commun. 46 (2010) 68016803.
[138] Q. Xiang, J. Yu, M. Jaroniec, Nitrogen and sulfur co-doped TiO2 nanosheets
with exposed {001} facets: synthesis, characterization and visible-light photocatalytic activity, PCCP 13 (2011) 48534861.
[139] Q. Xiang, J. Yu, W. Wang, M. Jaroniec, Nitrogen self-doped nanosized TiO2
sheets with exposed {001} facets for enhanced visible-light photocatalytic
activity, Chem. Commun. 47 (2011) 69066908.
[140] H.G. Yang, G. Liu, S.Z. Qiao, C.H. Sun, Y.G. Jin, S.C. Smith, J. Zou, H.M. Cheng, G.Q.
Lu, Solvothermal synthesis and photoreactivity of anatase TiO2 nanosheets
with dominant {001} facets, J. Am. Chem. Soc. 131 (2009) 40784083.
[141] J. Yu, J. Fan, K. Lv, Anatase TiO2 nanosheets with exposed (001) facets:
improved photoelectric conversion efciency in dye-sensitized solar cells,
Nanoscale 2 (2010) 21442149.
[142] J. Yu, L. Qi, M. Jaroniec, Hydrogen production by photocatalytic water splitting
over Pt/TiO2 nanosheets with exposed (001) facets, J. Phys. Chem. C 114 (2010)
1311813125.
[143] I.M. Arabatzis, P. Falaras, Synthesis of porous nanocrystalline TiO2 foam, Nano
Lett. 3 (2002) 249251.
[144] R.A. Caruso, M. Giersig, F. Willig, M. Antonietti, Porous Coral-like TiO2 structures produced by templating polymer gels, Langmuir 14 (1998) 63336336.
[145] X. Fan, H. Fei, D.H. Demaree, D.P.J.M. Brennan, St. John, S.R.J. Oliver, Polymer gel
templating of free-standing inorganic monoliths for photocatalysis, Langmuir
25 (2009) 58355839.
[146] G. Hasegawa, K. Kanamori, K. Nakanishi, T. Hanada, Facile preparation
of hierarchically porous TiO2 monoliths, J. Am. Ceram. Soc. 93 (2010)
31103115.
[147] G. Hasegawa, K. Morisato, K. Kanamori, K. Nakanishi, New hierarchically
porous titania monoliths for chromatographic separation media, J. Sep. Sci.
34 (2011) 30043010.
[148] J. Konishi, K. Fujita, K. Nakanishi, K. Hirao, Monolithic TiO2 with controlled
multiscale porosity via a template-free solgel process accompanied by phase
separation, Chem. Mater. 18 (2006) 60696074.
[149] J. Konishi, K. Fujita, K. Nakanishi, K. Hirao, K. Morisato, S. Miyazaki, M. Ohira,
Solgel synthesis of macromesoporous titania monoliths and their applications to chromatographic separation media for organophosphate compounds,
J. Chromatogr. A 1216 (2009) 73757383.
[150] S.O. Kucheyev, T.F. Baumann, Y.M. Wang, T. van Buuren, J.H. Satcher Jr.,
Synthesis and electronic structure of low-density monoliths of nanoporous
nanocrystalline anatase TiO2 , J. Electron Spectrosc. Relat. Phenom. 144147
(2005) 609612.
[151] Y. Lai, C. Lin, J. Huang, H. Zhuang, L. Sun, T. Nguyen, Markedly controllable
adhesion of superhydrophobic spongelike nanostructure TiO2 lms, Langmuir 24 (2008) 38673873.
[152] J. Liu, M. Li, J. Wang, Y. Song, L. Jiang, T. Murakami, A. Fujishima, Hierarchically macro-/mesoporous TiSi oxides photonic crystal with highly efcient
photocatalytic capability, Environ. Sci. Technol. 43 (2009) 94259431.
[153] S.I. Matsushita, T. Miwa, D.A. Tryk, A. Fujishima, New mesostructured porous
TiO2 surface prepared using a two-dimensional array-based template of silica
particles, Langmuir 14 (1998) 64416447.
[154] M. Ren, R. Ravikrishna, K.T. Valsaraj, Photocatalytic degradation of gaseous
organic species on photonic band-gap titania, Environ. Sci. Technol. 40 (2006)
70297033.
[155] P.K. Sharma, A. Ramanan, The role of N,N-dimethylaniline in the formation of
titania gel monolith by sol-gel method, J. Mater. Chem. 31 (1996) 773777.
[156] G. Tian, Y. Chen, W. Zhou, K. Pan, C. Tian, X.-r. Huang, H. Fu, 3D hierarchical
ower-like TiO2 nanostructure: morphology control and its photocatalytic
property, Cryst. Eng. Comm. 13 (2011) 29943000.

A. Ferreira
[157] S.A. Toms, O. Zelaya, R. Palomino, R. Lozada, O. Garca, J.M. Ynez,
da Silva, Optical characterization of sol gel TiO2 monoliths doped with Brilliant
Green, Eur. Phys. J. 153 (2008) 255258.
[158] J.E. Wijnhoven, Preparation of photonic crystals made of air spheres in titania,
Science 281 (1998) 802804.
[159] J.E.G.J. Wijnhoven, L. Bechger, W.L. Vos, Fabrication and characterization
of large macroporous photonic crystals in titania, Chem. Mater. 13 (2001)
44864499.
[160] B. Yao, L. Zhang, Preparation and characterization of mesoporous titania gelmonolith, J. Mater. Chem. 34 (1999) 59835987.
[161] J. Yu, L. Zhang, B. Cheng, Y. Su, Hydrothermal preparation and photocatalytic activity of hierarchically sponge-like macro-/mesoporous titania, J.
Phys. Chem. C 111 (2007) 1058210589.
[162] J.G. Yu, Y.R. Su, B. Cheng, Template-free fabrication and enhanced photocatalytic activity of hierarchical macro-/mesoporous titania, Adv. Funct. Mater.
17 (2007) 19841990.

[163] K. Hou, B. Tian, F. Li, Z. Bian, D. Zhao, C. Huang, Highly crystallized mesoporous
TiO2 lms and their applications in dye sensitized solar cells, J. Mater. Chem.
15 (2005) 24142420.
[164] S.H. Ahn, J.H. Koh, J.A. Seo, J.H. Kim, Structure control of organized mesoporous
TiO2 lms templated by graft copolymers for dye-sensitized solar cells, Chem.
Commun. 46 (2010) 19351937.
[165] W.-G. Yang, F.-R. Wan, Q.-W. Chen, J.-J. Li, D.-S. Xu, Controlling synthesis of
well-crystallized mesoporous TiO2 microspheres with ultrahigh surface area
for high-performance dye-sensitized solar cells, J. Mater. Chem. 20 (2010)
28702876.
[166] S. Ding, J.S. Chen, Z. Wang, Y.L. Cheah, S. Madhavi, X. Hu, X.W. Lou, TiO2 hollow
spheres with large amount of exposed (001) facets for fast reversible lithium
storage, J. Mater. Chem. 21 (2011) 16771680.
[167] J.M. Szeifert, J.M. Feckl, D. Fattakhova-Rohlng, Y. Liu, V. Kalousek, J.
Rathousky, T. Bein, Ultrasmall titania nanocrystals and their direct assembly into mesoporous structures showing fast lithium insertion, J. Am. Chem.
Soc. 132 (2010) 1260512611.
[168] P. Periyat, N. Leyland, D.E. McCormack, J. Colreavy, D. Corr, S.C. Pillai, Rapid
microwave synthesis of mesoporous TiO2 for electrochromic displays, J.
Mater. Chem. 20 (2010) 36503655.
[169] L. Yao, T.W. Haas, A. Guiseppi-Elie, G.L. Bowlin, D.G. Simpson, G.E. Wnek, Electrospinning and stabilization of fully hydrolyzed poly(vinyl alcohol) bers,
Chem. Mater. 15 (2003) 18601864.
[170] K.H. Lee, H.Y. Kim, M.S. Khil, Y.M. Ra, D.R. Lee, Characterization of
nano-structured poly(-caprolactone) nonwoven mats via electrospinning,
Polymer 44 (2003) 12871294.
[171] C.B. Almquist, P. Biswas, Role of synthesis method and particle size of nanostructured TiO2 on its photoactivity, J. Catal. 212 (2002) 145156.
[172] I.-D. Kim, A. Rothschild, B.H. Lee, D.Y. Kim, S.M. Jo, H.L. Tuller, Ultrasensitive
chemiresistors based on electrospun TiO2 nanobers, Nano Letters 6 (2006)
20092013.
[173] J.-A. Park, J. Moon, S.-J. Lee, S.H. Kim, T. Zyung, H.Y. Chu, Structure and CO
gas sensing properties of electrospun TiO2 nanobers, Mater. Lett. 64 (2010)
255257.
[174] J. Moon, J.-A. Park, S.-J. Lee, T. Zyung, I.-D. Kim, Pd-doped TiO2 nanober networks for gas sensor applications, Sens. Actuat. B: Chem. 149 (2010) 301305.
[175] M.Y. Song, D.K. Kim, S.M. Jo, D.Y. Kim, Enhancement of the photocurrent generation in dye-sensitized solar cell based on electrospun TiO2 electrode by
surface treatment, Synth. Met. 155 (2005) 635638.
[176] M.Y. Song, D.K. Kim, K.J. Ihn, S.M. Jo, D.Y. Kim, New application of electrospun
TiO2 electrode to solid-state dye-sensitized solar cells, Synth. Met. 153 (2005)
7780.
[177] A. Kumar, R. Jose, K. Fujihara, J. Wang, S. Ramakrishna, Structural and
optical properties of electrospun TiO2 nanobers, Chem. Mater. 19 (2007)
65366542.
[178] S. Mu Jo, M. Yeon Song, Y. Rack Ahn, C. Rae Park, D. Young Kim, Nanobril
formation of electrospun TiO2 bers and its application to dye-sensitized solar
cells, J. Macromol. Sci. Part A 42 (2005) 15291540.
[179] S.H. Nam, H.-S. Shim, Y.-S. Kim, M.A. Dar, J.G. Kim, W.B. Kim, Ag or Au
nanoparticle-embedded one-dimensional composite TiO2 nanobers prepared via electrospinning for use in lithium-ion batteries, ACS Appl. Mater.
Interfaces 2 (2010) 20462052.
[180] S. Ramakrishna, K. Fujihara, W.-E. Teo, T.-C. Lim, Z. Ma, An Introduction to
Electrospinning and Nanobers, World Scientic, Singapore, 2005.
[181] A.K. Alves, F.A. Berutti, F.J. Clemens, T. Graule, C.P. Bergmann, Photocatalytic
activity of titania bers obtained by electrospinning, Mater. Res. Bull. 44
(2009) 312317.
[182] T. Zhao, Z. Liu, K. Nakata, S. Nishimoto, T. Murakami, Y. Zhao, L. Jiang, A.
Fujishima, Multichannel TiO2 hollow bers with enhanced photocatalytic
activity, J. Mater. Chem. 20 (2010) 5099.
[183] C.A. Grimes, G.K. Mor, TiO2 Nanotube Arrays: Synthesis, Properties and Applications, Springer, 2009.
[184] R. Wang, K. Hashimoto, A. Fujishima, Light-induced amphiphilic surfaces,
Nature 388 (1997) 432.
[185] K. Nakata, M. Sakai, T. Ochiai, T. Murakami, K. Takagi, A. Fujishima, Antireection and self-cleaning properties of a moth-eye-like surface coated with TiO2
particles, Langmuir 27 (2011) 32753278.
[186] K. Nakata, S. Nishimoto, A. Kubo, D.A. Tryk, T. Ochiai, T. Murakami, A.
Fujishima, Fabrication and application of TiO2 -based superhydrophilicsuperhydrophobic patterns on titanium substrates for offset printing, Chem.
Asian J. 4 (2009) 984988.
[187] S. Nishimoto, A. Kubo, K. Nohara, X. Zhang, N. Taneichi, T. Okui, Z. Liu,
K. Nakata, H. Sakai, T. Murakami, M. Abe, T. Komine, A. Fujishima, TiO2 based superhydrophobic-superhydrophilic pattern: fabrication via ink-jet
technique and application to offset printing plate, Appl. Sur. Sci. 255 (2009)
62216225.
[188] Y. Suda, T. Shimada, M. Tabuchi, A new technology of reusable plate using
TiO2 photocatalysis, Proc. Tech. Assoc. Graphic Arts (2000) 125134.
[189] A. Fujishima, K. Hashimoto, T. Watanabe, TiO2 Photocatalysis, Bkc, Inc., Tokyo,
1999.
[190] K. Nakata, K. Udagawa, D.A. Tryk, S. Nishimoto, T. Ochiai, H. Sakai, T. Murakami,
M. Abe, A. Fujishima, Fabrication of micro-patterned TiO2 thin lm incorporating Ag nanoparticles, Mater. Lett. 63 (2009) 16281630.
[191] S.D. Gillmor, A.J. Thiel, T.C. Strother, L.M. Smith, M.G. Lagally,
Hydrophilic/hydrophobic patterned surfaces as templates for DNA arrays,
Langmuir 16 (2000) 72237228.

K. Nakata, A. Fujishima / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 169189
[192] H. Gau, S. Herminghaus, P. Lenz, R. Lipowsky, Liquid morphologies on structured surfaces: from microchannels to microchips, Science 283 (1999) 4649.
[193] G. Cao, Y. Wang, Nanostructures and Nanomaterials: Synthesis, Properties
and Applications, World Scientic Publishing Co. Pte. Ltd., Singapore, 2011.
[194] M. Khler, W. Fritzsche, M.K. Hler, Nanotechnology: An Introduction to
Nanostructuring Techniques, Wiley-VCH, 2004.
[195] X. Zhang, M. Jin, Z. Liu, D.A. Tryk, S. Nishimoto, T. Murakami, A. Fujishima,
Superhydrophobic TiO2 surfaces: preparation, photocatalytic wettability conversion, and superhydrophobicsuperhydrophilic patterning, J. Phys. Chem.
C 111 (2007) 1452114529.
[196] X. Zhang, M. Jin, Z. Liu, S. Nishimoto, H. Saito, T. Murakami, A. Fujishima,
Preparation and photocatalytic wettability conversion of TiO2 -based superhydrophobic surfaces, Langmuir 22 (2006) 94779479.
[197] R.S. Dibbell, G.R. Soja, R.M. Hoth, D.F. Watson, Photocatalytic patterning of
monolayers for the site-selective deposition of quantum dots onto TiO2 surfaces, Langmuir 23 (2007) 34323439.

189

[198] Z.-Z. Gu, A. Fujishima, O. Sato, Patterning of a colloidal crystal lm on a


modied hydrophilic and hydrophobic surface, Angew. Chem. 114 (2002)
21712174.
[199] Y. Lai, C. Lin, H. Wang, J. Huang, H. Zhuang, L. Sun,
Superhydrophilicsuperhydrophobic micropattern on TiO2 nanotube
lms by photocatalytic lithography, Electrochem. Commun. 10 (2008)
387391.
[200] K. Nakata, S. Nishimoto, Y. Yuda, T. Ochiai, T. Murakami, A. Fujishima,
Rewritable superhydrophilic-superhydrophobic patterns on a sintered titanium dioxide substrate, Langmuir 26 (2010) 11628
11630.
[201] K. Nakata, H. Kimura, M. Sakai, T. Ochiai, H. Sakai, T. Murakami, M.
Abe, A. Fujishima, UV/thermally-driven rewritable wettability patterns
on TiO2 -PDMS composite lms, ACS Appl. Mater. Interfaces 2 (2010)
24852488.

Vous aimerez peut-être aussi