Вы находитесь на странице: 1из 68

Surface Science Reports 63 (2008) 515–582

Contents lists available at ScienceDirect

Surface Science Reports


journal homepage: www.elsevier.com/locate/surfrep

TiO2 photocatalysis and related surface phenomena


Akira Fujishima a,∗ , Xintong Zhang b , Donald A. Tryk c
a
Kanagawa Academy of Science and Technology, 3-2-1 Sakado, Takatsu-ku, Kawasaki-shi, Kanagawa 213-0012, Japan
b
Center for Advanced Optoelectronic Functional Materials Research, Northeast Normal University, 5268 Renmin Street, Changchun 130024, China
c
Fuel Cell Nanomaterials Center, University of Yamanashi, Takeda 4-3-11, Koufu, Yamanashi 400-8510, Japan

article info a b s t r a c t

Article history: The field of photocatalysis can be traced back more than 80 years to early observations of the chalking of
Accepted 1 October 2008 titania-based paints and to studies of the darkening of metal oxides in contact with organic compounds
editor: Y. Murata in sunlight. During the past 20 years, it has become an extremely well researched field due to practical
interest in air and water remediation, self-cleaning surfaces, and self-sterilizing surfaces. During the same
Keywords: period, there has also been a strong effort to use photocatalysis for light-assisted production of hydrogen.
Titanium dioxide
The fundamental aspects of photocatalysis on the most studied photocatalyst, titania, are still being
Titania
TiO2
actively researched and have recently become quite well understood. The mechanisms by which certain
Self-cleaning surfaces types of organic compounds are decomposed completely to carbon dioxide and water, for example,
Superhydrophilic effect have been delineated. However, certain aspects, such as the photo-induced wetting phenomenon,
Anion doping remain controversial, with some groups maintaining that the effect is a simple one in which organic
Water splitting contaminants are decomposed, while other groups maintain that there are additional effects in which
Environmental cleaning the intrinsic surface properties are modified by light. During the past several years, powerful tools such
as surface spectroscopic techniques and scanning probe techniques performed on single crystals in ultra-
high vacuum, and ultrafast pulsed laser spectroscopic techniques have been brought to bear on these
problems, and new insights have become possible. Quantum chemical calculations have also provided
new insights. New materials have recently been developed based on titania, and the sensitivity to visible
light has improved. The new information available is staggering, but we hope to offer an overview of
some of the recent highlights, as well as to review some of the origins and indicate some possible new
directions.
© 2008 Elsevier B.V. All rights reserved.

Contents

1. Introduction........................................................................................................................................................................................................................516
2. Historical overview ............................................................................................................................................................................................................516
3. Properties of TiO2 materials ..............................................................................................................................................................................................519
3.1. Crystal structures ...................................................................................................................................................................................................519
3.2. Electronic properties .............................................................................................................................................................................................520
3.3. Surface structure studies.......................................................................................................................................................................................523
3.4. Surface chemical studies: Interactions with water .............................................................................................................................................523
3.5. Surface chemical studies: Interactions with dioxygen and other species .........................................................................................................527
3.6. Bulk chemistry—Hydrogen....................................................................................................................................................................................527
3.7. Electrochemical properties ...................................................................................................................................................................................529
3.8. Photoelectrochemical properties..........................................................................................................................................................................534
4. Fundamentals of photocatalysis........................................................................................................................................................................................534
4.1. Mechanisms of photocatalysis ..............................................................................................................................................................................534
4.1.1. Photoelectrochemical basis of photocatalysis ......................................................................................................................................534
4.1.2. Time scales ..............................................................................................................................................................................................538
4.1.3. Trapping of electrons and holes.............................................................................................................................................................541

∗ Corresponding author. Tel.: +81 (0)44 819 2020; fax: +81 (0)44 819 2038.
E-mail address: fujishima@newkast.or.jp (A. Fujishima).

0167-5729/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfrep.2008.10.001
516 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

4.1.4. Oxidizing species at the TiO2 surface ....................................................................................................................................................542


4.1.5. Role of molecular oxygen .......................................................................................................................................................................545
4.1.6. Effect of crystal face................................................................................................................................................................................547
4.1.7. Remote photocatalysis ...........................................................................................................................................................................549
4.2. Photocatalytic reactions ........................................................................................................................................................................................551
4.2.1. Decomposition of gaseous pollutants ...................................................................................................................................................551
4.2.2. Decomposition of aqueous pollutants...................................................................................................................................................552
4.2.3. Decomposition of liquid and solid films ...............................................................................................................................................553
4.2.4. Photocatalytic sterilization ....................................................................................................................................................................553
4.3. Visible-light-induced photocatalysis....................................................................................................................................................................555
4.3.1. Non-metal doping...................................................................................................................................................................................555
4.3.2. Origin of visible light photoactivity.......................................................................................................................................................556
4.3.3. Activity and stability of N-doped TiO2 photocatalysts .........................................................................................................................558
5. Fundamentals of the photo-induced hydrophilic (PIH) effect ........................................................................................................................................560
5.1. Overview ................................................................................................................................................................................................................560
5.2. Mechanisms of the PIH effect................................................................................................................................................................................562
5.2.1. Decomposition of organic films .............................................................................................................................................................562
5.2.2. Reductive mechanism ............................................................................................................................................................................563
5.2.3. Oxidative mechanism.............................................................................................................................................................................563
5.2.4. Combined redox mechanism .................................................................................................................................................................563
5.2.5. Visible-light-induced PIH effect.............................................................................................................................................................564
6. Brief review of applications...............................................................................................................................................................................................565
6.1. Self-cleaning surfaces ............................................................................................................................................................................................565
6.2. Water purification .................................................................................................................................................................................................567
6.3. Air purification .......................................................................................................................................................................................................568
6.4. Self-sterilizing surfaces .........................................................................................................................................................................................569
6.5. Anti-fogging surfaces.............................................................................................................................................................................................570
6.6. Heat transfer and heat dissipation........................................................................................................................................................................571
6.7. Anticorrosion applications ....................................................................................................................................................................................571
6.8. Environmentally friendly surface treatment .......................................................................................................................................................572
6.9. Photocatalytic lithography ....................................................................................................................................................................................572
6.10. Photochromism......................................................................................................................................................................................................574
6.11. Microchemical systems .........................................................................................................................................................................................574
7. Summary ............................................................................................................................................................................................................................575
Appendix. TiO2 film preparation methods ..................................................................................................................................................................576
References...........................................................................................................................................................................................................................576

1. Introduction have tried to put together an overview of some of the more


fundamental aspects, which are in their own right extremely
Photocatalysis is generally thought of as the catalysis of a scientifically interesting and which also need to be better
photochemical reaction at a solid surface, usually a semiconductor understood in order to make significant progress with applications.
[1–16]. This simple definition, while correct and useful, however, The review will be divided into several sections: 2. Historical
conceals the fact that there must be at least two reactions occurring overview; 3. Properties of TiO2 materials; 4. Fundamentals of
simultaneously, the first involving oxidation, from photogenerated photocatalysis; 5. Fundamentals of the photo-induced hydrophilic
holes, and the second involving reduction, from photogenerated effect; 6. Brief review of applications; 7. Summary, and Appendix
electrons. Both processes must be balanced precisely in order for (film preparation methods).
the photocatalyst itself not to undergo change, which is, after all,
one of the basic requirements for a catalyst.
It will be seen in this review of the fundamentals and selected 2. Historical overview
applications of photocatalysis, principally on titanium dioxide,
that there is a host of possible photochemical, chemical and We will give a brief overview of the early history of
electrochemical reactions that can occur on the photocatalyst photocatalysis, which will be based just on papers that we have
surface. The types of reactions occurring, their extent and their been able to access, which means that we will almost certainly
rates depend upon a host of factors that are still in the process be ignoring some important papers. The earliest work that we
of being unraveled. Furthermore, there can indeed be changes have been able to find is that of Renz, at the University of Lugano
that occur, involving the surface and bulk structure and even (Switzerland), who reported in 1921 [17] that titania is partially
decomposition of the photocatalyst, a fact that appears to stretch reduced during illumination with sunlight in the presence of an
the definition of the term. organic compound such as glycerol, the oxide turning from white
This topic started its early history as mostly a nuisance involving to a dark color, such as grey, blue or even black; he also found
the chalking of titania-based paints [17,18] and then gradually
similar phenomena with CeO2 , Nb2 O5 and Ta2 O5 . For TiO2 , the
transformed into a highly useful approach to the remediation of
reaction proposed was:
water and air and then into an approach to maintain surfaces clean
and sterile. Along the way, it has also transformed into an approach TiO2 + light → Ti2 O3 or TiO. (2.1)
to photolytically split water into hydrogen and oxygen [19–21]
and also an approach to perform selective oxidation reactions in Baur and Perret, at the Swiss Federal Institute of Technology, were
organic chemistry [22]. the first to report, in 1924, the photocatalytic deposition of a silver
Clearly, with so many varied aspects, photocatalysis is nearly salt on zinc oxide to produce metallic silver [23]. Even at this
impossible to review comprehensively. In the present review, we early date, the authors suspected that both oxidation and reduction
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 517

were occurring simultaneously. The reaction pathway proposed


was this:

ZnO + hν → h+ + e− (2.2)
1 1
h+ + OH− → O2 + H2 O (2.3)
4 2
e− + Ag+ → Ag0 . (2.4)

Three years later, Baur and Neuweiler proposed simultaneous


oxidation and reduction to explain the production of hydrogen
peroxide on zinc oxide [24].

2h+ + CH2 O + 2OH− → CO + 2H2 O (2.5)


2e + 2H + O2 → H2 O2 .
− +
(2.6)

It was not until many years later that this was absolutely
confirmed, however. In 1932, Renz reported the photocatalytic
reduction of silver nitrate to metallic silver and gold chloride to Fig. 2.1. Original data of Goodeve and Kitchener showing the photocatalytic
metallic gold on a number of illuminated oxides, including TiO2 and decomposition of a dye (‘‘chlorazol sky blue)’’ adsorbed on anatase powder under
UV illumination at 365 nm [26].
Nb2 O5 , [25] and discussed the results in terms of the Baur redox
© 1938, Royal Society of Chemistry.
mechanism.
It has been recognized for quite a long time that titania-based
of Baur and Neuweiler was finally confirmed, with the overall
exterior paints tend to undergo ‘‘chalking’’ in strong sunlight.
reaction:
This means that a non-adherent, white powdery substance tends
to form on the surface, similar to the chalk on a blackboard. RHOH + H2 O + O2 → H2 O2 + R(OH)2 . (2.7)
This effect was recognized to result from the actual removal of
Markham, first at the Catholic University of America and later
part of the organic component of the paint, leaving the titania
at St. Joseph’s College (USA), continued to study photocatalytic
itself exposed. With this background, Goodeve and Kitchener, at
reactions on ZnO, and her papers constitute an impressive, yet
University College, London, carried out an excellent study on the
underappreciated, body of work [28,30,31,35–39]. This work
photocatalytic decomposition of a dye on titania powder in air in
culminated in a highly intriguing study in which Markham and
1938, including absorption spectra and determination of quantum
co-worker Upreti constructed and studied a number of different
yields (Fig. 2.1) [26]. These authors proposed that titania acts as a
types of photo-assisted fuel cells, using illuminated ZnO as the
catalyst to accelerate the photochemical oxidation and also studied
photo-anode with formamide or several alcohols as the organic
a number of other oxides and speculated on the precise mechanism
substrates [39]. At the dark cathode (platinum), several different
[27]. In 1949, Jacobsen, at the National Lead Company (USA), also redox mediators were examined, with atmospheric oxygen
attempted to explain the paint chalking phenomenon in terms of ultimately being the electron acceptor. The authors may have been
a redox mechanism. He found a correlation between the tendency discouraged by the inevitable problem of ZnO photocorrosion,
of a number of different titania powders to undergo photo-induced which prevented this system from reaching practical application.
reduction in the presence of organic compounds to their chalking It was not until years later that the same basic ideas were re-
tendency [18]. The photo-induced reduction was measured as a examined with TiO2 . Unfortunately, Markham and Laidler, in their
loss of reflectivity, due to the discoloration of the powder upon initial study in 1953, examined TiO2 but subsequently abandoned
reduction, presumably to various oxygen-deficient forms, all the it, since it did not produce measurable amounts of hydrogen
way to Ti2 O3 . The author proposed a cyclic redox process in which peroxide [28].
the titania was reduced while the organic paint components were It is also interesting to note that Stephens et al. (Wayne
oxidized, followed by re-oxidation of the titania by oxygen from State University), in their study in 1955 of hydrogen peroxide
the air. The changes experienced by the titania were recognized to production on a large assortment of illuminated semiconductors,
be completely reversible, while those experienced by the organic but, unfortunately, not TiO2 , remarked that ‘‘zinc oxide and the
paint were recognized to be irreversible, leading to the formation other catalytic solids should not be abandoned as devices for
of water-soluble organic acids and CO2 . Even though Jacobsen was capturing solar energy in a form capable of transfer to some
apparently unaware of the work of Baur on the redox mechanism, chemical system’’ [32]. These authors found that CdS was the most
he referred to the 1921 paper of Renz on the photo-reduction of active photocatalyst, exceeding ZnO in activity.
metal oxides and proposed the same basic mechanism that had In a study reported in 1956 in Nature, Hindson and Kelly
been proposed by Baur; thus, a foundation was laid for later work (Defense Standards Laboratory) reported on the effects of various
on the redox mechanism. rot-inhibitors on tent fabrics for use in Australia. They examined
During the 1950s, the development of photocatalysis shifted to the effects of fabric strength after one year of exposure to sunlight.
zinc oxide. In 1953, two studies appeared in which the puzzling They stated: ‘‘The effect of anatase is startling. Fabrics containing
phenomenon of hydrogen peroxide production on zinc oxide 3% of this pigment lost 90% in strength’’.
illuminated with UV light was studied [28,29], followed by a series In 1958, Kennedy et al., at the University of Edinburgh, studied
of follow-up studies in ensuing years [30–34]. In these studies, the the photo-adsorption of O2 on TiO2 in order to try to more fully
overall reactions and mechanisms were completely clarified, and understand the photocatalytic process [40]. They concluded that
it became apparent that an organic compound was oxidized while electrons were transferred to O2 as a result of photoexcitation, and
atmospheric oxygen was reduced. Even in the earliest study, an the resulting reduced form of O2 adsorbed on the TiO2 surface.
overall reaction with phenol to produce catechol was proposed, These authors found a correlation between the ability of the TiO2
and the involvement of radical species such as the hydroxyl radical sample to photocatalytically decompose chlorazol sky blue (the
(•OH) was also speculated upon [28]. Thus, the original proposal same dye used earlier by Goodeve and Kitchener) and the ability
518 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

to photo-adsorb O2 . This phenomenon is certainly important for


photocatalysis and will be commented upon later.
During this period, researchers in Russia were active. Photo-
adsorption of O2 on illuminated ZnO was studied by Terenin and
Solonitzin at the University of Leningrad (now University of St.
Petersburg) [41]. In a very interesting early work, Filimonov, at
the same institution, compared the photocatalytic oxidation of
isopropanol to acetone on ZnO and TiO2 [42] and concluded that
the mechanism on TiO2 involved an overall reduction of O2 to H2 O,
while the reduction of O2 on ZnO only went as far as H2 O2 . On TiO2 ,
the surface reactions were proposed to be:

TiO∗2 + (CH3 )2 CHOH → (CH3 )2 CO + TiO + H2 O (2.8)


1
TiO + O2 → TiO2 . (2.9)
2
Thus, this mechanism is a more detailed version of the Baur cyclic
mechanism. It involves the removal of a surface lattice oxygen Fig. 2.2. Photocurrent vs. potential for illuminated rutile single crystal. SCE refers
atom, which would be a kind of reduction process. This mechanism to the saturated calomel electrode, which is 0.241 V vs. the standard hydrogen
will be discussed later also, in Section 3.3. electrode (SHE) (taken from Fujishima et al. [214]).
In Japan, at the Kyoto Institute of Technology, an early study © 1971, Chemical Society of Japan.
(1964) by Kato and Mashio also found that various types of
titania powders had different photocatalytic activities, specifically the key points in understanding photocatalysis, and we will return
to oxidize hydrocarbons and alcohols, simultaneously producing to it later.
hydrogen peroxide [43]. Interestingly, these authors found that During the late 1960s, one of the present authors, at the
anatase powders were more active than the rutile ones. University of Tokyo, began to study the photoelectrochemistry of
In further work at the University of Edinburgh, McLintock and titania and found that oxygen gas was evolved at potentials very
Ritchie, using gas-phase adsorption measurements, studied the much shifted from the thermodynamic expectation, for example,
photocatalytic oxidation of ethylene and propylene at TiO2 [44]. with an onset of ca. −0.25 V vs. the standard hydrogen electrode
This study is one of the first that we have found that shows that (SHE), compared to the standard potential of +0.95 V in pH 4.7
it is possible to oxidize organic compounds completely to CO2 and aqueous buffer (Fig. 2.2) [50,51]. At first, there was skepticism of
H2 O: this result, but then it slowly became accepted. One reason that
C2 H4 + 3O2 → 2CO2 + 2H2 O. (2.10) this result was difficult to understand is that the photoexcitation
process converts the photon energy to chemical energy with little
The mechanism was proposed to involve the production of loss, and thus the photogenerated hole has a very high reactivity, so
superoxide from oxygen: that it can react directly with either water or quite robust organic
and inorganic compounds. Subsequently, a number of studies were
2 .
O2 + e− → O•− (2.11)
carried out in which the photoelectrochemical oxidation process
Markham et al. had already proposed this reaction to take place on TiO2 was examined for the competitive oxidation of water
on illuminated ZnO [31]. Work on similar photo-reactions has to O2 with the oxidation of a variety of inorganic and organic
continued into more recent years [45]. substrates [52,53]. Both types of reactions, of course, involve the
In an important study for the relationship between photoelec- use of light energy to get over an energy barrier, either an overall
trochemistry and photocatalysis, which we will come back to later, uphill process, as in the case of O2 evolution, or an overall downhill
in Section 3.2, Lohmann, at the Cyanamid European Research Insti- process, as in the case of organic oxidations.
tute, in 1966 published a highly detailed study of the photoelectro- With the report of the ability to simultaneously generate hydro-
chemical (PEC) behavior of ZnO, both in the presence and absence gen gas in 1972 (see Fig. 2.3) [19], the PEC field started to receive
of redox couples, including ferro/ferricyanide and methylene blue much wider attention, due to its implications for solar energy con-
[46]. He clearly showed that the overall current at the ZnO elec- version [54,55]. From this point, also, photoelectrochemistry be-
trode under illumination is the sum of anodic and cathodic cur- came closely associated with photocatalysis. We shall return to this
rents, the anodic current being a combination of the dissolution of topic later, in Section 3.2, and more carefully describe the detailed
the ZnO itself and the oxidation of any redox species present. The relationships.
cathodic process was the reduction of O2 to H2 O2 . This same ap- In this overview, we briefly mention some of the early work
proach had been introduced in 1938 by Wagner and Traud, at the of Bard and co-workers at the University of Texas. Frank and Bard
Technical University of Darmstadt, to help explain the corrosion of were the first ones to propose that illuminated TiO2 could be used
metals, coupled with either hydrogen evolution or oxygen reduc- for the purification of water via the photocatalytic decomposition
tion [47,48]. of pollutants [56,57]. They suggested that cyanide and sulfite could
Another PEC study that we will mention in this overview is that be photocatalytically oxidized to cyanate and sulfate, respectively.
of Morrison and Freund, of the Stanford Research Institute, who In one of these studies, they found that photocatalytic oxidations
also studied ZnO [49]. These authors also demonstrated in detail could also occur at other illuminated semiconductors, such as
the various situations that arise in the presence and absence of ZnO, CdS, Fe2 O3 and WO3 . The most active semiconductor was
redox couples. They also showed that oxidation products of some found to be ZnO [57]. These authors expanded this study to a
organic compounds are different in the case of the PEC electrode long list of inorganic and organic species [58] and speculated that
poised at the open circuit potential, i.e., with both oxidation and photocatalysis could be a useful approach to both environmental
reduction currents balanced, compared to the case of a purely cleanup and photo-assisted organic synthesis. The Bard group
electrochemical oxidation. This difference was proposed to be due also suggested that each small illuminated semiconductor particle
to the presence of cathodically generated superoxide. This is one of could be considered as a PEC cell, with both photo-assisted
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 519

Fig. 2.3. Photoelectrochemical cell used in the photolysis of water [19].


© 1972, Nature Publishing Group.

oxidation and dark reduction reactions taking place [59]. The Bard
group also proposed photocatalysis as a way to remove toxic
metals from wastewater [60]. Fig. 2.4. (a) Citations per year of the 1972 Nature paper: ‘‘Electrochemical
For a period of several years, the photocatalysis area continued photolysis of water at a semiconductor electrode’’ [19]; (b) Numbers of research
articles appearing on photocatalysis per year: search results in the period of
to expand as a technology for both the selective oxidation of 1972–2007 with the ‘‘Web of Science’’ (a) by the keyword ‘‘photocataly∗ ’’ (blue bars)
organic compounds [22] and the unselective oxidation of organic and (b) the keywords ‘‘TiO2 AND photocataly*’’ (green bars). (For interpretation
compounds for purposes of water purification [1,2,61–64] and, of the references to colour in this figure legend, the reader is referred to the web
version of this article.)
to some extent, also air purification [65–69]. There have also
been reviews and listings of references of work on both air
3. Properties of TiO2 materials
and water purification [4,70–72]. For these technologies, it is
typically necessary to use powerful ultraviolet (UV) light sources.
For passive purification, without special light sources, it became 3.1. Crystal structures
apparent in the early 1990s that the amount of light present in
As often described, there are three main types of TiO2
either natural sunlight or artificial light was insufficient to process
structures: rutile, anatase and brookite. The size dependence of
large amounts of organic compounds. Therefore, attention was
the stability of various TiO2 phases has recently been reported
turned to applications in which a relatively small number of UV [77,78]. Rutile is the most stable phase for particles above 35 nm
photons could be used to carry out reactions at the TiO2 surface, in size [77]. Anatase is the most stable phase for nanoparticles
for example, to decompose thin organic films on solid surfaces or below 11 nm. Brookite has been found to be the most stable
to kill bacteria on surfaces [5,6,73–76]. Thus, the focus turned from for nanoparticles in the 11–35 nm range, although the Grätzel
water purification to passive, self-cleaning, self-sterilizing solid group finds that anatase is the only phase obtained for their
surfaces, which, with sometimes only slight modification, could nanocrystalline samples [79,80]. These have different activities
also be used to purify air. For these types of applications, it was for photocatalytic reactions, as summarized later, but the precise
necessary to develop ways to coat various materials with TiO2 reasons for differing activities have not been elucidated in detail.
films. Such applications included the self-cleaning glass cover for Since most practical work has been carried out with either rutile
highway tunnel lamps, as well as a number of others, which have or anatase, we will focus more attention on these.
been reviewed previously and which will also be reviewed briefly Rutile has three main crystal faces, two that are quite low
later in this article. in energy and are thus considered to be important for practical
The large number of applications has also generated a renewed polycrystalline or powder materials [81]. These are: (110) and
(100) (Fig. 3.1a, b). The most thermally stable is (110), and
scientific interest in photocatalysis, and indeed on photo-assisted
therefore it has been the most studied. It has rows of bridging
reactions on semiconducting metal oxides in general. One of the
oxygens (connected to just two Ti atoms). The corresponding Ti
ways that we have tracked this activity is by looking at the
atoms are 6-coordinate. In contrast, there are rows of 5-coordinate
number of citations of the 1972 Nature paper on water photolysis Ti atoms running parallel to the rows of bridging oxygens and
(Fig. 2.4(a)). This number of yearly citations has been climbing alternating with these. As discussed later, the exposed Ti atoms
steadily over the past ten years or so and of course is correlated are low in electron density (Lewis acid sites). The (100) (Fig. 3.1b)
with the number of publications appearing on photocatalysis surface also has alternating rows of bridging oxygens and 5-
(Fig. 2.4(b)). coordinate Ti atoms, but these exist in a different geometric
520 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 3.1. Schematic representations of selected low-index faces of rutile: (a) (110); (b) (100); and (C) (001).

relationship with each other. The (001) face (Fig. 3.1c) is thermally excess titanium, such as the Magneli phases, Tn O2n−1 , where n
less stable, restructuring above 475 ◦ C [81]. There are double rows can range from 4 up to about 12 and the titanium oxide layered
of bridging oxygens alternating with single rows of exposed Ti compounds, in which there can be as much as several percent
atoms, which are of the equatorial type rather than the axial type. excess oxygen. The oxygen-deficient Magneli phases, which also
Anatase has two low energy surfaces, (101) and (001) exist for V, Nb, Mo, Re and W, have been known for many years
(Fig. 3.2a, b), which are common for natural crystals [80,82]. [88–91]. In these compounds, oxygen vacancies are ordered and
The (101) surface, which is the most prevalent face for anatase lead to the slippage of crystallographic planes with respect to
nanocrystals [79], is corrugated, also with alternating rows of each other; this leads to formation of planes in which, instead of
5-coordinate Ti atoms and bridging oxygen, which are at the corner or edge-shared TiO6 octahedra, there are now face-shared
edges of the corrugations. The (001) (Fig. 3.2b) surface is rather octahedra. Fig. 3.4 shows a schematic diagram of this situation. The
flat but can undergo a (1 × 4) reconstruction [82,83]. The (100) corresponding Ti atoms are then unusually close and can interact
surface is less common on typical nanocrystals but is observed electronically [92]. It has been found recently that laser ablation of
on rod-like anatase grown hydrothermally under basic conditions a TiO2 rutile target can produce Magneli-phase nanoparticles [93].
(Fig. 3.2c) [80]. This surface has double rows of 5-coordinate Ti There are also quite a number of layered titanate compounds
atoms alternating with double rows of bridging oxygens. It can in which there is an apparent excess of oxygen. For example, the
undergo a (1 × 2) reconstruction [84]. layered protonic titanate Hx Ti−2 x/4x/4 O4 • H2 O has been prepared
Recently, the brookite phase, which is rarer and more difficult and exfoliated into single sheets, termed ‘‘titania nanosheets’’.
to prepare, has also been studied as a photocatalyst (see later). The Fig. 3.5 shows (a) a diagram of the layered structure and (b) TEM
order of stability of the crystal faces is (010) < (110) < (100)
and AFM images of single sheets.
(Fig. 3.3) [85].
Recently also, the discovery of high-pressure phases of TiO2
was made [86]. These are expected to have smaller band- 3.2. Electronic properties
gaps but similar chemical characteristics [87]. Their existence
was theoretically predicted and then experimentally proven; It was reported in 1942 by Earle that rutile and anatase TiO2
specifically, a form of TiO2 with the cotunnite structure was in the form of powders are n-type semiconductors and that
prepared at high temperature and pressure and then quenched in the conductivity decreases with increasing O2 partial pressure at
liquid nitrogen. It is the hardest known oxide. temperatures above 600 ◦ C [94]. The effect of O2 was explained on
There are actually quite a variety of different structures for the basis of an equilibrium involving thermal release of O2 from the
compounds with compositions close to TiO2 , including those with lattice. We recognize today that this leads to the creation of Ti3+
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 521

Fig. 3.2. Schematic representations of selected low-index faces of anatase: (a) (101); (b) (100); and (C) (001).

sites, which are responsible for the electronic conductivity. The possible that the migration involved interstitial hydrogen. This is
activation energy for the electronic conductivity was found to be an ambiguity that has persisted for many years.
1.75 eV for unsintered rutile powder and 1.7 eV for sintered rutile Strong reduction of various types of samples was examined,
powder. No evidence for ionic conduction was found. Cronemeyer at various temperatures between 300 and 1150 ◦ C. The strong
and Gilleo reported in 1951 that rutile single crystals exhibit reduction turns the samples blue–black. The activation energy
a band-gap energy of 3.05 eV [95]. Absorption spectra were for electronic conduction had already been reported to be
reported for both normal and slightly reduced crystals. For the 0.07 eV at room temperature, to produce a conductivity of ca.
latter, the blue color was based on a very broad absorption that 1 −1 cm−1 . The conductivities were found to increase with
peaked at 1.8 µm. In the following year, Cronemeyer published increasing reduction time. A ceramic sheet sample heated in
a very extensive study of the electronic properties of single hydrogen at 800 ◦ C was found to experience a weight loss of
crystal rutile in which the preliminary findings were substantiated 0.1%, corresponding to a release of oxygen that would provide
[96]. Detailed photoconductivity measurements were made. Dark 3 × 1020 electrons cm−3 . Hall effect measurements showed a close
conductivity and photoconductivity measurements were also agreement between the numbers of carriers and those calculated
made on a slightly reduced sample (reduction in H2 at 600 ◦ C). on the basis of the weight loss, indicating that all of the electrons
Interestingly, there was found to be a marked hysteresis in the dark were electrically active.
conductivity when the sample was raised from room temperature Breckenridge and Hosler also published extensive work on the
to 250 ◦ C and then cooled back to room temperature. After cooling electrical properties of rutile [97]. The effective electron mass was
with a high applied electric field, the blue color was found to be found to be anomalously large, 30–100 times greater than that of
concentrated at the negative electrode; the author ascribed this to the free electron. These authors presented convincing arguments
movement of oxygen vacancies, but this is not confirmed. It is also that the source of electronic conductivity in rutile is Ti3+ , which
522 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 3.3. Schematic representation of the brookite structure (taken from Beltran
et al. [85]).
© 2006, American Chemical Society.

Fig. 3.5. Layered lepidocrocite-like protonic titanate: (a) schematic representation;


(b) AFM image (taken from Shibata et al. [633] and Sasaki [634], respectively).
© 2007, Royal Society of Chemistry; 2007, Ceramic Society of Japan.

the neutral (fully reduced) vacancy being the lowest, followed


by the singly reduced and finally the unreduced vacancy. The
observed optical absorption edge was proposed to correspond to
the transition between the valence band (based on O2− ) to the
neutral oxygen vacancy, which was considered to be a narrow
impurity band, with a high effective electron mass. The band at 3.67
eV above the valence band was proposed to be due to Ti4+ . The
observed temperature and oxygen partial pressure dependences
were fully explained by the equilibrium:

1
O2 − = O2 + O2v+ + 2e− . (3.1)
2
The electronic properties of rutile and anatase thin films were
studied by Tang et al. [98]. There were large differences in the
electronic conductivities of the two types of films after reduction
by heating in vacuum at either 400 or 450 ◦ C. The anatase films
became essentially metallic, with no change in conductivity with
temperature. The rutile films, in contrast, retained measurable
activation energies, 0.076 eV for 400 ◦ C and 0.06 eV for 450 ◦ C. The
difference in behavior was considered to be due to the following
properties for rutile: the average static dielectric constant of ca.
100, the effective electron mass of 20 m0 , and the donor state radius
Fig. 3.4. Schematic representation of the crystallographic shear process to form of ca. 2.6 Å. Since the latter is similar to the distance between Ti4+
Magneli phases from rutile (taken from Marezio et al. [632]). sites, there is little overlap between donor wave functions. Anatase
© 2000, Elsevier Science. has the following properties: static dielectric coefficient of ca. 30,
and reduced effective mass of ca. 1 m0 , based on an estimated
results from the loss of oxygen, which produces oxygen vacancies donor state radius of ca. 15 Å. Based on optical absorption spectra,
Ov . It was proposed that these vacancies (valence +2) can have the band-gap energies were estimated to be 3.0 eV for rutile and 3.2
0, 1 or 2 electrons associated with them, with distinct energies, eV for anatase. Forro et al. reported on the electronic properties of
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 523

high purity anatase single crystals and found an activation energy


for electronic conduction of 0.004 eV [99].
Recently, Hendry et al. pointed out the problem that the exact
nature of electron transport had not been solved, given the very
wide range of values of Hall mobilities (0.01–10 cm2 V−1 s−1 )
and polaron (electron + accompanying lattice distortion) effective
masses (8me –190me , where me is the free electron mass) [100].
Part of the problem might involve the presence or absence
of dopants, which were found to decrease the mobility. These
workers, using THz spectroscopy with undoped rutile single
crystals, based on Feynman’s analysis [101], found intermediate
size polarons and a mobility of ∼1 cm2 V−1 s−1 . Thus, they
concluded that, in TiO2 films such as those used in dye-sensitized Fig. 3.6. Schematic diagram of the cleavage of rutile along the (110) plane (taken
nanocrystalline solar cells, the main limiting factor might be from Diebold [106]).
interparticle contacts. However, this assumes a zero dopant level, © 2003, Elsevier Science.
which is unrealistic. Clearly, even now, further work needs to be
done to clarify this basic issue. of the bridging O. However, a more recent experimental study that
Other recent studies have also focused on the electronic used LEED found that the bridging O is pushed out by 0.12 Å, in
properties from the standpoint of the dye-sensitized solar cell contrast to the earlier experimental study [108]. New theoretical
application (see, e.g., work of Aduda et al. [102]). For example, the calculations were mostly in agreement with the LEED study, except
effect of the morphology of porous titania films on the electron that the bridging O position was almost unchanged from the
drift mobility was studied. Other studies have been focused on unrelaxed structure [109].
the gas-sensing properties, e.g., for hydrogen gas [103–105]. In The titania surface may undergo significant structural changes
nanostructured form, titania has very high sensitivity for H2 , when it is exposed to water. Onishi and co-workers have shown
increasing greatly in electronic conductivity in the presence of that single crystal rutile surfaces that have been prepared by
low concentrations, for example, an increase of three orders standard methods used for UHV can actually erode and roughen
of magnitude upon introduction of 1000 ppm H2 [103]. The when exposed to an aqueous electrolyte [110]. Subsequently,
mechanism proposed for the increased conductivity was thought Nakato and co-workers reported on a method by which atomically
to involve the adsorption of hydrogen on the titania surface, flat single crystal surfaces could be prepared that were stable in
rather than incorporation into the bulk. The platinum electrodes aqueous electrolyte [111]; this method involved etching in 20% HF,
that were used to contact the surface in this study were also followed by air-annealing at 600 ◦ C.
possibly involved, acting to dissociate the H2 molecule, so that H The X-ray crystal truncation rod (CTR) technique has been
atoms could be produced and adsorbed more easily. The subject of used by Zhang et al. to examine the rutile (110) surface in the
hydrogen interactions with titania has been well studied and will presence of pure water and of 1 molal RbOH aqueous solution
be treated in more detail in Section 3.5. [112]. Interestingly, the five-coordinate Ti, which had been found
to be significantly depressed in UHV, was found to be depressed
3.3. Surface structure studies to a much smaller degree in pure water (0.051 Å) and only slightly
depressed in 1 m Rb+ . This is because the terminal position, which
It is quite difficult to separate work that has been carried out is empty in UHV, is occupied by a water molecule in aqueous
on the surface structure of titania from work that has been carried solution and by a hydroxide ion in the alkaline Rb+ solution. The
out on surface science in general and also on surface chemistry. The six-coordinate Ti, which had been found to be pushed out in UHV,
latter two subjects will be taken up in the two sections that follow. was found to be depressed to a small degree (0.002 Å in water and
In this section, we will briefly treat the stoichiometric rutile and 0.019 in Rb+ solution). The bridging O, which had been previously
anatase surfaces. A detailed treatment has been given as part of been observed to be depressed in the X-ray study and pushed out
Diebold’s extensive review on the surface science [106]. in the LEED study, was found to be pushed out, by 0.004 Å in water
Diebold shows how rutile can be cleaved to produce the and 0.010 in Rb+ solution. All of the displacements were smaller
commonly shown (110) surface, which is the most stable rutile than those found in vacuum, which the authors propose to be due
surface (see Fig. 3.6). Ramamoorthy et al. carried out theoretical the fact that either water molecules or Rb+ ions occupy positions
calculations on the rutile structure and found the (110) surface that would be occupied in the bulk lattice. Further conclusions
to be the most stable, based on the fact that it has the least from this study will be discussed in Section 3.4, in which we treat
dangling bonds [81]. The structure shown in Fig. 3.1a is that of interactions of titania with water.
the unrelaxed bulk and is rather flat, but these authors predicted Other predictions from the Ramamoorthy work were in regard
that this structure should pucker slightly upon relaxation, with the to the relative stabilities of the other rutile single crystal surfaces.
five-fold-coordinated Ti atoms depressed by 0.32 a.u. (0.169 Å), The order of stability was found to be (110) > (100) >
and the bridging oxygens also depressed, by 0.15 a.u. (0.079 Å). (011) > (001). This calculation is strictly only valid for 0 K and
Vogtenhuber et al., using similar calculations, found that the five- is for vacuum. Based on the results of the X-ray CTR study for
fold Ti atoms were depressed by 0.180 Å, the bridging O atoms aqueous solution, this ordering might be modified slightly, since
by 0.156, the planar O atoms by 0.115 Å and the six-fold Ti atoms the stabilities are based in part on the presence of dangling bonds,
by 0.049 Å [107]. An experimental study that made use of surface which would of course not be present any longer in the presence
X-ray diffraction found that the five-coordinate Ti was depressed of water.
by 0.16 Å, while the six-coordinate Ti was pushed out by 0.12 Å,
and the bridging O was depressed by 0.27 Å. A total of seven 3.4. Surface chemical studies: Interactions with water
theoretical studies were compared with the experimental surface
X-ray diffraction results in the review of Diebold [106]. Most of The interactions of titania surfaces with water have been stud-
these studies have agreed on the depression of the five-coordinate ied extensively and have been reviewed [106,113]. The earlier
Ti, on the pushing out of the six-coordinate Ti and the depression work involved conventional surface science methods. The studies
524 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 3.7. High-resolution electron energy loss spectra of rutile (110) at various
dosages of water, starting from the clean surface at bottom (taken from Henderson
et al. [114]).
© 1996, Elsevier Science.

reported during the past five years on this subject have involved
both theoretical and experimental studies, the latter including a
large number of scanning tunneling microscopy studies. The inter-
actions with water are important to understand, because water,
either liquid or vapor, is almost always present in photocatalytic
reactions. These interactions are especially important for the later
discussion of the photo-induced hydrophilic effect.
Much of the work that has appeared over the past decade has
been targeted at the question of whether water is adsorbed molec-
ularly or dissociatively. Going back to one of the pioneering works
on this subject, Henderson reported a high-resolution electron en-
Fig. 3.8. Schematic diagram of a mixed molecular water-dissociated water
ergy loss spectroscopy (HREELS)-temperature-programmed des- monolayer on the rutile (110) surface (taken from Lindan [127]).
orption (TPD) study that concluded that the adsorption of water © 2003, American Institute of Physics.
on rutile (110) is molecular on the stoichiometric surface and dis-
sociative on the reduced surface, which is conventionally produced main question is whether or not this is true for the non-reduced,
by heat treatment, presumably forming oxygen defects [114]. The stoichiometric surface. Theoretical studies have been divided into
progression of HREELS spectra as a function of water coverage is those that predict molecular adsorption [116–122], those that
shown in Fig. 3.7. In the background spectrum at the bottom, there predict dissociative adsorption [123–129], and those that also find
is a very small peak at 3690 cm−1 . This is due to the O–H stretch stability for mixed molecular-dissociative adsorption [122,125–
for OH groups that are not hydrogen-bonded, often called ‘‘iso- 127]. This is a particularly difficult problem, since the energy
lated’’ OH groups. This vibrational frequency is close to that for differences are rather small. One of the studies that has predicted
OH groups that stick out from the surface of liquid water, without dissociative adsorption also predicts a mixed layer of molecular
being hydrogen-bonded to any neighbors, as observed with sum- and dissociated water at higher coverages [127]. Fig. 3.8 is a
frequency generation (SFG) spectroscopy [115]. At higher cover- schematic diagram taken from this paper that shows how the
ages, the peaks that appear are shifted to lower wavenumbers, mixed monolayer is arranged; the two structures both include
indicating hydrogen bonding. There is no longer any evidence of hydrogen bonding between a water molecule adsorbed at a five-
the high wavenumber peak, except at the highest coverage. There coordinate Ti site and an OH group adsorbed at an adjacent 5-
is also a peak that appears at 1605 cm−1 , which is due to the H–O–H coordinate Ti site. There is also a weak interaction of the water
bending mode of liquid water. Thus, it is certain that there are wa- molecule with the bridging oxygen. For reference, the diagrams
ter molecules adsorbed. However, it is not certain whether there for purely dissociative and purely molecular adsorption are shown.
are also dissociated water molecules present that are hydrogen Zhang and Lindan have also calculated a theoretical vibrational
bonded to neighboring water molecules or to bridging oxygens. spectrum, which we show along with one of the HREELS spectra
There is not much doubt that water dissociates at oxygen from Henderson’s work (Fig. 3.9). Even though the simulated
vacancies that are produced by heating in vacuum [113]. The spectrum is significantly shifted upward in wavenumber, the two
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 525

Fig. 3.10. A series of three sets of infrared spectra for (a) anatase [635] and (b) [636],
Fig. 3.9. (a) A HREELS spectrum for water adsorbed on rutile (111), taken from (c) [637]) rutile powders acquired at various temperatures and water coverages. In
Henderson [114]; (b), (c) and (d) show simulated vibrational spectra for the various (a), the water coverage decreases with spectrum number, and in (b) and (c), with
types of submonolayers and monolayers studied by Zhang and Lindan [127]. spectrum letter. In (b) and (c), the original spectra were obtained in the transmission
© 2003, American Institute of Physics. mode; all spectra have also been replotted with increasing wavenumber. In (b), the
main peaks are listed.
peaks have an appearance that is similar to the experimental © 1988, Royal Society of Chemistry; 1987, American Chemical Society; 1971, Royal
Society of Chemistry.
spectrum. The strong peak at high wavenumber is due to an almost
completely non-hydrogen-bonded OH, presumably the terminal
hydroxyl group, and the weaker, lower wavenumber peak is due to [113]. One of the interesting aspects is that Ti3+ sites by themselves
the OH group of the water molecule that is bonded to the hydroxyl do not have special reactivity; for example, such sites on rutile
group. (100) and on Ti2 O3 are not reactive. Only on the (110) surface are
This type of double-peak structure is rather commonly they reactive.
observed in experimental infrared spectra for both rutile and The background of the STM work has also been discussed
anatase powders (Fig. 3.10). The quite sharp peak or peaks at in the thorough reviews of Henderson [113] and Diebold [106].
high wavenumber are due to isolated OH groups, and the broader, For example, the latter discusses the problems of distinguishing
lower wavenumber peak or peaks are due to hydrogen-bonded OH between oxygen vacancies and hydroxyl groups that have been
groups. In all of the spectra shown, there is some fine structure. produced as a result of a water molecule reacting with an
Although not certain, this could be due to the existence of different oxygen vacancy. A number of authors concluded that the medium-
crystal faces, with slightly different geometries for adsorption. brightness spots that they observed in STM between bright rows
Thus, it appears likely that, at least on powders, with coverages were due to oxygen vacancies. Diebold et al. pointed out that there
on the order of a monolayer, there could be mixed monolayers. appeared to be two types of defects that were observable on rutile
Certainly, there is molecular water, and there must also be (110), which they termed ‘‘A’’ and ‘‘B’’ [131]. Between the rows
hydroxyl groups, with the OH group pointing up, normal to the of 5-coordinate Ti atoms, which appear bright due to their high
surface, so that there is little opportunity for hydrogen bonding. electron density, there are darker rows that are due to the bridging
However, for powders, there are, of course, a variety of crystal oxygens. The ‘‘A’’ type were observed to be significantly brighter
faces exposed, and distinct situations might be found on each. This than the ‘‘B’’ type and were proposed to be oxygen vacancies. The
is expected from the work of Henderson in a comparison of the A defects were removed by scanning the tip at a voltage of +3 V,
(110) and (100) surfaces [130]. The latter was found to support while the B type remained. The A-type defects were also found
dissociative, while the former was found to support molecular to be quite mobile. Suzuki et al. subsequently reported similar
adsorption. images and also found that the brighter spots were removable with
Direct evidence for molecular adsorption on stoichiometric a scan at +3 V [132]. These authors found that the spots were also
rutile (110) can also be found in STM work that has been targeted removable by electron-stimulated desorption. They were also able
at interactions of water with oxygen vacancies. This general topic to produce additional spots by dosing with atomic hydrogen. Thus,
will be discussed next. they proposed that the bright spots were due to hydroxyl groups
The background of the work on the interaction of water with formed by hydrogen adsorption on bridging oxygens. Brookes et al.
oxygen vacancies on rutile (110) has been given by Henderson also carried out STM measurements on rutile (110) and found that,
526 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 3.11. STM images of rutile (110) showing (a) oxygen vacancies and (b) bridging hydroxyl groups (taken from Wendt et al. [129]).
© 2005, Elsevier Science.

when they purposefully dosed the surface with water, it adsorbed


without dissociation at 150 K and then dissociated at 290 K.
Significantly, no terminal features were observed, i.e., there was
no evidence for oxygens or hydroxyls adsorbed on 5-coordinate Ti
sites, which would have resulted if the water had dissociated on the
stoichiometric surface. Features were only observed on the dark,
bridging oxygen rows. Thus, the authors concluded that the water
had dissociated at oxygen vacancies.
Schaub et al. carried out further work and found that there
were two types of ‘‘A’’ defects, a smaller type and a larger, brighter
one [117]. The latter was assigned to an oxygen vacancy and the
former to a hydroxyl group, based on theoretical calculations.
These authors continued to support the idea of water being
dissociatively adsorbed only at oxygen vacancies, in line with
their theoretical calculations. In further work from this group,
however, the assignments were modified [129]. In this work, it was
recognized that the surfaces that had been examined earlier were
mostly hydroxylated. Special care was taken to achieve extremely
low levels of background water, and thus it became clear that the
darker features were the actual oxygen vacancies and the medium- Fig. 3.12. STM images of rutile (110) showing the dissociation of a water molecule
bright features were individual hydroxyl groups that had been at an oxygen vacancy (taken from [133]).
formed via water dissociation (Fig. 3.11). © 2006, American Physical Society.
Subsequent work showed even more clearly how a water
molecule moves along a row of 5-coordinate Ti sites and then to move, whereas the one that is produced at an adjacent bridging
reacts with a vacancy, first producing a pair of hydroxyl groups oxygen due to the addition of a proton, is quite mobile. This result
on neighboring bridging oxygens [133]. A second water molecule appears to be consistent with the report of Wendt et al., in which
can then further catalyze the splitting of the hydroxyl pair in it was shown that proton can jump several rows away [133].
an energetic reaction that can result in one of the protons The implication is that the electrons associated with the original
jumping several rows away. The initial water-dissociation process vacancy are rather localized. All the theoretical results have not
is shown in Fig. 3.12. Both the dissociation and splitting processes been in agreement with this picture.
are also available as movies [134,135]. This paper also corrects The work just described on the interactions of water with
the assignments that had been given in work focused on the vacancies on rutile (110) has a more general implication, in
interaction of oxygen (O2 ) with oxygen defects [136]. addition to the obvious one. Some of the studies that have been
A paper by Bikondoa et al. appeared in early 2006 [137], carried out over the years that have discussed reactions of oxygen
apparently written without the knowledge of the one by Wendt vacancies may have been actually dealing with bridging hydroxyl
et al., which appeared in 2005 [129]; this paper also clearly showed groups. Henderson had already pointed out this effect in 1996
the whole situation regarding previously published assignments [114].
by the various groups. These authors can be credited with having We note also the work of Mezhenny et al., which was focused on
recognized, from the beginning, the fact that the bright spots the question of whether or not UV light produces oxygen vacancies
that were being observed by various groups were in fact due to on the rutile (110) surface [139]. This work showed little effect
hydroxyl groups, either single or double, that had been formed via of ordinary intensity levels of UV light, i.e., similar to those that
water reaction at oxygen vacancies. are present in sunlight, in producing oxygen vacancies on the
Additional work has recently appeared on the STM observation surface. It is likely that this work might have also suffered from
of the reaction of water molecules with oxygen vacancies. Zhang unrecognized background levels of water, since they report STM
et al. reported that the two bridging OH groups that are produced images that are characterized by the brighter spots that have been
from the reaction are actually not identical [138]. The one that is assigned by later workers to bridging hydroxyls. Nevertheless, if
produced at the site of the original oxygen vacancy is not observed oxygen vacancies had indeed been produced, there would have
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 527

been an increase in the number of such hydroxyls, whereas


no increase was observed. This result is in contrast to earlier
results obtained with second harmonic generation and X-ray
photoelectron spectroscopy that concluded that large numbers of
defects were in fact produced and were reactive toward dioxygen
[140]. These issues will be discussed at greater length in the next
section and in Section 4, which deals with the photo-induced
hydrophilic effect.
Work on the interaction of water with titania surfaces has also
been carried out with other techniques. For example, anatase pow-
der was examined with gas-chromatography–mass spectrometry
(GC–MS) and quantum chemical calculations. Experimental evi-
dence was found for water dissociation, which was consistent with
the theoretical calculations; the latter showed that the adsorption
is molecular on the anatase (101) surface and dissociative on the
(001) surface.
The previously mentioned X-ray CTR study of Zhang et al.
provides rather clear evidence for molecular adsorption for pure
water on the stoichiometric rutile (110) surface and dissociative
adsorption in alkaline aqueous solution (pH 12) [112].

3.5. Surface chemical studies: Interactions with dioxygen and other


species

It was realized in 1998 that dioxygen does not only react with
oxygen vacancies (or possibly, as discussed above, bridging hy-
droxyls) to produce a near-stoichiometric surface at temperatures
above 600 K, but also, at temperatures below 600 K, it can leave be-
hind an oxygen atom adsorbed at a 5-coordinate Ti site [141]. This
realization led to doubts concerning previously published work
that had found dissociative water adsorption at rutile (110). It also
led to a reassessment of what had been an accepted procedure for Fig. 3.13. Schematic diagram of an O2 molecule reacting at an oxygen vacancy on
the preparation of high quality, clean surfaces. The scheme that rutile (110) and dissociating, with further reaction with a water molecule (taken
Epling et al. proposed to explain the interactions of O2 with oxygen from Epling et al. [141]).
vacancies and subsequent reaction with water is shown in Fig. 3.13. © 1998, Elsevier Science.
In the same paper, these authors also found evidence to support a
similar end product that resulted when water was present initially, We include here a brief mention of surface photochemical
so that bridging hydroxyls had already been formed. reactions involving O2 . The work of Thompson and Yates has
Henderson et al. reported later that O2 can adsorb at a reduced, employed the photodesorption of O2 as a means of monitoring
i.e., vacancy-containing, rutile (110) surface without dissociation the arrival of photogenerated holes to the surface of a rutile single
at temperatures below 150 K [142]. One of the more interesting crystal with exposed (110) face [145]. This process is essentially
aspects was the observation that O2 can adsorb, probably as O•− 2 ,
the reverse of the photo-adsorption process just alluded to, which
at a ratio of up to three molecules per oxygen vacancy, which requires a trapped electron, creating a partially or fully reduced
necessarily means that it does not have to interact directly with O2 , i.e., O•−
2 . The presence of methanol as a hole trapping agent
the vacancy but can reside on an adjacent cation site. Another significantly decreased the photodesorption. In further work by the
paper from the same group appeared more recently exploring the same authors, they proposed a fractal rate law to fit the observed
reaction of O2 with bridging hydroxyl groups in more detail [143]. kinetics of the reaction of trapped electrons with trapped holes
These authors conclude that the role played by O2 in photocatalysis [146]. It was assumed that the electrons were associated with
involves specifically this reaction. They also found, in agreement oxygen vacancies, but this picture may be in some doubt, based on
with their earlier work, that a second monolayer of water blocks the ability of trace water to convert these to bridging hydroxyls.
the access of O2 to the bridging OH groups, effectively impeding the
electron transfer. On the basis of these results and other studies in 3.6. Bulk chemistry—Hydrogen
which superoxide was generated both on thermally reduced titania
and on UV-illuminated titania, the authors proposed that bridging In this section, we briefly review the literature on the
hydroxyls are a key intermediate in the photocatalytic process. bulk chemistry of titania. This subject is mostly limited to the
We agree with this proposal and also propose (see later) that such incorporation of elemental hydrogen. It also can include the
bridging hydroxyls can be generated electrochemically. incorporation of lithium or sodium, but this is beyond the present
The effect of gas-phase O2 on nanocrystalline titania films has scope. The subject of hydrogen incorporation can also include
been studied in terms of the gas-sensing application [144]. It electrochemically-induced processes; these will be treated in the
was found that the film conductivity decreased in the presence next section. The characteristics of hydrogen as a bulk impurity
of O2 . This effect could also be related to the effect discussed in titania are central to the understanding of its electrical,
above but is more likely to involve the scavenging of bulk trapped electrochemical and photoelectrochemical behavior, which is, in
electrons, as discussed in the next section. The authors also turn central to the understanding of the photocatalytic behavior.
observed photo-induced adsorption of O2 , a phenomenon that had An early paper on the optical and infrared absorption spectra
been described by Kennedy et al. in 1958, as mentioned in the of rutile single crystals by Soffer showed evidence for the
historical overview [40]. incorporation of H, already present in the as-received crystal,
528 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 3.14. (a) H binding site within the rutile structure and (b) proposed diffusion path within the solid structure, along a c-channel (taken from Bates and Perkins [154]).
© 1979, American Physical Society.

and D, which was introduced by heating at 900 ◦ C in D2 ; these compared with 8 × 10−8 cm2 s−1 at 698 ◦ C along the a-axis. The
were evidenced by the appearance of IR bands at 3277 (main) diffusion coefficient estimated for room temperature was 1.8 ×
and 3322 cm−1 due to an O–H stretch and at 2442 cm−1 for the 10−13 cm2 s−1 , which was in reasonable agreement with the value
corresponding O–D stretch [147]. The author remarked that the very roughly estimated by Chester and Bradhurst, in the range
bands are unusually narrow for a solid-state O–H stretch and also 10−11 –10−13 cm2 s−1 , based on electrochemical insertion.
discussed the possibility of a hydrogen-bonding-related shift. Bates and Perkins measured the infrared frequencies for H, D
Prior to the work of Hill in 1968 [148], a number of different and T in rutile TiO2 and carried out a detailed structural analysis,
studies had suggested that non-stoichiometry in rutile TiO2 comparing the results with theory for anharmonic oscillators and
was associated with increased electrical conductivity, but the also with that of hydrogen bonding [153]. The agreement with the
mechanism was not clear. It had also been suggested that hydrogen latter was poor. Bates et al. later published a much more detailed
acts as a dopant in rutile. In Hill’s work, heating rutile crystals study, including a review of the literature up to 1979 [154]. They
in hydrogen below 600 ◦ C led to increases in the bulk hydrogen carried out a detailed analysis of the mechanism of diffusion of H in
concentration, measured by IR. Heating above 650 ◦ C in vacuum rutile. The binding site within the lattice for the proton is shown in
led to decreases in hydrogen concentration, coupled with oxygen Fig. 3.14a, and the proposed path for diffusion in Fig. 3.14b. The
loss to produce water. This led to increased conductivity, probably binding site was later confirmed by Klauer and Wöhlecke using
due to the formation of faults involving Magneli phases. For the polarized Raman [155]. The understanding of this system achieved
heated crystals, the electrical behavior was modeled involving in this work is excellent. It was concluded that the wavenumber
a series network of two parallel RC circuits, with one being shift of the IR absorption was not due to hydrogen bonding, which
associated with an ‘‘exhaustion layer’’, i.e. either a depletion layer is consistent with the observed sharpness of the band. Instead, it
or layer that is very low in carriers, as discussed in the next section. was proposed to be due to the electrostatic environment within
Johnson et al. reported further detailed work on the optical and the lattice (however, see below). Further work was also published
infrared spectra for H and D incorporated in rutile single crystals on tritium diffusion [156].
[149]. This paper referred to the earlier work of von Hippel et al. Peacock and Robertson have carried out quantum chemical
that provided presumably more accurate values for the absorption calculations for H in a variety of oxides that are considered as
maxima: 3276 and 3317 cm−1 for O–H and 2435 and 2463 cm−1 high dielectric constant oxide gate materials [157,158]. They find
for O–D. That paper had proposed that the peak splitting was due that the H0 energy level lies above the conduction band in ZnO,
to slightly differing O–Ti distances. Johnson et al., however, denied TiO2 and SrTiO3 , consistent with the fact that H is a shallow
this possibility due to the symmetry of the structure. H and D donor in all the three. Work of Park et al. also confirmed these
doping was carried out by heating in an atmosphere of H2 O or results for rutile [159]. Koudriachova et al. have also carried out a
D2 O, plus O2 at 850 ◦ C, or in some cases, H2 or D2 below 550 ◦ C. recent ab initio quantum chemical calculation on H incorporation
Conduction-band electrons produced a broad absorption band at in rutile [160]. These authors sought to recheck the binding site,
1.5 µm. This was remarked to not be due to conventional free due to the difficulty already mentioned, i.e., the O–H–O bond
electron behavior, which should produce no peak. These authors distances were not consistent with established rules relating them
carried out a detailed analysis of the various possible binding to vibrational frequencies. The new calculation found a distortion
sites for H or D within the crystal. The same group reported the in the cage surrounding the H atom such that the distances
use of the IR absorption band in the precise determination of H became highly consistent with the correlation. They found that
and D concentrations in rutile [150]. DeFord and Johnson studied the H atoms are most favorably located at ordered positions, as
the H/rutile system in detail from the viewpoint of theoretical shown in Fig. 3.15a. The lattice expands linearly with increasing
semiconductor and thermodynamic properties [151]. Later, they H incorporation (Fig. 3.15b).
made measurements of H and D diffusion using the isotope It is also appropriate to mention here theoretical calculations
exchange technique in order to avoid internal electric fields [152]. that were carried out on the surface adsorption of hydrogen as
The diffusion was carried out simply by heating the samples in the H2 [161]. In that work, it was found that up to one monolayer is
appropriate atmosphere (see above) for various times and then adsorbed, with all of the bridging oxygens becoming hydroxylated
measuring the H or D concentrations via the IR absorption. The and the underlying Ti4+ ions being reduced to Ti3+ . We might note
diffusion coefficients for H varied from ca. 3 × 10−8 cm2 s−1 at that this type of surface could in principle be produced thermally
350 ◦ C to ca. 1.7 × 10−6 cm2 s−1 at 700 ◦ C along the c-axis, by removing half of the bridging oxygens, followed by exposure to
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 529

Panayotov and Yates recently reported on experiments in


which they reduced titania pellet samples (Degussa P25) with
a source of H atoms [164]. They were able to observe the
broad, featureless background visible-IR spectrum expected for
conduction-band electrons. In addition, there were increases in
electronic conductivity. However, there was no discernable IR
absorption at ca. 3280 cm−1 for the internally bound O–H stretch
already discussed above. The activation energy for the diffusion of
H into the solid was estimated to be 0.09 eV.

3.7. Electrochemical properties

The intrinsic electrochemistry of TiO2 has been studied


continuously for a long period, since the first report of Boddy in
1968 on the oxygen evolution reaction [165]. This work made use
of single crystal rutile electrodes. Interestingly, this work contains
a figure in which the photocurrent vs. potential behavior is given,
one year prior to the reports of one of the present authors [50,166].
This will be discussed further, in the next section.
Many aspects of the behavior of TiO2 can be explained on the
basis of a semiconductor model, as already discussed. One of the
ways of characterizing a semiconductor is to measure its flat-band
potential electrochemically, for example, with capacitance. In our
original work reported in 1969, we described such measurements
and concluded that the flat-band potential EFB for rutile (001) is
pH-dependent, with a relationship similar to the following [167]:

TiO(O) + H+ + e− = TiO(OH) (3.2)


with the resulting Nernst relationship:

F
EFB = 0.00 + ln[H+ ] = −0.0591pH at 25 ◦ C. (3.3)
RT
This result has essentially been confirmed by subsequent workers,
with the value at pH 0 being +0.01 ± 0.05 V vs. SHE for rutile
(001) [169,171,175]. The pH dependence is typical of the behavior
of most oxide semiconductors and has generally been considered
to be due to a surface acid–base equilibrium for these oxides. The
value for anatase is more negative: −0.20 V vs. SHE [168].
There has been a certain amount of discussion devoted to the
question of exactly how the capacitance measurements should be
Fig. 3.15. (a) Illustration of a high stability ordered arrangement of interstitial conducted and how the results should be interpreted, even for
hydrogen atoms in the rutile structure, with stoichiometry H1/4 TiO2 (taken from single crystals. This discussion is interesting, because it displays
Koudriachova et al. [160]). (b) Plot of lattice volume for various numbers of the convergence of an ideal, simple theoretical model with a real,
interstitial H (open circles) and Li atoms (filled circles) (taken from Koudriachova
non-ideal complicated material. The model, already discussed,
et al. [160]).
© 1994, American Physical Society. involves an ideal semiconductor with a space charge region whose
thickness is dependent upon the potential difference between the
water. It could also be produced in principle electrochemically (see Fermi level, to which we can assign an electrochemical equivalent
Section 3.7). EFL and the conduction-band-edge energy, again given on the
It has been found that titania nanotubes respond to the electrochemical scale, ECB . The energy difference ξ between ECB and
presence of hydrogen in the gas phase, as already discussed in EFL deep within the bulk of the material is dependent upon the
carrier concentration. The space charge capacitance CSC based on
the section on electronic properties. In that work, it was not
this simple model is given by
considered that hydrogen could actually be absorbed. However,
work of Lim et al. showed that for nanotubes that were prepared 1 2(E − EFB )
hydrothermally, there was a reversible uptake of ca. 2% [162]. The = (3.4)
2
CSC εε0 eND
incorporation led to an increase in the IR absorption (3427 cm−1 ),
which is significantly higher than that for single crystal rutile. Only where ND is the bulk concentration of donors, with the conse-
75% of the uptake was reversible at room temperature with the quence that a plot (Mott–Schottky) of C −2 vs. potential yields the
remaining 25% requiring temperatures up to 130 ◦ C to desorb. carrier concentration from the slope and the flat-band potential
A relatively detailed study has been carried out on small from the intercept with the potential axis. Many workers have
rutile crystals by the use of optical and IR absorption and Raman found that these plots are either non-linear or that the intercept
scattering [163]. The effect of neutron irradiation was also studied. gives a result that is not reasonable, for example, more negative
The incorporation of H in minerals is of interest to geologists, than that given in Eq. (3.3). Such results have been explained in
because it affects the macroscopic properties and can be a way various ways, including (1) non-uniform depth profile of carriers
by which water is incorporated in minerals that normally do not and (2) deep trap levels. There appears to be some consensus on
absorb water. the merits of the first explanation.
530 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Finklea has given an excellent summary of the practical


methods of obtaining linear Mott–Schottky plots with correct
intercepts [175]. The opposite situation can also arise if a surface
layer exists with a higher concentration of carriers compared to
the bulk. This leads to a shallow slope near the flat-band potential,
followed by a steeper slope at more positive potentials.
In certain cases, if the carrier concentration is sufficiently
high, the space charge capacitance can approach that of the
electrical double layer, and the Helmholtz capacitance must also
be considered.
It can be argued in general that the behavior of TiO2 is simply
not ideal, and, in each individual case, a full impedance treatment
is necessary to extract the space charge capacitance or even to
evaluate whether or not it exists. Similar situations also exist for
oxide films; the situation is particularly complicated for oxide films
Fig. 3.16. Mott–Schottky plots for a rutile (001) surface in pH 4.7 buffer: (a) raw grown on metals, for example, iron and zirconium [176].
data for a more lightly doped sample (circles), with a solid line showing the fit For certain types of TiO2 electrodes, it may be doubtful that the
to Eq. (3.5) (including both space charge CSC and passive layer CPL contributions); behavior can be strictly described with a semiconductor model.
(b) calculated intrinsic CSC behavior, after removing the effect of CPL ; (c) raw data For example, the Mott–Schottky plots may exhibit a significant
for a more heavily doped sample (squares), with a solid line showing the fit to
Eq. (3.5); (d) calculated intrinsic CSC behavior, after removing the effect of CPL . SCE
frequency dispersion, either with or without non-linearities. This
refers to the saturated calomel electrode, which is 0.241 V vs. the standard hydrogen can signal the fact that a different type of model might be
electrode (SHE) (based on [167]). more appropriate. In most cases, it is advisable to examine the
full impedance spectrum over a range of potentials in order
A model that can explain the appearance of a steep slope near to determine whether or not the behavior does in fact follow
the intercept and a shallower slope at more positive potentials that expected for a semiconductor. One specific type of behavior
is that of a thin layer near the surface that has a lower carrier that has been observed is that of an electrochromic film, in
concentration. This model is reasonable, because, as has been which all of the charge injected into the film is associated with
reported often, the effect can arise as a result of oxidizing etching increased absorption of visible light (see Refs. [177–181] and later
treatments, which could remove electrons from such a surface discussion).
layer, especially if the treatment is conducted for a relatively The electrochemical impedance spectral (EIS) behavior can
short time. In this case, the overall capacitance behaves as if be quite powerful in elucidating the behavior of electroactive
there is a smaller, potential-independent capacitance due to the materials, i.e., those that can undergo electron transfer reactions
passive layer CPL , in series with the potential-dependent space within pores and even within the solid material itself. Nogami
charge capacitance. The slope of the upper portion of the curve examined the EIS behavior of several single crystal rutile samples
is somewhat increased from that which is characteristic of the and obtained sets of straight lines, with varying slopes in Bode
carrier concentration in the bulk of the material, and the intercept amplitude (log |Z | vs. log frequency) plots [182]. He explained
is shifted to the negative. The overall behavior is described by this behavior on the basis of a disordered layer on the surface
1/2  of the single crystal. The Bode plots show straight lines for the
2(E − EFB ) 2(E − EFB )
   
1 1 1
= + + (3.5) logarithm of the magnitude of the impedance (absolute value of the
2
CTOT εε0 eND εε0 eND CPL 2
CPL complex impedance) vs. the logarithm of the frequency. At more
which is derived by substituting positive potentials, the slopes of the lines are close to 0.7, while,
at more negative potentials, the slopes are close to 0.5. For a pure
1 1 1 capacitance, the slope should be 1.0, while, for a pure resistance,
= + (3.6)
CTOT CSC CPL it should be 0.0. For intermediate cases, there are both resistance
into Eq. (3.4). In our original paper, we reported curves with this and capacitance distributed through the material. The simplest
type of double slope (see Fig. 3.16) and made a mathematical electrical case is that of the uniform semi-infinite transmission
correction of this type in order to estimate the true flat-band line, in which there is a ‘‘ladder’’ of constant resistances and
potential [167]. Our explanation at that time invoked a Helmholtz capacitances (Ref. [183] and references therein). This type of
layer capacitance, which was later correctly disputed by Dutoit behavior has been shown to be followed by a porous medium
et al. [169]. Based on our nitric acid etching procedure used with cylindrical pores. A rate-limiting diffusion process in which an
in that work, however, it is quite reasonable to invoke the electroactive species diffuses through a uniform medium behaves
passive layer model just described. DeGryse et al. also found fault electrically in the same way, and thus there is some ambiguity as to
with the involvement of the Helmholtz layer and proposed an which process is actually operating. In either case, the slope should
inhomogeneous carrier concentration [170]. Tomkiewicz proposed be 0.5, as observed by Nogami for more negative potentials.
a model based on surface states with energies in the middle of For the general case, in which the slope of the Bode plot can
the band-gap [171]; this model was criticized by Ullman, who have an arbitrary value between 0.5 and 1, the same type of
proposed a surface passive layer [172]. RC ladder is also valid, and the physical picture is also similar
It should be noted that this same model has been discussed by to that for the uniform network, but with pores that are non-
Schoonman et al. [173], but the equation given in that paper was cylindrical. For example, Wang and Bates have shown that horn-
different, in that the cross-term was not included, leading to the shaped pores can give rise to Bode slopes in this range [184]. Thus,
erroneous result that the curve is simply raised but has the same the results of Nogami could be explained by a situation in which
slope. If the Mott–Schottky behavior does not fit this simple model, horn-shaped pores of larger diameter (lower roughness factor)
it is still possible that it can be fitted with a more complicated exist at higher potentials; at intermediate potentials, the diameters
model, in which the slope at any given point in the curve can become smaller (higher roughness factor); finally, at the most
be related to the carrier concentration at a certain depth; thus a negative potentials, the pores behave as if they are cylindrical and
complete concentration profile can be estimated [174]. longer than the penetration length of even the lowest frequencies
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 531

Fig. 3.17. Electrochemical impedance spectrum for a nanocrystalline anatase film on a conductive support: (upper left) experimental results from Cao et al. [187]; (upper
right, lower right) simulated log of the impedance amplitude vs. log frequency and phase angle vs. log frequency, respectively.
© 1995, American Chemical Society.

measured. It should be noted that impedances with arbitrary compared to the experimental result shows that there is some va-
slopes are often referred to as constant phase elements, meaning lidity for this model. The higher frequency capacitive behavior,
simply that the phase angle, which is 0◦ for a pure resistance, 90◦ with a capacitance of ca. 10 µF cm−2 , is typical for the flat surface
for a pure capacitance and 45◦ for the uniform transmission line, is of the film. The lower frequency capacitance (ca. 25 µF cm−2 ) is
constant but is not one of these standard ones. In any case, it seems proposed to involve a bulk reduction of Ti4+ to Ti3+ , as discussed
clear that Nogami’s conclusion, i.e., that the frequency dispersion of below. The diffusion-like process might involve either actual diffu-
the Mott–Schottky plots is due to the presence of a disordered layer sion in pores or combined gradients of resistance and capacitance,
on the single crystal surface, might be valid. Frequency dispersion or even a combination.
is a result of the mixing of resistive (i.e., electronically conducting) Muñoz et al. studied the impedance behavior of titania
character with capacitive (i.e., ionically conducting) character. nanotube arrays using EIS and also found behavior with similar
Furthermore, his proposal that the behavior involves surface aspects [188]. The nanotubes were prepared via anodic oxidation
states might also be valid. Wang also has shown that constant- of Ti metal; the titania produced in this way is amorphous
phase angles between 90◦ and 45◦ can be obtained if the and can be converted to anatase with heat treatment. The
resistances and capacitances in the ladder are non-uniform, impedance of both forms was measured. For the amorphous films
e.g., with the resistance increasing and the capacitance decreasing at intermediate potential (0.0 V vs. Hg/Hg2 SO4 ), there were two
in a logarithmic manner [183]. Even a 45◦ angle can result with capacitive regions, as in the work described above, with a transition
a non-uniform situation. A non-linear gradient of resistance and region at intermediate frequencies that could be attributed to a
capacitance could well result if there is a similar gradient of Warburg impedance. At higher potential (+1.0 V), the impedance
concentration of Ti3+ sites within the film. Cahan and Chen have was almost purely capacitive, and, at lower potential (−0.5 V),
discussed such gradients in the passive oxide film on iron. They it was capacitive at lower frequencies, with a relatively low
have also pointed out that these gradients can naturally give rise impedance (higher capacitance) and predominantly resistive at
to linear log current vs. potential behavior (linear Tafel slope; see higher frequencies. The capacitive components are discussed in
below) that mimics that for rate-limiting electron transfer [185, terms of a space charge capacitance due to the oxide film at the
186]. base of the nanotubes, the film being in contact with the metal,
Cao et al. have reported impedance results for a nanocrystalline and a Helmholtz (double layer) capacitance of the pore walls.
anatase TiO2 film at a single, intermediate potential (0.24 V vs. SHE) Interestingly, for the annealed compact film, without nanotubes,
[187]. In this case, the Bode plot exhibited three regions: capacitive the impedance also exhibited two distinct capacitive regions, just
behavior at high frequency, diffusive behavior at intermediate fre- as found by Cao et al., with little dependence upon potential. In
quencies, and capacitive behavior at lower frequencies (Fig. 3.17). contrast, after annealing, the spectra for the nanotube films were
This behavior can be simulated approximately by the equivalent almost entirely capacitive, with a single component.
circuit shown, with the resulting Bode plot shown on the right. This general behavior is similar to that for other oxides of
Here, we have not attempted a perfect fit, which would have re- transition metals of the early transition series, e.g., MoO3 and
quired the introduction of constant-phase elements to simulate WO3 . Our own work with amorphous WO3 films clearly showed
the non-ideal capacitive behavior, which results in phase angles frequency regions in which the behavior was more capacitive
considerably below 90◦ . The similar shape of the calculated plot and regions in which it was more diffusive [189,190]. There
532 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

even appeared to be two types of diffusion, which we attributed


to diffusion in pores and diffusion within particles. In recent
work on MoO3 films, impedance results also showed diffusive
behavior at intermediate frequencies and capacitive behavior at
lower frequencies; the latter was attributed to an electrochemical
reduction process [191]. With impedance, in which the potential
variation is usually only a few millivolts, an electrochemical
process can behave exactly like a capacitance and is often termed
as pseudocapacitance.
The use of impedance measurements can clearly yield much
information about TiO2 photocatalytic films. It is clear from a
comparison of the reported impedance results that more work
is needed in order to fully elucidate the physical significance.
Further studies could establish the relationships between the film
nanostructure and the electrochemical behavior.
The electrochemical behavior of TiO2 in the more negative
region has been studied by several groups with the more
commonly used cyclic voltammetric (CV) technique, on single
crystal rutile [192], single crystal anatase [168] and nanocrystalline
anatase [187,193]. In all cases, there is a distinctive feature that
involves a cathodic voltammetric peak (reduction reaction) as
the potential is swept negatively, followed by an anodic peak
Fig. 3.18. Compendium of Tafel plots for oxygen reduction at various types of TiO2
(anodic reaction) as the potential is swept back in the positive
electrodes, with curves at the left in alkaline solution (pH 13 or 14) and curves at the
direction. Similar voltammetric features have been obtained for right for acid solution (pH 0): (a) anodized Ti in 1 M NaOH (O2 ) [203]; (b) anodized
lithium ion-containing non-aqueous electrolytes and thus have Ti in 0.05 M H2 SO4 (O2 ) [203]; (c) rutile (001) in 1 M KOH (air) [192]; (d) rutile (001)
been attributed to the intercalation of lithium [194,195]: in 1 M NaOH (air) [204]; (e) anatase film in 0.1 M NaOH (O2 ) [206]; (f) rutile (001) in
0.1 M NaOH (O2 ) [206]; (g) anatase (001) in 0.1 M NaOH (O2 ) [206]; (h) rutile (001)
TiO2 + xLi+ + xe− → Lix TiO2 . (3.7) in 1 M NaOH (O2 ) [205]. In cases in which air was used (curves c and d), the curves
have been multiplied by a factor of 5 to make them more directly comparable with
The proton counterpart has the same form and has been argued to those in which pure O2 was used.
also involve intercalation (see below):
H0 = −8 to H− = +23, i.e., 31 log units of proton activity
TiO2 + xH+ + xe− → Hx TiO2 . (3.8) or pH. Quartz crystal microbalance (QCMB) measurements were
The latter reaction has been associated with the filling of electron also carried out from H0 = −6 to pH = 12. In fact, the QCMB
trap sites, with an elevation of the Fermi level. These changes can measurements were consistent with the mass corresponding to the
be followed spectroscopically, by an increase in the absorption of uptake of a proton, or to a deuteron in the case of D2 O solution. It
light in the wavelength region from 380 to 600 nm. However, there was argued that only intercalation (Eq. (3.8)), rather than surface
has been some controversy involved with this idea. Some workers protonation–deprotonation (Eq. (3.2)) can account for the variation
conclude that the coloring of TiO2 films occurs as a result of the of ECB with proton activity and mass changes.
filling of the conduction band, with the absorption of light exciting The electrochemical insertion of protons was used later by
electrons from lower to higher energy levels within the CB. This Pelouchova et al. as a means of controlling the doping level
is a more physical view, which has been advanced by Fitzmaurice of anatase single crystals and the electrochemical behavior of
and others [196–201]. Other workers have concluded that the the electrode in the presence of a redox couple, in this case,
coloring process does indeed involve reaction (3.8) and that the methylviologen [180]. Specifically, the electrochemical doping
absorption of light involves electronic transitions associated with increased the rate of electron transfer to the oxidized form of the
the Ti3+ ion. This is a more chemical view, which has been methylviologen. There have been no other results reported thus far
advanced by Meyer and co-workers [187]. It has been difficult for other redox couples, but similar results could be expected.
to conclude which is correct, because the absorption spectrum The behavior of the oxygen reduction reaction (ORR) on TiO2
includes aspects that can be explained in both ways. Specifically, electrodes as a dark reaction has been examined by several groups
if the electrons are not trapped at specific sites, the absorption over the past 30 years: Clechet et al. used an anodized Ti electrode
should exhibit a steadily increasing absorbance with increasing and made measurements in both acid and alkaline solutions [203];
wavelength, as was observed by Panayotov and Yates, as discussed Parkinson et al. [192], Kesselman et al. [204], Tafalla and Salvador
earlier [164]. This is because there are many, closely spaced energy [205], and Tsujiko et al. [206] all used rutile single crystal (001)
levels that are available, with the probability being larger to absorb electrodes, and Tsujiko et al. also made measurements with an
a smaller amount of energy. If, on the other hand, the electrons anatase (001) single crystal [206]. Most of these studies used
alkaline solution but also examined the behavior at other pH
are trapped at specific, relatively well-defined sites, there should
values. In Fig. 3.18, all curves except one (b) are shown for alkaline
be specific, widely spaced energy levels, which would lead to
solution, either pH 13 or 14, while curve b is for pH 0.
absorbance peaks. Cao et al. argue that, since there is a broad peak
Clechet measured oxygen reduction of anodized Ti in both
in the absorbance at ca. 1000 nm, which corresponds to a specific
acidic and basic solutions and found a large difference in the
absorption process for Ti3+ , the electrons are essentially trapped at
behavior [203] (Fig. 3.18, curves a and b). The most striking
these sites [187].
difference is the large difference in the slope. In basic solution, the
In situ reflectance measurements have been used by Lyon
inverse of the slope (known as the Tafel slope) is quite close to that
and Hupp to follow the variation of the conduction-band energy
expected for a one-electron transfer as the rate-determining step
ECB with pH over an extremely wide range for nanocrystalline
(rds), i.e.,
anatase films [202]. They used solutions with standard Hammett
acidity functions H0 and H− , from H0 = −10 to H− = +27. dE (ln 10)RT
These workers found that Nernstian behavior was observed from = = 0.1183 V decade-1 (3.9)
d log i α nF
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 533

Table 3.1
Exchange current densities (log i0 , A cm−2 ) and Tafel slopes (V decade−1 ) for the reduction of O2 to O•−
2 at various types of TiO2 electrodes.

Electrode type log i0 , A cm−2 Slope, V dec−1 Reference

Anodized Ti in 1 M NaOH (O2 ) −5.2 0.12 Clechet et al. [203]


Anodized Ti in 0.05 M H2 SO4 (O2 ) −3.4 0.40 Clechet et al. [203]
Rutile (001) in 1 M KOH (air) −5.7 0.12 Parkinson et al. [192]; currents were multiplied by a factor of 5.
Rutile (001) in 1 M NaOH (air) −10.2 0.06 Kesselman et al. [204]; currents were multiplied by a factor of 5.
Anatase film in 0.1 M NaOH (O2 ) −5.2 0.12 Tsujiko et al. [206]
Rutile (001) in 0.1 M NaOH (O2 ) −9.3 0.12 Tsujiko et al. [206]
Anatase (001) in 0.1 M NaOH (O2 ) −8.0 0.06 Tsujiko et al. [206]
Rutile (001) in 1 M NaOH (O2 ) −8.4 0.10 Tafalla et al. [205]

where i is the current density in A cm−2 , E is the potential,


α is the transmission coefficient (usually 0.5), F is the Faraday
constant (96 485 C mol−1 ), R is the gas constant (8.314 J K−1 mol−1 ,
and T is the absolute temperature. This slope is consistent with
the rds being superoxide production (Eq. (2.11)). The standard
potential for this reaction has been reported as −0.284 V vs. SHE
[207], as shown at the bottom of the figure. In contrast, the slope
for acid solution is approximately 0.40 V decade−1 . There is no
simple interpretation for slopes larger than 0.118 V decade−1 ; one
possibility is that there is a thin passive surface layer, which would
lead to an effective transmission coefficient less than 0.5, in this
case, 0.148. Tsujiko et al. have discussed this dramatic difference in
behavior between acid and base and have concluded that, in basic
solution, there is an adsorption of O2 on the negatively charged
TiO2 surface [206], whereas adsorption does not take place in acid,
due to the protonation of the surface at pH lower than the potential
of zero charge.
Tafalla and Salvador proposed that the ORR on rutile takes place
with the involvement of surface states in a manner similar to
that proposed by Vandenmolden et al. for outer-sphere electron
transfer to various metal complexes [205]. The model states that
the first step is the relatively fast production of surface states,
assumed for the moment to be essentially Ti3+ in nature, in an Fig. 3.19. Comparison of Tafel plots for oxygen reduction on rutile (001) at various
electrochemical step, written here as an equilibrium: pH values: (a) pH 14 [204]; (b) pH 10.3 [204]; (c) pH 7.2 [204]; (d) pH 13 [206];
(e) pH 11.1 [206]; (f) pH 14 [205]; (g) pH 3 [205]. Curves a, b and c, measured in
air, were multiplied by 5 to make them more directly comparable with the other
Ti4+ + e− ↔ Ti3+ (3.10) curves, which were measured in pure O2 .
followed by a rate-determining step that is a purely chemical
reaction with oxygen: current densities, which are given in Table 3.1. Based on these
data, one is impressed by the wide variation in the results for
O2 + Ti3+ → O•−
2 + Ti
4+
. (3.11) similar surfaces (rutile (001)) on the one hand and the similarity
of the results for seemingly different surfaces, e.g., rutile (001) and
This model should result in a Tafel slope of 0.06 V decade−1 ,
anodized Ti.
however, not 0.12 V decade−1 , as observed experimentally. The
In comparing results for rutile (001) in various pH solutions
0.06 V decade−1 slope results from the fact that since the
from different groups, we see the variations even more graphically
electron transfer step is sufficiently fast to be considered to be in
(Fig. 3.19). Curves a, b and c show a regular positive shift in
equilibrium, its potential dependence is described by the Nernst
potential with decreasing pH [204], all with similar slopes, which
equation:
again can be approximately fitted to the standard 118 mV slope.
E = E0 − (RT /F )(ln([Ti3+ ]/[Ti4+ ])). (3.12) Curves f and g, from a different group (Tafalla and Salvador [205])
also show a positive shift in going from pH 14 to pH 3, but the
The fact that the electron transfer occurs at potentials that are as shift in comparing curve g for pH 3 with curve c for pH 7.2
much as 0.4 V positive of the flat-band potential is not consistent is unexpectedly small. The behavior obtained by Tsujiko et al.
with purely semiconducting behavior. Instead, it can be reasonably recently shows much different behavior: in going from pH 13 to pH
proposed that the electrode is sufficiently highly doped so that it 11.1, there is actually a negative potential shift. One must believe
behaves like a metal, and the band-edge energies become mobile that the surface preparation in the more recent work is superior,
or ‘‘unpinned’’. given the access to advanced surface characterization techniques.
However, curve b, for pH 0, follows a much shallower slope, This would explain the much lower absolute currents, presumably
approximately 400 mV decade−1 , i.e., an α value of ca. 0.15. For due to a nearly atomically smooth surface. The opposite trend in
the ORR in base, the first ET would involve the production of pH behavior is perplexing, however [206].
superoxide (Eq. (2.11)), which has a standard redox potential E 0 The explanation of the wide variation in electron transfer
of −0.284 V [207], indicated on the potential axis of the figure. The rates from titania electrodes to oxygen to produce superoxide
curves appear to show an onset in the region near this potential, (Eq. (2.11)) in alkaline solution may necessarily involve the bulk
while having no apparent correlation with the flat-band potential, electronic properties close to the surface. The doping level can
which is ca. −0.83 V vs. SHE (also indicated). The intercepts of the certainly vary over many orders of magnitude. From the pure
straight lines with 118 mV slopes with the E 0 yield the exchange semiconductor standpoint, Pleskov and Myamlin, among others,
534 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

have summarized this problem, concluding that the ET rate is


proportional to the density of electronic states and the electron
energy distribution function [208]. As we have seen, the doping
level can continue to increase above the point at which the
behavior is no longer like that of a semiconductor. In this regime,
characterized by the presence of surface states, one could expect
that the ET rate must also continue to rise.

3.8. Photoelectrochemical properties

Although the field of semiconductor photoelectrochemistry


was introduced to a large extent by workers at Bell Laboratories
as part of their work on semiconductor electrochemistry [165,
209–211], the field was developed to a very high level by the
work of Gerischer on materials such as Ge, GaAs, CdS and ZnO
[212]. Our early efforts were very much influenced by this work.
These semiconductors were fundamentally interesting, but the
possibilities for practical applications seemed limited due to
their susceptibility to photocorrosion. Our work [19,50,167,213,
214], showed that TiO2 is much less susceptible and could thus
be considered for applications such as solar energy conversion.
Interestingly, Boddy, as part of a study on anodic oxygen evolution, Fig. 3.20. Energy bands on TiO2 as a function of pH. Note that the rutile CB line is
had already just published a photocurrent vs. potential curve superimposed on the reversible hydrogen potential H+ /H2 at all pH.
(similar to Fig. 2.2) when our studies appeared, but the full
implications of the result were not recognized [165]. The photoelectrochemical oxidation of organic compounds
The photoelectrochemistry of TiO2 can be summarized by has been studied extensively. In a recent example, Villareal
referring to Fig. 3.20. It can be seen that the conduction-band et al. studied the photoelectrochemistry of polycrystalline anatase
energy ECB of rutile is essentially coincident with the reversible electrodes as a model system for photocatalysis and found that
hydrogen potential at all pH values, whereas that for anatase is it was possible to distinguish between direct hole-mediated
more negative by 0.20 V. At lower pH values, the ECB for rutile oxidation and •OH radical-mediated oxidation [223]. For example,
is also coincident with the reversible potential for O2 reduction methanol, which is less strongly adsorbed, was found to be
to hydroperoxyl radical HO2 •, but, at higher pH, ECB continues to oxidized via an •OH radical, while formic acid, which is more
become more negative, while the potential for O2 reduction to strongly adsorbed, was oxidized directly.
superoxide radical anion O•− 2 remains constant at −0.284 V [207].
This implies that this process should become more favorable in the 4. Fundamentals of photocatalysis
alkaline pH region, as already discussed in the previous section.
The fact that the ECB for anatase is sufficiently negative to reduce 4.1. Mechanisms of photocatalysis
O2 to O•−2 over a wide pH range is consistent with its higher
photocatalytic activity. The valence-band energies EVB for both 4.1.1. Photoelectrochemical basis of photocatalysis
rutile and anatase lie at approximately the same potential, which is As described in the Historical Overview, it became recognized

sufficiently positive to produce free •OH radicals (lower pH) or O• by several researchers that photocatalysis is based on ‘‘back-to-
radicals (higher pH). However, it is only positive enough to oxidize back’’ or short-circuited photoelectrochemical and electrochem-
2-propanol to its radical cation, in principle, in the acidic pH region. ical reactions, involving electrogenerated electrons and holes. At
The EVB for both materials is more than sufficiently positive the most global level, these can be written:
to oxidize water, by approximately 1.8 V, at all pH values. The
potentials for water oxidation are also much less positive than hν → e− +
CB + hVB (4.1)
those for the reactions involved in photocatalysis, for example, +
2H2 O + 4hVB → O2 + 4H +
(4.2)
the production of •OH radicals and the direct hole-mediated
oxidation of organic compounds. Nevertheless, some of the same O2 + 4H + 4eCB → 2H2 O.
+ −
(4.3)
reactions and intermediates may also be involved. Various workers Reactions (4.2) and (4.3) can be designated as the oxygen
have studied the detailed mechanisms of oxygen photoevolution photoevolution reaction (OPER) and the oxygen reduction reaction
reaction (OPER) on TiO2 [111,215–221]. The originally proposed (ORR), respectively. Of course, the latter can occur either, as
mechanism involved the reaction of photogenerated holes with shown, on the titania surface itself or on a separate electrode.
adsorbed hydroxyl groups (OH− ) to produce •OH radicals, which These processes can be examined separately with standard
then coupled to form H2 O2 , which was further oxidized to O2 electrochemical methods and can provide an overall rationale for
[215,216]. In order to explain the photo-etching of TiO2 , Nakato the energetics, as seen in Fig. 4.1. The point at which the anodic and
and co-workers have proposed alternative pathways involving the cathodic (flipped up to view the crossing point) currents are equal
reaction of photogenerated holes with bridging Ti–O–Ti moieties, is a good indication of the local current that could be expected
with breakage of the Ti–O bond [111,217–220] (see Section 4.1.4). under a given set of conditions. We first showed such crossing-
Recently, Neumann et al. have proposed a further modification point diagrams in a paper by Inoue et al. to show the short-
of this model [221]. Qu et al. have carried out quantum chemical circuit current for PEC cells [224]. The basic idea has been used
calculations related to these mechanisms; these workers found for many years to quantitatively analyze the energetics of local
large energy barriers to the diffusion of O-adatoms and thus electrochemical cells on corroding metals [47,48].
questioned whether O2 could be evolved as a result of the The set of results in Fig. 4.2, from Lewis and Hoffman and co-
recombination of O-adatoms [222]. workers [204] shows that, in 1 M NaOH, the crossing point for the
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 535

the photocatalytic process by adding catalysts for the ORR, such as


Pt and Pd, to TiO2 . Kesselman et al. noted that the extension of their
work to the case of real photocatalysts was not a simple one, due
to differences, for example, between rutile and anatase [204]. This
is an important point, which will be taken up again later.
This picture can then be related directly to the energy-level
diagram, in which the Fermi levels for the photo-anode (reaction
(4.2)) and cathode (reaction (4.3)) are shown as being equal (at
the crossing point in Fig. 4.2), which they must be, in a short-
circuited PEC cell (Fig. 4.1). This is helpful in understanding the
energetics and demonstrates that the reactions are expected to
proceed spontaneously. Of course, we are ignoring the detailed
mechanisms at the moment and will return to these later.
Surprisingly, the photocurrent at a single crystal rutile electrode
(Fig. 4.2, curve A), which can be totally consumed in generating
O2 , is not significantly affected by the presence of an organic
compound, as shown by Izumi et al. [227] and again by Kesselman
et al. [204]. This is because the photocurrent is essentially
determined by the characteristics of the semiconductor. At
increasingly positive potentials (low Fermi levels), the band
Fig. 4.1. Energy band diagram showing a photoelectrochemical cell, at open circuit, bending becomes sufficient for the photogenerated electrons and
in which oxygen photoevolution is occurring at a rutile electrode and oxygen
reduction is occurring at a platinum electrode. This shows the maximum open
holes to be efficiently separated, while, at increasingly negative
circuit potential that could be expected with a highly active catalyst. potentials, the band bending continues to decrease to the flat-band
condition, and then, contrary to the expected behavior for an ideal
diode (Fig. 2.2, curve 1–3), in the absence of oxygen, there is a
gap of several tenths of a volt before cathodic current starts. For
the moment, we are also ignoring the fact that there is actually
some transient cathodic current as one sweeps the potential in the
negative direction [204], which could be due to the intercalation of
protons into the TiO2 structure (see later).
A series of papers were published in which a microelectrode
was placed at various distances above an illuminated anatase film,
which was deposited on an indium–tin oxide (ITO) substrate [228–
233]. Measurements that were carried out with the microelectrode
technique showed that, even when an anatase film is actively
oxidizing an electron donor (in this case, ferrocyanide) and even
while there is oxygen reduction occurring at a bare section of the
ITO substrate, there is still a quite significant reduction current
passing at the anatase surface (Fig. 4.3) [232].
Fig. 4.2. Current–potential curves for (a) oxygen photoevolution on a rutile single At this point, it is possible to propose a tentative working
crystal; (b) hydrogen evolution on the same rutile electrode, where the cathodic mechanism for photocatalysis in terms of the balancing of anodic
current is flipped up to show the point at which it crosses curve A; (c) hydrogen
and cathodic reactions. The anodic reaction can be oxygen
evolution on a roughened platinum electrode, cathodic current and flipped up to
show the crossing point; and (d) oxygen reduction on the roughened platinum evolution, hydroxyl radical production or organic oxidation. The
electrode, cathodic and [187] flipped up to show the crossing point (based on cathodic reaction can either be oxygen reduction, to superoxide,
Kesselman et al. [204]). hydrogen peroxide or water, or reduction of the titanium dioxide
itself, via hydrogen insertion, for example. On a polycrystalline
OPER on the illuminated rutile TiO2 single crystal surface with the film, certain crystal faces may be favored as reduction sites. This
ORR also on TiO2 is at ca. 2 µA cm−2 , or ca. 25% of the maximum is not certain at the present stage of understanding. Nevertheless,
photocurrent available (Fig. 4.2, curves +B and −B, the latter being it is certain that the oxygen reduction consumes protons, thereby
flipped up to show the crossing point). Interestingly, if we show making the adsorbed water alkaline, as long as there are some
the hydrogen evolution reaction (HER) on a platinum electrode cations present. We already know that the potentials at which
(curve +C and −C), which is similar to a situation in which a TiO2 ORR proceeds are more negative in alkaline solution. Thus, the
particle is platinized, but with no oxygen to reduce, the cross-over reaction becomes progressively more difficult. However, there are
is slightly higher, ca. 3 µA cm−2 , or ca. 37.5% of the maximum. two mitigating factors: (1) the presence of CO2 in the atmosphere
This situation is also essentially the same as that in a complete PEC and its absorption could buffer the pH; and (2) O2 reduction occurs
cell and is consistent with our original work on the photo-assisted more easily in alkaline solution, i.e., larger changes in current per
electrolysis of water; this was reported for the 1 M NaOH case in unit potential change compared to acid, as already mentioned in
1975 [55], as well as with measurements of the flat-band potential Section 3.8.
[167,169,171]. There are two other factors that could help in offsetting the
If we now look at what happens when we couple the OPER on inhibiting effects of pH changes. First, the trapping of electrons
rutile TiO2 with the ORR on platinum, i.e., similar to the situation in restricted areas of the TiO2 surface and consequent increase
on the platinized TiO2 particle in the presence of dissolved oxygen, in the local electrical conductivity can lead to an increase in the
we can see that the crossing point is now near the maximum ability of that area to transfer electrons to solution phase species.
(ca. 7.5 µA cm−2 ). This idea was first proposed by Bard and co- If the concentration of trapped electrons becomes high enough,
workers [225] and has prompted several groups of researchers, the local area can become like a metal. At that point, the fact that
principally Heller and Gerischer and co-workers [226], to enhance the conduction band edge should shift to more negative potentials
536 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.3. (a) Schematic diagrams of the microelectrode monitoring system, with the microelectode situated far from the surface (left) and close to the surface (right), where
it can sense the oxygen reduction reaction that is occurring in parallel with ferrocyanide oxidation; (b) current vs. time curves showing the current when the electrode is
105 µm from the electrode (left) and when it is 15 µm from the surface; the concentration of ferrocyanide is 5 mM; the microelectrode diameter is 15 µm; supporting
electrolyte, 0.1 M K2 SO4 ; UV light intensity, 8.5 mW cm−2 , electrode potential, +0.70 V vs. SCE (taken from Maeda [232]).
© 1999, American Chemical Society.

due to a pH effect becomes meaningless, and the electrons are not This model is basically an extension or modification of that
required to pass over such an energy barrier (Fig. 4.4). At neutral of Gerischer and Heller [226,234–236], in which relatively small
pH, the conduction band energy for anatase should be ca. −0.6 V areas of TiO2 particle surfaces are proposed to act as reducing
vs. SHE, and the Fermi level should be ca. −0.4 V vs. SHE. This is sites. The novel aspect of the present modification is that it
just negative enough to reduce O2 to O•− 2 (E 0 = −0.284 V vs. provides a specific mechanism by which small reducing centers
SHE) [207]. We also note here that, for rutile, the energetics are could develop, i.e., more or less randomly selected portions of
not as favorable, i.e., the conduction-band energy would be ca. the TiO2 surface are proposed to provide starting points for
−0.4 V and the Fermi level ca. −0.2 V, which is not quite negative conductive regions to begin to develop within the bulk of the
enough to reduce O2 , as discussed in Section 3.7. If the local area particle or film. This can occur as a result of electrochemical
in which the electrons are trapped is relatively small and the hydrogen insertion (Fig. 4.5). As illumination proceeds, and more
remaining area, in which oxidation takes place is relatively large, electrons are produced, the latter can be partially consumed in the
the anodic current density in this remaining area will be small. In generation of such bulk conductive regions. The reason that such
such a case, the pH change will be smaller than expected, i.e., the structures would be spontaneously generated is that the electrons
conduction-band lowering due to acidification will be minimized, would be transported to the surface more efficiently in a more
and thus there may be minimal enhancement of the tendency to conductive region (‘‘wire’’). In other words, the first electron that
accumulate electrons at the surface (see Fig. 4.5). Now, we have a goes to the surface and produces an interstitial hydrogen atom
possible scenario in which the effect of the natural pH changes that makes it more likely that a second electron might be directed to
could occur in a weakly buffered solution is largely avoided, and the same area. The probability depends on the spatial distribution
it becomes energetically possible to reduce oxygen to superoxide of the electron density for this first bulk trapped electron. This
in areas that cover a small fraction of the surface while oxidation effect could explain why only relatively small surface regions
of organic molecules or other reactions (e.g., peroxide or oxygen might exist. Furthermore, if the electrons are channeled in this
production) can occur in areas that cover a large fraction of the way, they not only avoid electrostatic repulsion, which would tend
surface. to disperse them, but also localize them and thus protect them
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 537

reduction, so the H-rich region does not have an opportunity


to grow.
The scenario described here could be considered to be an
example of a simple self-organizing system, i.e., a system that
spontaneously organizes itself into an array (in this case, two-
dimensional) of two different types of regions with distinct
properties. There are a number of examples of such systems
in chemistry [237]. The ideal system does not assume any pre-
existing heterogeneity; this ideal could be approached on a single
crystal surface. It is quite likely that the pattern formation that
we observed on a rutile single crystal surface (see Section 5) was
of this type, in which cathodic and anodic regions might have
formed spontaneously. However, on a perfectly homogeneous
surface, the pattern formation may not be robust, and the regions
might tend to move around, based on statistical fluctuations. In
a real photocatalytic film, heterogeneity does in fact exist and
can affect the pattern of regions that might develop. The process
of formation of small conductive regions might involve a kind of
nucleation and subsequent growth, as in the deposition of crystals
or electrodeposition of metals. If the nucleation is based on specific
conditions that tend to favor it, for example, a particular crystal
face or the presence of trace impurities, the pattern could be quite
robust. This is an area that will be developed further.
Fig. 4.4. Modified band diagram for photocatalysis, showing that the cathodic As mentioned above, this is basically a modification of the
reduction of O2 takes place through surface states, with no impediment from the
Gerischer–Heller model, in which relatively small areas on the
increasingly alkaline pH at these sites.
surface trap electrons, which subsequently react with oxygen
to produce superoxide [226,234–236]. These authors envisioned
from recombination. These processes are being considered more
single, stationary surface traps on titania nanoparticles. They argue
quantitatively and will be reported elsewhere.
that the rate of electron transfer from surface traps to solution-
The insertion of hydrogen atoms would tend to increase
phase oxygen is significantly below the rate at which holes are
the volume, as already mentioned in Section 3.6, as confirmed
produced, and therefore, the number of electrons builds up within
theoretically by Koudriachova et al. [160]. This could help in
the particle to levels at which most of the holes are consumed by
explaining results that have been reported recently involving
recombination. They argue further that, in order to accelerate the
the effects of compressive vs. tensile stress and others involving
oxygen reduction reaction, electrocatalysts such as Pd or Pt could
changes in volume and roughness (see Section 5).
be used.
Several specific cases can be envisioned (Fig. 4.5). First, as a
The merits of this approach were also demonstrated by
reference case, let us say that only water, either liquid or vapor,
Kesselman et al. in their work with rutile single crystal electrode,
is present at the surface (Fig. 4.5a). A photoproduced electron is
in which they showed that the oxygen reduction current–potential
shown hitting a hydrogen-rich region and being conducted along
behavior on a Pt-covered rutile electrode surface was indeed
it to the surface, where it reacts with water, producing another
increased to the point where it matched the anodic photocurrent
interstitial hydrogen atom, which is essentially a kind of trapped
[204]. Aside from the practical problems of this approach, one
electron.
of the fundamental chemical problems is that it ignores the
If a hole trapping agent such as an alcohol is present at the
need to produce superoxide as a reaction intermediate in the
surface, the holes will be more effectively trapped, and there will
photocatalytic process. In fact, Heller and co-workers [226,
be less recombination (Fig. 4.5c). In this case, the photoproduced
234–236] and other groups, including our own, have proposed
electrons will have a greater opportunity to react and produce
mechanisms for catalysis in which superoxide plays a key role.
interstitial hydrogen. If this process continues indefinitely, a large
portion of the TiO2 particle, or even macroscopic solid, can be We propose here that oxygen reduction to superoxide and
reduced, so that the blue color produced by the presence of Ti3+ is hydrogen peroxide can occur at TiO2 at significant rates under
visible. This explains the first early report of Renz in which is noted certain conditions, specifically, those under which small near-
the discoloration of TiO2 samples exposed to organic compounds surface regions become sufficiently highly doped that the electron
and sunlight [17]. transfer approaches the limit predicted by the Marcus theory. This
When oxygen is present as well, it can be reduced to superoxide can involve doping with either oxygen vacancies, as in the case of
or hydrogen peroxide by the photoproduced electrons. In this case, bulk electrodes, or it can involve hydrogen doping, which can, in
conventional photocatalysis can take place, since the superoxide principle, be induced electrochemically, as already discussed.
can participate in the photocatalytic reactions (Fig. 4.5d). When In summary, this slightly modified Gerischer–Heller model
oxygen is present but no organic compound (Fig. 4.5c), the takes into account the requirement of electroneutrality, i.e., that
situation becomes like that for the photo-induced hydrophilic both oxidation and reduction must take place at exactly equal
effect (see Section 5). The superoxide produced by oxygen rates. It explains how oxygen reduction can proceed even at
reduction and the hydroxyl radicals produced by water oxidation significant rates without an added catalyst, and even though the
can both react further, the superoxide via disproportionation, and local pH might tend to slow down the reaction. It also explains
the hydroxyl radicals likewise via disproportionation, to produce a how photogenerated electrons can be channeled to the surface
surface that has a high coverage of hydrogen peroxide. In this case, efficiently, avoiding recombination with holes. Finally, it helps
again, the hole trapping process is not perfect, and recombination in explaining changes in bulk lattice parameters, which would
can occur, and also some electrons are removed via oxygen otherwise be difficult to explain.
538 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.5. Combined reductive–oxidative model for the photo-induced hydrophilic (PIH) effect, in which reducing and oxidizing regions are spontaneously formed: (a) in
the absence of oxygen, protons are reduced to produce interstitial hydrogen atoms (no organic compound present). (b) in the presence of oxygen, the latter is reduced to
superoxide or hydrogen peroxide (no organic compound present); (c) in the absence of oxygen, protons are again reduced, while an organic compound is oxidized; (d) in
the presence of oxygen, the latter is again reduced, while an organic compound is oxidized. In (a) and (b), holes react with water to produce a variety of possible hydrophilic
moieties.

Fig. 4.6. Processes occurring on bare TiO2 particle after UV excitation.

4.1.2. Time scales hole. The electron and hole ‘‘walk randomly’’ to the surface of
When a TiO2 photocatalyst absorbs a photon with energy equal the photocatalyst and are trapped there (Fig. 4.6); the trapped
to or greater than its band-gap, an electron–hole pair is generated; electron e− +
tr and hole htr react with acceptor (step c) or donor
subsequently, the pair is separated into a free electron and a free molecules (step d), respectively, or recombine at surface trapping
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 539

Table 4.1
Some measured characteristic time for primary processes in TiO2 photocatalysis.
Primary process Characteristic time Sample description References

Charge carrier generation


TiO2 + hν → e− + h+ fs
Charge trapping
h+ → h+ tr <200 fs
e− → e−tr <150 fs
Nanoporous TiO2 film Tamaki et al. [262]
+
hs−tr → h+ d−tr (relaxation) ∼100 ps
e− −
s−tr → es−tr (relaxation) ∼500 ps
Charge recombination
tr , h + etr , or e + h
e− + h+ 1 µs
+ − − +
Nanoporous TiO2 film (water) Yoshihara et al. [246]
→ heat (or hν ) 25 µs Nanoporous TiO2 film (air) Peiro et al. [238]
Interfacial charge transfer
∼300 ps (methanol oxidation) Nanoporous TiO2 film Tamaki et al. [264]
∼3 ns (2-propanol oxidation) Nanoporous TiO2 film Tamaki et al. [264]
<2 µs (water oxidation) Degussa P25 powder Yamakata et al. [256]
h+ (or h+
tr ) + Red →Red
+
No water oxidation within 80 µs Nanoporous TiO2 film Murai et al. [609]
e−
tr + O2 → O2

<100 ns Nanoporous TiO2 film Yoshihara et al. [246]
e− + O2 → O− 2 10–100 µs Degussa P25 powder Yamakata et al. [256]
e− (or e− −
tr ) +O2 → O2 ∼10 µs Nanoporous TiO2 film Peiro et al. [238]
e− + Pt → Pt · · · e− 2.3 ps Degussa P25 powder Iwata et al. [265]

Fig. 4.7. Time scales in photocatalysis.

sites (step a). The electron and hole can also be trapped at bulk
trapping sites and recombine there with the release of heat (b).
The processes of back-reaction between the oxidized donor and
electron, or reduced acceptor and hole, can occur after the initial
charge transfer, especially when the species are strongly adsorbed
on the TiO2 surface [238–240]. These processes together with the
characteristic surface reactions, as illustrated in Fig. 4.6, occur at
the time scales shown in Fig. 4.7 and listed in Table 4.1. Since the
band-gap transition is of an indirect nature, deexcitation through
light emission is not an important process for TiO2 [6,74].
The photogenerated charge carriers, either free or trapped,
can be probed by various spectroscopic techniques. Trapped
holes and electrons absorb light in the visible and near-infrared
spectral regions [238,241–253], whereas free electrons absorb in
the infrared or microwave regions [254–260]. These absorptions
allow one to accurately measure the time scale of photocatalytic
processes by means of transient spectroscopies, such as transient
absorption (TA) spectroscopy [238,243–249,252,253], transient
diffuse reflectance (TDR) spectroscopy [242,250,251,255,256,258],
and time-resolved microwave conductivity (TMRC) measurements Fig. 4.8. Absorption spectra of (a) trapped holes in Refs. [252,253,261], and
[254,259,260]. Typical spectra for these species are shown in (b) trapped electrons in Refs. [243,245,252,253]. (c) The spectra of trapped holes,
free electrons, and trapped electrons identified by Yoshihara et al. [246].
Fig. 4.8. © 2004, American Chemical Society.
540 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

In an early study, Serpone et al. compared the TA results of three


TiO2 sols excited by a 30-ps-width UV laser, with particle sizes of
2.1, 13.3, and 26.7 nm [247]. They observed that for all three kinds
of sols, the transient spectra were fully developed and comprised
spectra of trapped holes and trapped electrons at the end of the
30 ps laser pulse. Thus, they estimated that the trapping times
for electrons and holes should be in the ≤ 1–10 ps range. Later,
Yang et al., by means of femtosecond TA spectroscopy, estimated
the hole trapping within 50 fs and electron trapping within 260 fs
for TiO2 particles of several-nm diameter [261]. The ultrafast hole
trapping was interpreted in terms of the difference in the effective
mass of hole (0.8me ) and electron (>10me ) that results in a 40
times faster transit time from the interior to the surface of the TiO2
nanoparticle for the hole than that for the electron. Very recently, Fig. 4.9. (a) Absorption spectra of TiO2 nanocrystalline film in the ground state and
Tamaki et al. studied the trapping dynamics of electrons and of trapped hole in the film. (b) Time profiles of transient absorption of TiO2 in air,
holes in a nanoporous TiO2 film by femtosecond TA spectroscopy water, methanol, ethanol, and 2-propanol at 400 nm [264].
[262,263]. They observed that shallowly trapped electrons and © 2006, American Chemical Society.
holes were generated at a time constant of 100 fs, deeply trapped
electrons at 150 fs, and deeply trapped holes at 200 fs. In addition, [264]. They recorded the time profile at 400 nm after a laser
they assigned the time constant of 100 ps to trapped holes relaxing pulse (160 fs pulse width); at this wavelength, deeply trapped
from shallow traps to deep ones, and the time constant of 500 ps to holes contributed 90% of the absorption, while trapped electrons
trapped electrons relaxing from surface shallow traps to bulk deep contributed another 10%. In the case of the TiO2 nanoporous film in
traps. These results suggest that in TiO2 photocatalysis the species air, where there is no scavenger for trapped holes, the time profile
that lead to surface redox chemistry should be the trapped, rather showed an additional rise in TA up to 20 ps as a minor contribution
than free, charge carriers. after the instantaneous rise within the time-resolution (200 fs) of
Yoshihara et al. studied the transient adsorption of TiO2 instrument. They attributed this additional rise to the relaxation
nanoporous film in a wide wavelength range (400–2500 nm) from shallowly trapped holes to deeply trapped holes. A similar
[246]. They used a low-intensity pulsed laser to excite a TiO2 rise was observed for a film immersed in water. The TA did not
film so that less than one electron–hole pair was generated in show a decay up to 1 ns, either in air or in water, suggesting
each particle, on average, per pulse; this ensured them obtaining that electron–hole recombination and the oxidation of water are
information of the transient species under conditions close to negligible on this time scale. However, in alcohol media, the TA
those in a real photocatalytic system, i.e., with UV light below the of trapped holes decayed after excitation, clearly indicating that
1 mW cm−2 level. They found that, for a nanocrystalline TiO2 film, trapped holes reacted with the alcohols, and the reaction was
the half-life for electron–hole recombination is ca. 1 µs for a film competitive with the trapping of holes. The lifetimes of trapped
dipped in N2 -saturated deuterated water. This result is in contrast holes in the films in methanol, ethanol, and 2-propanol were
to the fast electron–hole recombination (∼10 ns) suggested by 300, 1000, and 3000 ps, respectively; the reaction efficiencies
the earlier transient spectroscopic studies on TiO2 colloids and were close to unity. The decay kinetics of the trapped holes that
powders in which high power laser pulses are used [6,259,260], react with alcohols is monoexponential on the nanosecond order,
but is consistent with recent studies by Durrant and co-workers different from the non-exponential recombination kinetics for the
[238] and Yamakata et al. [256], in which similar low-intensity film in air. This suggests that the hole transfer from trapping sites
laser pulses were employed for excitation. to alcohol molecules is the rate-limiting step for the former.
Yoshihara et al. identified the characteristic absorption of There were several studies that dealt with the electron transfer
trapped and free electrons in nanoporous TiO2 films: trapped within Pt/TiO2 photocatalysts. In a femtosecond TDR study, Furube
electrons showed a broad absorption centered at 770 nm, whereas et al. observed a new decay component of a few ps under 390 nm
free electrons showed a structureless absorption in the near- excitation in addition to the normal charge recombination kinetics
infrared and infrared region whose intensity increased with for Pt/TiO2 photocatalyst; the larger the Pt loading on the TiO2
wavelength (λ) by a function of λn [246]. They found that trapped surface, the more efficient the electron migration observed [241].
electrons reacted with oxygen more rapidly than free electrons: They thus assigned the fast decay process to the migration of
the absorption of trapped electrons decreased clearly at 100 ns electrons in TiO2 to Pt. Iwata et al. measured the femtosecond TA
after excitation when the TiO2 film was immersed in air-saturated spectra of Pt-loaded TiO2 in the spectral region of 0.9–1.5 µm. They
methanol, whereas no change was observed for free electrons at observed a decay process of 2.3 ps, which represents the transfer
the same time scale. of electrons from TiO2 to Pt. Despite this fast electron transfer,
Durrant and co-workers studied the photochemical reduction Iwata et al. suggested that a fraction of the electrons remain in the
of oxygen adsorbed on nanoporous TiO2 film by means of TA TiO2 particle, since considerable signal intensity remains even after
spectroscopy [238]. They observed that in a nitrogen atmosphere, hundreds of ps [265]. Yamakata et al. traced the absorption of free
the TA signal probed at 900 nm decayed with a half-life of electrons in TiO2 and Pt/TiO2 particles after a 10-ns-width laser
about 25 µs. The signal, however, exhibited a decay with a pulse by means of time-resolved IR absorption spectra [256]. They
half-life of about 12 µs in air, indicative of oxygen reduction observed that the absorbance on Pt/TiO2 decreased more quickly
competing directly with the electron–hole recombination reaction. in the 0–1 µs range than that on TiO2 . The rapid decay on Pt/TiO2
Yamakata et al. studied the water-induced and oxygen-induced suggests that the migration of electrons to Pt clusters occurred
decay kinetics of free electrons in TiO2 by means of time-resolved within 1 µs, and the electrons were equilibrated among the Pt
infrared absorption spectra [256]. They observed that on TiO2 clusters and TiO2 substrates after 1 µs. The three studies coincide
(Degussa P25), O2 from the gas-phase captured the electrons and with each other on the fact of fast electron transfer from TiO2 to Pt
accelerated the decay rate at a delay time of 10–100 µs, and water clusters.
vapor reacted with holes within 2 µs. The above-mentioned studies focused on TiO2 nanoparticles
Tamaki et al. studied the oxidation of adsorbed alcohols with and nanoporous films. Further studies need to be conducted on
trapped holes by means of femtosecond TA spectroscopy (Fig. 4.9) polycrystalline TiO2 films, which are also important photocatalyst
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 541

materials. However, their results are still valuable for us to


determine some important time scales in TiO2 photocatalysis,
as summarized in Fig. 4.7 and Table 4.1. These time scales,
as summarized in the table, suggest that: (1) the charge
recombination in TiO2 photocatalysts could be a slow process
due to the efficient charge trapping; (2) the interfacial charge
transfer processes are competitive to charge recombination; (3)
photocatalytic reactions could be completed within µs to ms
after the generation of the electron–hole pair. The last point
suggests that in the photocatalytic processes under low-intensity
UV illumination (1 mW cm−2 , ∼1015 photons s−1 cm−2 ), a photon
is absorbed by the TiO2 photocatalyst only after the electron–hole
pair generated by the previously absorbed photon has already
undergone reaction or charge recombination. Fig. 4.10. Schematic diagram of the reaction of trimethylacetic acid (TMA) on a
rutile (110) single crystal via adsorption at two adjacent 5-coordinate Ti4+ sites,
with simultaneous reduction of 6-coordinate Ti4+ sites to Ti3+ and subsequent re-
4.1.3. Trapping of electrons and holes oxidation by O2 to Ti4+ [280].
Charge carrier trapping is needed in the photocatalysis © 2003, American Chemical Society.
process to suppress recombination and increase the probability
of interfacial charge transfer [6,74]. As mentioned before, the above the valence-band edge, and assigned it to the recombination
recombination can be delayed, with a lifetime on the scale of of conduction-band electrons with deeply trapped holes [219,220].
µs in nanocrystalline TiO2 films [238,246]. This makes the slow No energetics information for shallowly trapped holes is available,
charge transfer of electrons to molecular oxygen competitive with probably due to their short lifetime.
the recombination. The electron paramagnetic resonance (EPR) Two different types of surface Ti–OH groups have been
technique is frequently used to trace trapped electrons and holes proposed to play important roles in trapping charge carriers,
in TiO2 nanoparticles after UV illumination [266–271]. At low forming Ti4+ –• OH radicals (hole trapping) or Ti3+ –OH groups
temperature, trapped electrons can be detected in the form of (electron trapping) [257]. This point has been investigated both
Ti3+ . Adsorption of O2 at surface has been found to remove the experimentally and theoretically recently. In an IR spectroscopic
Ti3+ signal. Trapped holes are also observed and are considered study, Panayotov et al. observed the depletion of conduction-
to be localized deep trap states. However, there is no consensus band electrons in reduced TiO2 powders with water adsorption
on the exact nature of the state: several types of species, such [279]. Compared with the oxidized TiO2 sample, the reduced
as sub-surface oxygen radicals connected to surface hydroxyls sample shows an increase in the background absorbance due to the
(Ti–O• –Ti–OH) [266,271] surface oxygen radicals (Ti–O–Ti–O• ) production of free conduction-band electrons. Introducing water
generated from basic surface hydroxyls [267,270], lattice O•
− vapor into the IR system caused a decrease of the background
radical, etc. [266,268] are postulated to exist in various cases. In a absorption. This apparent electron withdrawal was believed to be
− the result of the formation of surface hydroxyls by dissociative
recent EPR study by Berger et al., a lattice O• radical was observed
absorption of water on defects on the TiO2 surface. This is an issue
on the surface of dehydrated 13-nm TiO2 nanocrystals; the signal
− that clearly merits further study.
of O• was completely removed by the addition of a large amount In another study, Henderson et al. sought to identify surface
of O2 due to the exchange effect between the O2 molecule and the sites associated with charge transfer and trapping during the

O• species [268]. photodecomposition of a trimethyl acetate (TMA) adlayer on the
The energetics of electron trap states of various TiO2 materi- rutile (110) surface, using scanning tunneling microscopy (STM),
als have been evaluated by a variety of experimental methodolo- electron energy loss spectroscopy (EELS), and photodesorption
gies. For example, for single crystal rutile, trap sites have been (Fig. 4.10) [280]. In ultra-high-vacuum (UHV) conditions, they
shown by thermoluminescence and thermally stimulated current observed the decomposition of a TMA molecule that bridged two
measurements to exist 0.21–0.87 eV below the conduction-band Ti4+ sites into isobutene and CO2 , with the accumulation of Ti3+
edge [272–274]. For nanocrystalline TiO2 film, trap states at about cations bound to surface bridging OH groups. The Ti3+ signal,
0.5–0.7 eV below the conduction-band edge have been reported detected as a feature at about 0.8 eV in the EELS spectra, was largely
[197,201,275]. By time-resolved photoacoustic spectroscopy, elec- removed by postexposure to O2 .
tron trapping states in nanocrystalline colloidal titanium dioxide By means of density functional theory (DFT) calculations on
were determined to exist on average 0.8 eV below the conduction- hydroxylated and reduced rutile TiO2 (110) surfaces, Di Valentin
band edge [276]. Photoluminescence measurements on anatase et al. [281] demonstrated the electron trapping nature of bridging
nanoparticles revealed the presence of four shallow electron trap- Ti–OH groups, which is consistent with Henderson’s study [280],
ping levels at energies of 0.4, 0.5, 0.64 and 0.86 eV below the but found no evidence that these groups should also act as hole
conduction-band edge, ascribed to the presence of oxygen vacan- traps by formation of Ti4+ –• OH radicals. As calculated, the oxygen
cies [277,278]. Rutile particles show almost the same photolumi- vacancy introduced two localized Ti3+ 3d1 states about 1 eV
nescence from shallow traps as anatase particles, indicating the below the conduction-band edge (Fig. 4.11). These states were not
same origin of shallow traps, i.e., the presence of oxygen vacancies removed upon dissociation of a water molecule and formation
in rutile as well as anatase particles. of a pair of hydroxyl groups. Instead, the excess electrons of
The energetics of deeply trapped holes has been evaluated the hydroxylated surface were trapped by two Ti4+ ions, which
indirectly by several methods. Lawless et al. studied the reaction were reduced to Ti3+ . These two ions were found to be of two
of TiO2 particles with • OH radicals that were generated by pulse types: a six-fold-coordinated Ti located between the two bridging
radiolysis [249]. The product of the • OH reaction with the particles hydroxyls, with an energy 1.6 eV below the conduction band
was identified as a deeply trapped hole at the particle surface, (3.4 eV band-gap calculated) denoted as Ti3br+−OH , and a nearby five-
whose redox level was approximately 1.3 eV above the valence fold-coordinated Ti, with an energy 1.2 eV below the conduction
band. Nakamura et al. observed a photoluminescence centered at band, denoted as Ti35c+ . The O 2p states of the bridging OH
840 nm from the atomically smooth TiO2 (100) surface, ca. 1.5 eV groups were spread within the O 2p band, providing very little
542 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.12. Correlation of O− and Ti3+ EPR signal intensity as a function of UV


exposure time at 90 K. The intensity values for the Ti3+ and O− species were
normalized with respect to equal integral intensities of the respective spin center
signals [268].
© 2005, American Chemical Society.
Fig. 4.11. Total and projected density of states for the hydroxylated (top) and
reduced (bottom) TiO2 (110) surface, calculated using the B3LYP hybrid functional.
The Ti3+ states are localized on: (a) the Ti ion between the two bridging OH groups,
under real photocatalytic conditions nanosize TiO2 photocatalysts
Ti3br+−OH ; (d) the Ti ion nearest to the oxygen vacancy, Ti3br+−v ; (b),©Ti35c+ , top view. with highly hydrated surfaces are used. Under these conditions,
The vertical dotted line in the PDOS denotes the position of the Fermi energy [281]. there are large numbers of hydroxyl groups in a variety of different
© 2006, The American Physical Society. environments, as shown by the infrared spectra in Fig. 3.10. More
work is still needed to unravel the roles of the various surface
contribution to the top of the band. This suggests that these Ti–OH groups in photocatalytic processes.
OH groups are not very efficient hole traps. On the contrary, One important point is that, whereas almost all of the
the Ti atom of the Ti–OH group is a good electron trap, as the photogenerated holes are trapped in deep or shallow trapping
corresponding energy level for Ti3br+−OH was found 1.6 eV below the sites, a large fraction of the photogenerated electrons are in a
CB by calculation. Once the electron is trapped, the Ti3+ –OH can nearly free, untrapped state in the interior of the TiO2 particle,
attract and neutralize a hole, i.e., behave as a recombination center. as observed by a structureless absorption in the near-IR and IR
These studies have provided solid evidence for the contribution region [258]. In a recent EPR and IR study, Berger et al. found
of surface Ti–OH groups to electron trapping. For the stoichiomet- that even at the low temperature of 90 K, the major fraction,
ric (unreduced), hydrated TiO2 surface, there are at least two dif- over 90%, of photoexcited electrons remained in the conduction
ferent types of surface Ti–OH groups: one is the bridging Ti–OH band and was silent in EPR spectra under high-vacuum conditions
(actually, considered to be adsorbed protons, i.e., Ti–OH+ –Ti, as (Fig. 4.12) [268]. This may explain the slow oxygen reduction on
discussed in Section 5.2.4), and the other is the terminal Ti–OH, the irradiated TiO2 surface, since trapped electrons react with
(considered simply as OH− ions that are adsorbed at 5-coordinate molecular oxygen much faster than free electrons [246]. The
Ti4+ sites, as discussed in Section 5.2.4, or, alternatively, as the reaction of the trapped electron with O2 will be discussed also in
product of the dissociative adsorption of water at these sites (see Section 4.1.5.
also Section 3.4) [282]. The former is more acidic than the latter,
which leads to different reactivities. One can consider two path- 4.1.4. Oxidizing species at the TiO2 surface
ways for electron trapping at bridging oxygens (given formally as Many efforts have been devoted to clarifying the oxidizing
Eq. (3.2)): either the electron is trapped first and the proton trans- species generated at the irradiated TiO2 surface, which is
fer follows, or vice versa. As outlined above, there is a consensus essential for understanding the mechanism of photocatalysis and
that the neutral bridging Ti–OH group is the final product, with the for designing photocatalyst systems for environmental cleanup
electron itself trapped at the 6-coordinate Ti4+ site. There is also applications. The ever-clarified oxidizing species include holes,
evidence that some electron density exists at the 5-coordinate Ti4+ either free or trapped, • OH radicals, O•− 1
2 , and O2 , among others [6,
sites (see Sections 3.5 and 4.1.5), since both superoxide and hydro- 10,22,283,284]. H2 O2 and O2 are also involved in the photocatalytic
gen peroxide can adsorb at these sites. oxidation processes in various mechanisms [5,6,10].
More controversial is the role of the more basic terminal Holes. Holes are the primary oxidizing species in photocatalytic
Ti–OH− group, which has been postulated to trap holes by reactions. As mentioned above, photoproduced holes are trapped
some workers, but this has been disputed by others (see within picoseconds at the TiO2 photocatalyst surface [262]; this
Section 4.1.4). At this point, we should note that the role of indicates that most of the primary oxidation processes are caused
hydroxyl groups in general can become complicated, because, by trapped holes. TA studies have revealed that at least two
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 543

Fig. 4.13. Sum-frequency generation (SFG) spectral evolution of methanol on TiO2 (a) with introduction of water vapor (b) and subsequent evacuation (c) [291].
© 2004, American Chemical Society.

kinds of different trap sites exist for holes on the surface of by means of in situ sum-frequency generation (SFG), a highly
TiO2 nanocrystalline photocatalysts: the deep hole exhibits an sensitive surface characterization technique (Fig. 4.13) [291].
absorption around 520 nm, while the shallow hole exhibits an When only methanol vapor was in contact with the TiO2 film,
absorption around 1200 nm [246]. The shallowly trapped holes are two kinds of species, molecular methanol and methoxide, which
easily excited thermally into the valence band so as to establish were produced by dissociative chemisorption of methanol, were
equilibrium with the free holes [252]. Shallowly trapped holes will detected on the TiO2 surface. When a relatively large amount of
therefore have a comparable reactivity and mobility to free holes. water vapor (water–methanol ratio, ca. 300) was introduced into
The deeply trapped holes, however, are more or less localized at the system, the surface methoxide signal dropped below the SFG
deep traps and exhibit lower oxidizing potential [252,277]. The detection limit, but physisorbed methanol remained. Evacuating
precise chemical nature of these is still in question, however. the system containing a large amount of water and a relatively
Shallowly trapped holes react very rapidly with chemisorbed small amount of methanol caused the water SFG peak to drop
substances; the oxidation process is even competitive to the below the SFG detection limit. The methanol peak remained.
ultrafast charge trapping [264]. Deeply trapped holes, however, Interestingly, a substantial methoxide SFG peak reappeared. These
prefer to react with more mobile physisorbed substances, due
workers suggested that the oxidation pathway of methanol in the
to their localized nature; the reaction is often a slower process.
presence of water should depend on the ratio between water and
For instance, NMR studies by Xu and Raftery showed that for the
methanol.
oxidation of ethanol and 2-propanol, the chemisorbed species,
In a purely photoelectrochemical study, Villareal et al. found
i.e., ethoxide or propoxide, were oxidized much more rapidly than
the H-bonded species [285,286]. Shkrob et al. observed the rapid that the oxidation of methanol was never mediated by holes,
photo-oxidation of glycerol by the holes on TiO2 nanoparticles; because the conditions were aqueous solution, so that there was
ca. half of these holes are scavenged within the duration of a sufficient water available to shift the adsorption equilibrium away
3.3-ns fwhm (full width at half maximum), 355-nm excitation from methanol. In contrast, they found that formic acid oxidation
laser pulse; the others are scavenged at a slower rate over 200 ns was always mediated by holes, due to strong chemisorption of the
after the photoexcitation pulse [287,288]. They suggested that the acid on the TiO2 surface [223].
reaction with chemisorbed and physisorbed glycerol may account •OH radicals. Previous product analyses on various photocatalytic
for the prompt and the slow hole decays, respectively. Tojo et al. reactions led to the postulate that •OH radicals, produced by
studied the one-electron oxidation processes of aromatic sulfides oxidation of surface hydroxyl or adsorbed water, play an important
in TiO2 colloidal solution by laser flash photolysis [289]. They found role in initiating oxidation reactions, especially for substances that
that the 4-methylthiophenylmethanol cation radical showed a adsorb weakly on the TiO2 surface [6,11]. This oxidation pathway
broader absorption at 1 µs after excitation than that at 10 µs is sometimes designated as indirect oxidation, in comparison to
after excitation. They assigned the former to the absorption due to the direct oxidation by holes. The presence of •OH radicals at
surface-bound cation radicals and the latter to the absorption due
the irradiated TiO2 surface was demonstrated by spin trapping
to free cation radicals in solution.
experiments with EPR spectroscopy [292–295].
One important point is that even for the same substance, the
Some recent studies, however, suggest that the role of the
adsorption states may change, depending on the environment. By
means of NMR, Nosaka et al. studied the adsorption of ethanol on
•OH radical is probably overestimated. The photocatalytic one-
TiO2 photocatalysts [290]. When there was a physisorbed water electron oxidation studies by Fox and co-workers [296,297] and
layer on the surface of the photocatalyst, ethanol was found to Tachikawa et al. [298–300] showed that for many substances the
prefer to stay in the mobile water physisorbed water layer, and primary oxidation should be initiated by free or trapped holes.
thus the NMR signal for ethanol was sharp. When the physisorbed A fluorescence probe investigation by Ishibashi et al. showed
water molecules vaporized completely at elevated temperature, that free •OH radicals, detected by terephthalic acid probe, were
ethanol molecules reached the surface and reacted with the surface produced in a quantum yield three orders smaller than that of
titanols to form ethoxide, resulting in remarkable line broadening trapped holes, detected by iodide ions [301]. Nosaka et al. argued
due to the restricted mobility. With rehydration and recovery of that the common spin trapping reagents, such as 5,5-dimethyl-
the water layer, the ethoxide was hydrolyzed to form ethanol 1-pyrroline-N-oxide (DMPO), might simultaneously detect both
molecules, which moved back to the physisorbed water layer and trapped holes and surface-adsorbed •OH radicals [302]. They
produced sharp NMR signals. suggested that water reacts with the one-electron oxidation
In another study related to this, Wang et al. observed the product of the spin trapping reagent (Eqs. (4.4), (4.5)) and should
competitive adsorption between methanol and water on TiO2 film give the same product as the adduct formation between •OH
544 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.14. Reaction scheme for the oxygen photoevolution reaction on TiO2 (rutile)
in contact with an aqueous solution of pH of 1 to about 12 [219,220].
© 2004, American Chemical Society.
Fig. 4.15. Emission spectra observed after the laser pulse irradiation at 355 nm
for TiO2 (P25) suspension in water (), ethanol ( ), and mixture of water–ethanol
radicals and the spin trapping reagent (Eq. (4.6)) [302]. (M). Copyright The Owner Societies 2004. (b) Schematic illustration of the possible
formation mechanism of singlet 1 O2 on UV-excited TiO2 particles.
DMPO + h+ → DMPO•+ (4.4)
+ − can produce both surface-adsorbed •OH radicals and free •OH
DMPO• + OH → DMPO–OH • (4.5) radicals.
•OH + DMPO → DMPO–OH • . (4.6)
O2 + 2e− + 2H+ → H2 O2 (4.15)
Similarly, some reactions, which were considered to be initiated by •− •−
O2 + O2 + 2H +
→ H 2 O2 + O2 (4.16)
•OH radicals previously, could be initiated by trapped holes with
− −
the production of the same products [302]. One such example is H 2 O2 + e → •OH + OH (4.17)
the oxidation of methanol (Eqs. (4.7)–(4.10)): H 2 O2 + O2 •−
→ •OH + OH + O2 −
(4.18)
By •OH radicals:
H2 O2 + hν → •OH + •OH. (4.19)
CH3 OH + •OH → •CH2 OH + H2 O (4.7) •− •−
O2 . Superoxide O2 is less important in initiating oxidation
•CH2 OH → HCHO + H+ + e− . (4.8) reactions. Although there are several reports that suggested that
O•−
2 initiated some type of oxidation reaction, for example, the
By trapped holes: oxidation of trichloroethylene [307], the examples are quite few
compared to those concerning holes and •OH radicals. The role of
CH3 OH + h+ → + CH3 OH → •CH2 OH + H+ (4.9) O•− in photocatalytic oxidation processes was discussed mainly
2
•CH2 OH → HCHO + H + e . + −
(4.10) within the framework of participation in the total mineralization of
organic substances through reaction with organoperoxy radicals,
•OH radicals have been proposed to be generated by the the production of H2 O2 by disproportionation reaction, and the
oxidation of either water or adsorbed hydroxide ion with anti-microbial activities [5,6,10]. Recently, Nosaka et al. proposed
photogenerated holes [6,215,216,303,304]. Nakato and co-workers that the reaction between O•− 2 and a trapped hole would produce
and others have found problems with this simple electrochemical singlet oxygen 1 O2 , a strong oxidant (see below) [283].
process (e.g., failure to explain the 840-nm photoluminescence 1
O2 . Singlet oxygen 1 O2 is an important reactive oxygen species
peak) [111,217–220,305] and have proposed an alternative in atmospheric, biological and therapeutic processes, and is also
pathway in which •OH radicals are produced by a nucleophilic used as a reagent in organic synthesis. The formation of 1 O2 in
attack of water on a hole trapped at a surface lattice oxygen TiO2 photocatalysis was predicted twenty years ago [308], but
((4.11)–(4.14)). only recently 1 O2 generated on irradiated TiO2 surface could be
detected. Nosaka et al. first reported the detection of the near-
[Ti–O–Ti]s + h+ + H2 O → [Ti–O • HO–Ti]s + H+ (4.11) infrared phosphorescence at 1270 nm of 1 O2 in a photocatalytic
+ TiO2 aqueous suspension system by means of a gated photon
[Ti–O • HO–Ti]s → [Ti–O–O–Ti]s + H (4.12)
counting method (Fig. 4.15a) [283]. They suggested that oxidation
[Ti–O–O–Ti]s + H2 O → [Ti–O–OH • HO–Ti]s . (4.13)
of O•− 2 with holes at the TiO2 surface should be a plausible
When the O–O bond in TiO–OH breaks mechanism to produce 1 O2 (Fig. 4.15b). The lifetime of 1 O2 was
determined to be 2 µs, which is rather short as compared with
[Ti–O–OH • HO–Ti]s → [Ti–O • HO–Ti]s + •OH. (4.14) those of •OH radicals (ca. 10 µs) and trapped holes, possibly due
to the fast deactivation of 1 O2 at TiO2 surfaces.
A similar scheme is shown in Fig. 4.14, in which the overall water Daimon and Nosaka measured the quantum yields for 1 O2 over
oxidation process is shown. Quantum mechanical calculations ten commercial TiO2 photocatalysts in air, and found they ranged
have also been carried out by Qu and Kroes recently to check these from 0.12 to 0.38, while the lifetimes ranged from 2.0 to 2.5 µs
proposed pathways [222]. These authors examined a large number [309]. The short lifetime suggests that the generated 1 O2 mainly
of different possible geometries for adsorbed oxygen atoms and remains on the TiO2 surface without being released to the air or
peroxide groups. Neumann et al. examined this mechanism solution. Since the quantum yield is significantly high, 1 O2 may
experimentally with electrochemical mass spectroscopic and contribute to the oxidation of organic molecules at the irradiated
photovoltage measurements and found problems with it [221]. TiO2 surface.
They proposed a slightly modified pathway for water oxidation. Hirakawa et al. studied the generation of 1 O2 by the photocat-
Moreover, there are several reactions of H2 O2 , produced by alytic reaction of TiO2 particles, either anatase or rutile, dispersed
reduction of molecular oxygen or the disproportionation reaction in solvents [310]. The estimated quantum yield by both types of
of superoxides, could produce •OH radicals (Eqs. (4.15)–(4.19)) [6, TiO2 was about 0.02 in ethanol. The 1 O2 emission was completely
306]. It should be noted that reaction (Eq. (4.17)) only produces quenched by the addition of superoxide dismutase, suggesting that
surface-adsorbed •OH radicals, whereas reactions (4.18) and (4.19) 1
O2 is produced via oxidation of O•− 1
2 . The O2 was found to be
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 545

adsorbed at the TiO2 surface and could not oxidize nicotinamide


adenine dinucleotide, which has no affinity for the TiO2 surface.
However, it can oxidize the water-soluble protein bovine serum al-
bumin, which can be adsorbed on the TiO2 surface. The generation
of 1 O2 was enhanced in the phospholipid membrane, suggesting
that the phospholipid membrane is an important environment for
the phototoxic reaction mediated by 1 O2 generation in TiO2 pho-
tocatalysis.
Janczyk et al. reported the generation of 1 O2 at surface-modified
TiO2 particles [311]. They found that TiO2 particles, after surface-
modification with organosilane, could oxidize cyanuric acid, a
stable material resistant to TiO2 photocatalysis. They concluded
that cyanuric acid was oxidized by 1 O2 and suggested an energy
transfer pathway for the generation of 1 O2 on hydrophobically
modified TiO2 particles.
Fig. 4.16. Time-resolved accumulation of trapped holes and scavenged electrons
4.1.5. Role of molecular oxygen upon UV exposure at different oxygen pressures: (a) 0.1 mbar O2 and (b) 10 mbar
Molecular oxygen plays an indispensable role in TiO2 photo- O2 . (c) UV exposure at P < 10−6 mbar and subsequent addition of O2 . Note that
the profiles of O− and O•− − •−
2 coincide for a given pressure, as the ratio of O to O2
catalysis, especially in reactions related to environmental cleanup
equals unity throughout all experiments [312].
applications [5,6,10]. It assists the charge separation in TiO2 by © 2005, Wiley-VCH Verlag GmbH & Co.
capturing TiO2 electrons; it generates active species such as O•− 2 ,
H2 O2 , 1 O2 , etc. that participate in reactions; it itself participates greater concentration of O•− 2 at the water–TiO2 interface compared
in the reaction and accelerates the mineralization of organic sub- to the air–TiO2 interface was thought to be due to the lower inter-
stances; it also helps the TiO2 photocatalyst to maintain stoi- facial energy barrier for the electron transfer in water.
chiometry during photocatalytic reactions. These effects have been As discussed in Section 3.7, the reaction of TiO2 electrons with
either postulated or already demonstrated for many years, but re- molecular oxygen is a process that is still poorly characterized,
cent progress in both theoretical simulations and experimental varying over many orders of magnitude in rate, even for samples
techniques have greatly improved our understanding. that are nominally identical. It has been considered as the rate-
Improved charge separation. Berger et al. found that molecular determining step in photocatalytic processes [234,236], although
oxygen could not only capture the photogenerated electrons this is still not certain. The kinetics for oxygen reduction on the
to suppress charge recombination, but also improve the photo- TiO2 surface can probably be improved by depositing a noble metal
stimulated charge separation in TiO2 nanoparticles, i.e., more holes co-catalyst [226], although this is a far from practical solution.
could be trapped within the particles in the presence of molecular It will be interesting, both fundamentally and practically, to try
oxygen [312]. They designed two experiments to demonstrate to understand the detailed reasons for the wide variations in
this idea: in one experiment, they illuminated dehydrated TiO2 rate. As discussed in Section 3.8, the driving force for electron
nanocrystals (13 nm) and recorded the EPR spectra before and transfer is quite low, as the potential difference between the
after the addition of molecular oxygen; in another experiment, conduction-band edge of TiO2 and the potential for O2 reduction
they illuminated dehydrated TiO2 nanocrystals and recorded the to superoxide is quite small, which may result in a slow electron
EPR spectra in the presence of molecular oxygen. They observed transfer [238]. This driving force is most likely decreased even
that, in the former experiment, the concentration of trapped holes further, due to the fact that the electron transfer takes place from
(O− ) was the same as the superoxide, indicative of the depletion surface electron traps.
of photogenerated electrons with molecular oxygen; thermally Generation of active oxygen species. Superoxide (O•−2 ) and hydrogen
stable charge separation could be maintained for quite a long time peroxide are two main products of O2 reduction in TiO2
after the formation of superoxide. In the latter experiment, they photocatalysis and are involved in the production of other active
observed that the concentrations of O•− −
2 and O were related to the oxygen species such as •OH, 1 O2 , as described in Section 4.1.4.
oxygen pressure during UV excitation, first increasing with oxygen While superoxide is the product of one-electron reduction of
pressure, and finally saturating at 25 mbar O2 (Fig. 4.16). This molecular oxygen, hydrogen peroxide can be produced by either
indicates that either the number of adsorption sites for O•− 2 ions two-electron reduction of molecular oxygen or disproportionation
on the nanocrystal surface or the surface capacity for hole trapping of two superoxides [6]. The generation of superoxide has been
and O− formation limits the final concentration of EPR-active recognized for a long time based on the EPR [268,269,312,315,316]
species. Most interestingly, at 140 K, a maximum concentration or chemiluminescence studies [313,314,317,318]. The detection
of ten O− /O•−2 pairs per particle was obtained after O2 photo- of surface-peroxo species, however, is difficult, and only recently
adsorption, but, under high-vacuum conditions, UV excitation was achieved by Nakamura et al. by means of multiple internal
resulted in only one e− /O− pair per particle. reflection infrared spectroscopy [319].
Ishibashi et al. quantitatively studied the kinetics of pho- Tilocca et al. studied the adsorption of O2 on hydroxylated
togenerated O•− 2 under very low-intensity light illumination rutile TiO2 (110) using density functional theory total energy
(1 µW cm−2 ) at the air–TiO2 interface as well as the water–TiO2 calculations and first principles molecular dynamics simulations
interface by measuring the chemiluminescence from luminol so- [320]. Their results suggested that the electron transfer from Ti–OH
lution pipetted onto TiO2 film photocatalysts [313,314]. The total to O2 does not need to pass through a chemisorbed O2 state.
number of steady-state O•− 2 on the TiO2 surface was determined Instead, scavenging by O2 starts as soon as the oxygen molecule
from the chemiluminescence intensity: about 1 × 1014 cm−2 at captures two hydrogen atoms from the bridging hydroxyls
the air–TiO2 interface and 2 × 1014 cm−2 at the water–TiO2 inter- (Fig. 4.17, a–c); this produces a hydroperoxide species, which
face. These steady-state concentrations were independent of light is eventually chemisorbed onto the five-fold-coordinated surface
intensity (1 µW cm−2 –15 mW cm−2 ), indicating that the amount Ti (Ti5c ) atoms (Fig. 4.17, d–f). The H2 O2 –Ti5c molecule was
of O•−
2 is restricted by the number of adsorption sites on TiO2 . The further transformed to HO2 –Ti5c after transferring a proton back
546 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.17. Selected snapshots from the MD trajectory started with an O2 molecule above the surface, illustrating the formation of H2 O2 , its adsorption and further
transformation. Slab Ti and O atoms are represented as white and dark sticks, while adsorbed O and H atoms are represented as large dark and small white spheres,
respectively; dashed lines denote hydrogen bonds [320].
© 2005, American Chemical Society.

to a bridging oxygen (Fig. 4.17, g–i). The theoretical calculations


support the experimental hypothesis of Henderson that electron
scavenging could happen between weakly bound (physisorbed) O2
and bridging hydroxyls [143].
Atwood et al. compared the oxygen-centered radicals gener-
ated on surfaces of Degussa P25 powder and two rutile powders
by EPR spectroscopy [321]. They found that stable superoxide
radicals are generated on either dehydroxylated P25 powder or
dehydrated rutile powders under UV excitation; the signal can per-
sist up to several days. On the fully hydrated P25 surface, no signals
(O− , O•−
2 , •O2 H) are observed in the EPR spectrum, even after pro-
longed UV irradiation at 100 K. On the partially hydrated P25 sur- Fig. 4.18. Photocatalytic decomposition of several kinds of organic compounds
face, the dominant signal of O− ; and traces of O•− 2 and •O2 H are
through the tetraoxide mechanism suggested by Schwitzgebel et al. [235].
observed; both were thermally unstable and quickly decayed to © 1995, American Chemical Society.
diamagnetic species such as H2 O2 upon warming to 300 K. How-
ever, thermally stable O•−2 and •O2 H could be observed on rutile
Although valence-band holes of TiO2 were required for the
powders, even in the case of full hydration. It was believed that on oxidation to be initiated, the actual oxidizer was molecular
the P25 surface, which is mainly anatase, superoxide anions can be oxygen. O2 was suggested to react with organic radicals, generated
sufficiently mobile on the fully and partially hydrated surfaces such upon the hole or •OH radical reaction with the reactant, to
that disproportionation reactions (forming H2 O2 ) occur. However, produce an organoperoxy radical (ROO•), and subsequently the
on the hydrated rutile surface, some surface sites capable of O•− 2 reactant was degraded and mineralized through thermochemical
stabilization remain. reactions, probably by the tetraoxide pathway (Fig. 4.18). A similar
Consistent with Atwood’s study, Goto et al. reported a conclusion was also drawn from photoelectrochemical studies
preferred generation of hydrogen peroxide on anatase during [230,231] and isotope studies [323].
the photocatalytic degradation of 2-propanol in water, while for Although the existence of the organoperoxy radical was
rutile photocatalysts, the generation of superoxide was preferred predicted in photocatalytic processes, its detection was difficult
[322]. They observed this trend on several anatase and rutile due to its short lifetime [324–326]. In a recent EPR study, Murphy
photocatalysts; the data are shown in Table 4.2. et al. observed that a number of different organoperoxy species
Participation in reactions. Heller suggested that the photocatalytic are involved in the heterogeneous decomposition of ketones over
oxidation of organic pollutants is mostly mediated by molecular dehydrated polycrystalline TiO2 [327]. These radical intermediates
oxygen, not by photogenerated holes or •OH radicals [5,235]. were identified as organoperoxy-based species of general formula
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 547

Table 4.2
Physical properties and estimated production rates of hydrogen peroxide and superoxide ion on several TiO2 powders [322].

TiO2 powder Anatase component (%) Surface area (m2 /g) Purity (%) H2 O2 (10−7 mol min−1 ) 2 (10
O•− −7
mol min−1 )

NS-90 <1.0 0.9 >99 0 1.0


CR-EL <0.1 7.1 >99 0.1 0.9
TIO-3 <1.0 48 >99 0 1.6
TIO-2 >99.9 18 98.5 2.9 0.4
ST-21 >99.9 56 95 5.3 0.2
ST-11 >99.9 71 95 2.7 0.5

reactions [329–337]. Under O2 -deficient conditions, these O atoms


would gradually be lost during photocatalytic reactions; this
results in a loss of stoichiometry of the TiO2 photocatalyst
and also its photocatalytic activity [329–333]. Yoshida et al.
recently studied the structural changes of a TiO2 film during
the photocatalytic oxidation of an organic film in vacuum by a
unique in situ TEM/ EELS system [331–333]. The organic film
was deposited on a 60-nm-thick TiO2 film supported on a Cu
grid for TEM observation. During UV irradiation, the organic
material was oxidized and decomposed, desorbing from the TiO2
surface (Fig. 4.19). At the same time, single crystalline TiO2
was transformed into polycrystals, and additional lattice fringes,
Fig. 4.19. High-resolution transmission electron micrograph of the hydrocarbon/ belonging to substoichiometric forms of TiO2 , appeared. The TiO2
TiO2 films that were illuminated with UV light for 3 h. (1) A low-magnification film gradually became a network of aggregates after one week of
image and the corresponding selected area electron diffraction pattern (inset). irradiation. A detailed EELS analysis revealed that the structural
(b) A high-magnification image. Additional lattice fringes (d = 0.45 nm and 0.69
changes were associated with the loss of oxygen atoms in the TiO2
nm) which do not appear in a perfect rutile crystal, were observed in the inset of (a)
(marked by arrows) and (b) [333]. crystal lattice [331–333].
© 2006, American Institute of Physics. Muggli and Falconer studied the photocatalytic oxidation of
formic acid and acetic acid in a helium atmosphere [329,330]. They
ROO• and RCO3 •. They were formed by reaction of photogenerated observed the fast deactivation of the TiO2 photocatalyst and the
charge carriers with the adsorbed ketones in the presence of recovery of activity by storage in O2 in the dark. They suggested
molecular oxygen. The organoperoxy intermediates are thermally that the loss of surface lattice oxygen caused the deactivation of
unstable and decompose at temperatures in the region of the photocatalyst, and O2 was able to replenish the lattice oxygen
180–250 K. This work demonstrates the importance of free-radical vacancies.
pathways involving both organoperoxy and superoxide radicals for
the photo-oxidation processes over TiO2 photocatalysts. 4.1.6. Effect of crystal face
Maintenance of the stoichiometry of TiO2 . Surface coordinatively Morris Hotsenpiller et al. studied the photochemical reduction
unsaturated O atoms of TiO2 photocatalysts, especially those at of Ag+ ions in an aqueous solution on oriented TiO2 (rutile)
surface irregularities such as kinks and edges, are labile under surfaces [338]. They observed the oriented photochemical reaction
UV illumination due to hole trapping, and could be exchanged rates and quantum yields, i.e., the (101), (111), and (001)
with molecular oxygen [328] or abstracted during photocatalytic surfaces had higher Ag-deposition rates than the (100) and (110)

Fig. 4.20. (a) AFM image (1 × 1 µm2 ) of ultra-smooth TiO2 (110) surface. (b) Ag formed on as-polished (dark bar) and ultra-smooth (slashed bar) TiO2 surfaces with various
orientations. The Ag/Ti intensity ratio was evaluated by EDX [340].
© 2005, The Japan Society of Applied Physics.
548 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.21. (a) A SEM image of n-TiO2 (001) surface of (001) after photoelectrochemical etching under 2.0 V anodic polarization [218]. (b) A SEM image of TiO2 polycrystalline
electrode after photoelectrochemical etching under 1.0 V anodic polarization [344].
© 1998, The Electrochemical Society, Inc.

(111) > (110) [340]. Note that the Ag reduction reaction


is accompanied by water photo-oxidation in their experiments.
Thus, the above reactivity sequence should represent the sum of
reduction and oxidation reactions. In another study, Nakamura
et al. measured the flat-band potential of atomically smooth TiO2
rutile single crystal surfaces [111]. They found that the flat-band
potential of the (100) surface is 0.1 V more negative than that of
the (110) surface, indicative of the higher reduction ability of the
former surface.
Nakato et al. observed the anisotropic etching of TiO2 single
crystal surface during anodized polarization in H2 SO4 electrolyte
solution [217,341,342]. The etching left rectangular holes on the
rutile (001) surface (Fig. 4.21a) [218]. The holes, 50–200 nm wide,
extend in the (001) direction, with the (100) face or equivalent
exposed selectively at the walls. For the (110) surface, long
rectangular grooves were produced at the surface in the (001)
direction, with the (100) face or equivalent exposed at the walls
of the grooves. For the (100) surface, the results were quite
similar to those for the (110) surface, except that the walls of
the photoproduced grooves were either parallel or perpendicular
to the electrode surface. This also shows that the (100) face or
equivalent is exposed at the walls. A similar phenomenon was also
observed on polycrystalline TiO2 electrodes during anodization
Fig. 4.22. (a) SEM image of a rutile particle on which Pt fine particles were
polarization in H2 SO4 solution, with uniform rectangular cells
deposited selectively on (110) face by photocatalytic reactions. (b) SEM image of created on the surfaces of microcrystals (Fig. 4.21b) [343,344].
an anatase particle (center) on which Pt fine particles (indicated by arrows) were These etching results were considered to be abnormal, since the
deposited by photocatalytic reactions. (c) and (d) SEM images of rutile and anatase (110) face is the most stable for rutile, rather than the (100) face,
particles, respectively, co-deposited with Pt and PbO2 by photocatalytic reactions.
but this could be explained by the difference in hole reactivity with
© 2002, The Royal Society of Chemistry and the Centre National de la Recherche
Scientifique [345]. water for various crystal faces [218].
Matsumura and co-workers studied the crystal-face-dependent
photocatalytic reactivity of rutile and anatase particles [345]. The
surfaces. In a later study, Lowekamp et al. used polished and TiO2 powder used in their study consisted of 1-µm size rutile
annealed surfaces of randomly textured rutile polycrystals to and anatase particles with well-developed crystal faces. The rutile
photochemically reduce Ag+ to Ag metal in an aqueous solution particles showed a tetragonal prism structure with four planes,
[339]. By correlating the surface orientation of more than 100 which are assigned to the (110) faces. Each end of the prism is
individual crystallites with the amount of deposited Ag, they capped by four planes, which were assigned to the (011) faces. The
concluded that the most reactive orientations lie near the (101) anatase particle had a tetragonal bipyramid structure consisting of
surfaces. These studies, however, did not involve atomically eight (011) faces. Both vertexes were removed and capped by (001)
smooth surfaces, and thus might be open to question (see below). faces. After photocatalytic oxidation of water on the powder using
In a recent study, Yamamoto et al. reconsidered the anisotropic hexachloroplatinate ions as the electron acceptors, Pt deposits
photochemical reactivities of rutile with atomically smooth rutile were observed mostly on the rutile particles, especially on the
single crystal surfaces (Fig. 4.20) [340]. They observed only little (110) face, indicating that the reduction sites of rutile particles
orientation dependence of the photochemical reactivity for Ag+ were on the (110) face (Fig. 4.22a). When 2-propanol was added
reduction on the group of as-polished surfaces. In contrast, the to the solution, Pt was deposited on both the anatase (Fig. 4.22b)
photo-reactivity on the group of ultra-smooth surfaces exhibited and rutile particles. Using the thus-prepared Pt-deposited TiO2
strong orientation dependence, although their reactivities were powder, Pb2+ ions were photocatalytically oxidized to PbO2 , which
one to two orders of magnitude smaller than those on the as- were deposited on the TiO2 particles (Fig. 4.22, c, d). They observed
polished surfaces. The reactivity of ultra-smooth surfaces for Ag+ the selective deposition of PbO2 on the (011) face of rutile particles,
reduction followed the sequence: (101) > (100) > (001) > and for anatase particles, PbO2 was deposited in a larger amount on
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 549

carried out in which highly well-defined experimental results are


compared with high accuracy quantum mechanical calculations.

4.1.7. Remote photocatalysis


Recent studies have shown that photocatalytic reactions could
be initiated at a distance away from the TiO2 photocatalysts. Some
examples include the remote killing of bacteria [347], the remote
decoloration of dye films [348–352], the remote decomposition
of organic or polymeric films [350,353–358], and the remote
oxidation of inorganic materials (Cu, C, SiC, etc.) [359,360]. In
these examples, the impact distance of the TiO2 photocatalyst was
Fig. 4.23. SEM images of (a) an anatase particle after UV irradiation in aqueous found to be as long as several millimeters, and the reaction media
solution containing 1 mM H2 PtCl6 and 0.52 M 2-propanol and (b) an anatase particle could be air, water, or solid surfaces; in addition, several types
after sequential UV irradiation in the same solution as in (a) and then in an aqueous of oxidizing species were involved in the photocatalytic reactions
solution of 0.1 M Pb(NO3 )2 [346].
© 2003, The Royal Society of Chemistry and the Centre National de la Recherche (Fig. 4.24).
Scientifique. Tatsuma et al. first observed the remote oxidation of methylene
blue film at a distance of 500 µm away from a TiO2 film [352].
Later studies showed that aromatic or even aliphatic substances
can be oxidized and decomposed into CO2 remotely by TiO2
films (Fig. 4.25), and the impact distance could be up to 2.2 mm
[350]. It was clear that the oxidation could not be caused by
holes on the TiO2 surface. Similarly, it was hard to believe that
hydroxyl radicals generated on the TiO2 surface could travel such
a long distance in air, considering their high activity [350]. Some
neutral oxygen species generated on illuminated TiO2 surfaces,
such as HO2 , H2 O2 , can diffuse a long distance in air. However,
Fig. 4.24. Active oxygen species and some processes related to remote these species are not reactive enough with aliphatic hydrocarbons
photocatalysis. Neutral species including H2 O2 , HO2 , and 1 O2 are supposed to to produce CO2 . Tatsuma et al. thus postulated that hydroxyl
diffuse farther away from irradiated TiO2 surface than the highly reactive •OH, radicals should participate in the remote photocatalytic reactions
through air, water, or surface.
described above, but a major proportion of these hydroxyl radicals
should be generated in the gas phase by the UV-photolysis of H2 O2
the (001) face than on the (011) face. This means that the (011) face [350]. The possible remote oxidation mechanism thus includes
provides the oxidation site for rutile particles, while for anatase (1) the diffusion of H2 O2 in the gas phase, (2) the photolysis of
particles the (001) face is more oxidative than the (011) face. H2 O2 into hydroxyl radicals in the gas phase under UV irradiation
In another study, Matsumura and co-workers etched the (λ < 365 nm), and (3) oxygenation and decomposition of target
same anatase particles with HF solution [346]. By this process, materials by hydroxyl radicals. Also, H2 O2 can be generated in the
the edge between two (011) faces was selectively etched and disproportionation of HO2 . Thus, either H2 O2 or HO2 can be the
as a result, eight new faces were generated on each particle, diffusing species [350].
which were assigned as (112) faces. The new anatase particles Kubo and Tatsuma designed several experiments to verify
were photocatalytically deposited with Pt and PbO2 in sequence. the proposed double-excitation mechanism. In one of these
Interestingly, most of the Pt was deposited on the (011) faces, experiments, air was forced to flow through a TiO2 -loaded cell that
while PbO2 was deposited on the newly generated (112) faces, as was irradiated with UV light and an aqueous solution of peroxidase
shown in Fig. 4.23, indicating the reductive function of the (011) in sequence. By this experiment, they proved the existence of
faces and the oxidative function of the (112) faces for the anatase H2 O2 in the gas phase, released from the irradiated TiO2 film
particles. These results suggest that crystal faces could assist in the [361]. In another work, they observed that H2 O2 vapor, when
separation of electrons and holes, and thus, it may be possible to illuminated with UV light shorter than 365 nm, could oxidize
obtain high photocatalytic activity through control of the surface ODS-modified glass [362]. Importantly, the reaction rates of the
atomic arrangement of TiO2 photocatalysts. More work needs to be H2 O2 –UV system were in close agreement with those for the

Fig. 4.25. Time course of water contact angle of (a) octadecyltriethoxysilane (ODS) modified glass surface and (b) polystyrene-coated glass surface during the remote
oxidation experiment. The polyimide spacer was 50 µm thick and the light source was an Hg–Xe lamp (light intensity 10 mW cm−2 ) [350].
© 2001, American Chemical Society.
550 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

involved in the bleaching of the fluorescence of dye molecules


[351]. But they also argued that it is hard to exclude the
contribution of other active oxygen species, such as 1 O2 . They
designed an experiment that used terrylenediimide (TDI) as a
fluorescent probe for 1 O2 . The molecule reacts to 1 O2 to form
a strongly fluorescent TDI diepoxide with blue-shifted emission.
They found that 1 O2 generated on UV-excited TiO2 film could
diffuse in air over 2 mm distance, and oxidize TCI molecules at that
distance (Fig. 4.27) [348].
Haick et al. reported another type of remote photocatalytic
oxidation by surface diffusion of oxidizing species [357,358]. They
prepared microstripes of TiO2 on an Si substrate and observed
the mineralization of aliphatic chains anchored to inert silicon
domains within minutes, even when these chains were located as
far as 20 µm away from the TiO2 microstripes [358]. The hydroxyl
radical is believed to be the oxidizing species in this system. In
another study, they covered the TiO2 film with a photomask and
Fig. 4.26. Correlations between the remote oxidation ability and the H2 O2 observed the decomposition of aliphatic chains in the dark region
generation ability for various photocatalysts (). TiO2 ‘‘A’’ and ‘‘B’’ are anatase sols,
[357]. Very interestingly, the occurrence of the photocatalytic
with average particle sizes of 20 nm and 7 nm, respectively. Pt (1.0 wt%) and
Ag (0.0055 wt%) were deposited photocatalytically. The correlation between the reaction in the dark region did not cause any decrease in the
oxidation ability and the H2 O2 amount for the H2 O2 –UV reaction is also shown (4) reaction rate in the illuminated region, even though competition
[362]. may exist in sharing the oxidizing species. Kawahara et al. studied
© 2006, American Chemical Society. the surface diffusion behavior of photogenerated oxidizing species
on TiO2 films and concluded that the diffusion length was as long
photocatalytic remote oxidation, suggesting the similarity of the as 75 µm [356].
active species in both systems (Fig. 4.26). They also compared the The observation of remote photocatalysis, especially that
remote oxidation of octadecyltriethoxysilane (ODS)-modified glass induced by surface diffusion of oxidizing species, may have a
with various photocatalysts, and found that the oxidation rates significant effect on the design of photocatalytic materials. The
followed the sequence: Pt-loaded TiO2 > TiO2 > ZnO > WO3 , efficiency of photocatalysis is restricted by recombination. In
the same sequence for the decomposition activity of methylene the illuminated state, on the TiO2 surface, electrons, holes, and
blue in conventional photocatalysis [362]. Thus, they suggested active oxygen species maintain a dynamical equilibrium. When
that the conventional photocatalytic activity could be a measure there is spillover of an oxidizing species, light can be utilized
for roughly predicting the remote oxidation activity of a given more efficiently for chemical conversion. This implies that a
photocatalyst [362]. hybrid of TiO2 and an adsorbent (e.g., zeolite, active carbon) may
Naito et al. studied the remote photocatalysis by means of exhibit performance better than that of pure TiO2 . The role of the
single-molecule fluorescence microscopy [348,351]. They found adsorbent is not restricted to the adsorption of pollutants; it can
that the double-excitation process described above should be also act as a reaction site via the out-diffusion of oxidizing species.

Fig. 4.27. Fluorescence images of single terrylenediimide (TDI) molecules spin-coated on a polymethyl methacrylate (PMMA)-coated coverslip before (a) and after (b) UV
irradiation for 5 min (scale bars are 10 µm). The air gap between TDI layer and TiO2 film is 12.5 µm. The bright spots correspond to TDI diepoxide. The UV irradiation area
is inside the white circle in the images, in this area TDIs are bleached by direct UV irradiation. (c) Scheme for the fluorescent detection of airborne 1 O2 with TDI [348].
© 2006, American Chemical Society.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 551

Fig. 4.28. (a) Quantum yield dependence on absorbed photons at various initial 2-propanol concentrations; (b) quantum yield dependence on normalized absorbed photon
number (Inorm ) at various initial 2-propanol concentrations [366].
© 1997, American Chemical Society.

The observation of remote photocatalysis may cause concern way. Ohko et al. measured the quantum yield (QY) of a
regarding the safety of TiO2 photocatalysts. However, the reaction simple photocatalytic reaction, 2-propanol oxidation, on a TiO2
rates for remote photocatalysis are lower than those for ordinary film in ambient air, for a series of 2-propanol concentrations
photocatalysis, and they decrease with distance. For a 50-µm (1–1000 ppmv) and a wide range of light intensity (tens of
diffusion distance, the rate of remote photocatalysis via air nW cm−2 –1 mW cm−2 ) [366]. They found that the QY increased
diffusion is ca. 2 orders of magnitude lower than that of ordinary gradually with decreasing light intensity I, and reached a
photocatalysis, while that through surface diffusion is ca. 1 order of maximum value when the light intensity was extremely low. They
magnitude lower, as roughly estimated by the published data [350, designated the condition for maximum QY as light-limited. They
356,357]. Therefore, the TiO2 surface should not be harmful even also found that when the 2-propanol concentration was decreased,
when human skin is in close proximity to it, in the low-intensity the maximum QY was obtained at lower light intensity. They
UV light contained in solar light or interior illumination. replotted QY as a function of a normalized light intensity (Inorm ),
Remote photocatalysis through air diffusion of oxidizing species which was defined as the ratio of the number of photons absorbed
may have applications in lithographic techniques. Tatsuma et al. per second to the number of adsorbed 2-propanol molecules,
[360,363] have investigated the feasibility of this new technique both normalized to the geometric area. The plots obtained for
and have observed the patterning of copper plates and organic different initial concentrations of 2-propanol fell on the same
monolayer-modified glass, with a resolution of 10 µm or better, by curve, which is a very interesting result (Fig. 4.28). QY increased
placing a TiO2 film on the surface to be patterned and irradiating as Inorm decreased and finally became constant at ∼28% for Inorm
the film through a photomask. The technique requires neither values below 10−4 s−1 . The maximum QY varied from film to
photo-resist nor special equipment other than a light source. film, which indicated that there are recombination losses that are
intrinsic to each film. As interesting as these results are, they are
4.2. Photocatalytic reactions yet to be adequately explained.
The mechanism of photodecomposition of 2-propanol does
As mentioned above, several kinds of active oxidative species, not involve chain reactions. Only one photon participates in

including free and trapped holes, •OH radicals, O•2 , and 1 O2 , are the generation of one molecule of acetone. Interestingly, the
involved in the initiation of photocatalytic oxidation reactions [6, QY of acetaldehyde degradation, for which there is the possible
11,15,16,283,302,311,317,364]. This offers the TiO2 photocatalyst’s involvement of radical chain reactions, exhibited the same trend
superior ability to oxidize almost all kinds of organic and polymer as 2-propanol [367]. The curves for both acetaldehyde and 2-
materials, kill microbes, and mineralize these substances with propanol coincided over a wide range of Inorm values (>10−4 s−1 ),
the aid of molecular oxygen. These reactions mainly occur on or except for the region of Inorm below 10−4 s−1 , where the QY
very close to the surface of the TiO2 photocatalyst. The remote values for acetaldehyde degradation continued to increase with
photocatalysis also contributes to the oxidation reactions, but as decreasing Inorm and reached 180% for the value of Inorm of 3 ×
mentioned above, at rates one or two orders of magnitude lower. 10−5 s−1 due to the involvement of the radical chain pathway.
It should be pointed out that, aside from the light-limited con-
4.2.1. Decomposition of gaseous pollutants dition, over a wide range of experimental conditions, photocat-
Under UV illumination, TiO2 is able to oxidize gaseous alytic reactions on TiO2 follow reasonably well-known behavior,
pollutants in quantum yields ranging from 1% to over unity, which i.e., first-order kinetics in the adsorbed concentration of the organic
has important applications in cleaning indoor and outdoor air [11]. compound and an order of α in light intensity [6].
Given that there is a sufficient supply of O2 and UV illumination, r = kΓ I α . (4.20)
gaseous organic substrates can be completely degraded into CO2 ,
H2 O, and mineral acids on a TiO2 thin film. However, the removal Here, r is the reaction rate, k the first-order rate constant, Γ the
rate is influenced by numerous parameters; these include light concentration per unit real surface area, and I the light intensity.
intensity, substrate concentration, O2 partial pressure, humidity, Under light-rich conditions, the value of α is between 0 and
substrate type, and so on. Also, in many cases, the generation of 1 [239]. This was explained by considering the recombination
relatively stable reaction intermediates lowers the removal rate of charge carriers to surface-adsorbed intermediate products
and even stops the reaction through blocking active sites [365]. [239]. Conversely, the light-limited reaction rate is represented
Air–solid photocatalytic reactions are influenced by substrate as α = 1. Many organic compounds, as well as noxious gases
concentration and incident light intensity in a cooperative such as H2 S and NH3 , follow Langmuir adsorption behavior, so
552 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

that this simple equation can be converted into the familiar amounts of carbonaceous intermediates and byproducts on the
Langmuir–Hinshelwood form by substituting the expression for Γ : catalyst surface. Sano et al. reported that Pt–TiO2 could be free from
deactivation in the removal of acetaldehyde and toluene through
Γ −1 = Γ0−1 + (Keq c )−1 (4.21) a combination of photocatalysis and thermocatalysis [371]. They
where Γ0 is the surface concentration at full coverage, Keq the combined Pt–TiO2 photocatalysts with a solar concentrator, by
adsorption equilibrium constant and c the gas-phase or liquid- which the photocatalyst could be heated to about 200 ◦ C. At
phase concentration [6,11,15]. this temperature, Pt could effectively decompose the intermediate
The O2 concentration also greatly influenced the mineralization products and prevent the TiO2 from deactivating.
of gaseous organic substances. Durrant et al. suggested that
an oxygen concentration of 5% is sufficient to suppress charge 4.2.2. Decomposition of aqueous pollutants
recombination through capture of photogenerated electrons [238]. Most of the organic pollutants in water can be completely de-
However, the study of Lichtin et al. clearly showed that high O2 composed and mineralized at the surface of UV-excited TiO2 pho-
concentration facilitates the complete mineralization of organic tocatalysts; these include alkanes, haloalkanes, aliphatic alcohols,
pollutants [368]. Lichtin et al. measured the molar ratio of CO2 carboxylic acids, alkenes, aromatics, haloaromatics, polymers, sur-
produced to vaporized CH3 OH removed after 15 min of irradiation factants, herbicides, pesticides, and dyes [11,365,379–382]. While
in different O2 concentrations. They observed that in 20% O2 , only UV light and O2 are necessary for the reactions, many factors
generation of the stoichiometric yield of CO2 took twice as long such as light intensity, pH, ions, photocatalysts, kinds and concen-
as complete removal of CH3 OH, while in 100 mol% O2 the times trations of substrates, etc., have a great influence on the efficiency
required for complete removal of CH3 OH vapor and for formation of the mineralization process. As in the gas–solid system, photocat-
of CO2 were similar. This supports the Gerischer–Heller proposal alytic reactions in water work best at room temperature, and thus
that the photocatalytic oxidation of organic pollutants is mostly no heating is needed.
mediated by molecular oxygen, as mentioned in Section 4.1.5. Serpone and co-workers studied the quantum yields of
Humidity influences the removal rate in a complicated way liquid–solid photocatalytic reactions on TiO2 slurry photocatalysts
and appears to be dependent on the concentration and type of [239,383–385]. Similarly to the study of Ohko et al., they found
organic compound. Obee et al. reported that ethylene oxidation that maximum quantum yields could be obtained under light-
rates decreased significantly as the water vapor concentration was limited conditions, that is, low light intensity and relative high
increased from 1000 ppmv [369]. The influence of water vapor on substrate concentration [239]. The maximum QY (365 nm) for
the reaction rate derived from the low adsorption of ethylene due phenol degradation was measured as 14% using Degussa P25
to its low adsorption affinity relative to water. Sano et al. observed as a photocatalyst at pH 3, and this value varied over a range
that the removal rate of acetaldehyde was higher in dry air than in of 3.5%–30% for six types of TiO2 photocatalysts tested [383,
humidified air (50% relative humidity) [370]. However, Einaga et al. 384]. Aside from the light-limited condition, over a wide range
found that in the cases of benzene, toluene, and hexane, removal of experimental conditions, liquid–solid photocatalytic reactions
rates decreased with decreasing humidity, due to the increasing on TiO2 can also be described by Eq. (4.20), the same as for
amounts of carbon deposits on TiO2 [371]. Blount and Falconer gas–solid photocatalytic reactions [239]. The QYs measured in
also observed the same phenomenon in toluene photocatalytic these conditions varied over the wide range of 0.1%–10%, which
oxidation [372]. Recently, Jo and Park reported that in the case was several times lower than those of gas–solid photocatalytic
of very low-level contamination in indoor air (ppb level), there is reactions [4,11,379]. The low QY should restrict the applications
no appreciable dependency on humidity for removal of volatile of TiO2 photocatalysis in water purification. The intrinsic reasons
organic compounds (VOCs), which results from the competitive for the lower QY in water vs. air are yet to be explained in detail.
adsorption of water and the substrate compounds on the TiO2 Pichat et al. studied the influence of the sintering process on the
surface [373]. photocatalytic removal rate of various organic pollutants in water
The removal rate is also influenced by the type of substrate com- [386–388]. While the increase of sintering temperature is expected
pound. A comparison of reaction rates between 0.1 to 100 ppmv for to decrease the recombination rate of charge carriers; it does re-
toluene, n-butane, 1,3-dichloroethene, n-hexane, methyl-ethyl- sult in a decrease in surface area. Their purpose was to examine
ketone, n-decane, propanal, 1-butanol, trichloroethylene, ethy- the net effect of the sintering process on photocatalytic reactions.
lene, ammonia and phosgene showed that the rates, from the most In one study, they compared the removal rate of three chlorophe-
to least reactive, differed by about a factor of fifteen at any particu- nolic compounds and one chloroaliphatic acid compound on four
lar concentration [11]. Substrates such as trichloroethylene, whose TiO2 samples, which were all obtained identically by TiOSO4 ther-
degradation concerns radical propagation, are often degraded at a mohydrolysis with subsequent calcinations at various tempera-
high rate. The degradation of aromatic substrates is slow and suf- tures [388]. They found that the removal rate increased with the
fers photocatalyst deactivation due to the accumulation of reaction sintering temperatures for the three chlorophenolic compounds,
intermediates. whereas it was the opposite for the aliphatic acid compound. They
As often mentioned in the literature, photocatalyst deactivation suggested that the hole attack mechanism for carboxylic acids is
is a general problem in the removal of gaseous pollutants, due much more sensitive to surface area variation than would be the
to the accumulation of intermediate products or products on the •OH radical mechanism for cholorophenolic compounds, which
TiO2 surface [8]. Some deactivations are reversible; the deactivated can react in the near-surface solution phase. Their studies demon-
photocatalyst could recover activity after UV illumination in clean strate suggest the difficulty in finding a high-efficiency photocata-
humid air [8,372,373]. Some deactivations are irreversible; these lyst versatile for all types of pollutants in water [386].
are often observed in the removal of gaseous pollutants containing Generally, loading of noble metal co-catalysts such as Pt, Au, and
heteroatoms such as S, N, P, and Si, etc. [374–377]. In such cases, Pd, etc., enables accelerating liquid–solid photocatalytic reactions
mineral acid products block the active surface sites, and thus water [389]. These co-catalysts could enhance the charge separation
washing is needed to regenerate the deactivated photocatalyst. and catalyze the oxygen reduction reaction. Recent studies
Loading noble metals can retard the deactivation of TiO2 suggest that addition of electron receptors, such as hydrogen
photocatalysts. Einaga et al. reported that Rh0 could greatly peroxide, ozone, persulfate, etc., also works well for improving
improve the catalyst durability in benzene photo-oxidation [378]. reaction rates [382,390,391]. These electron receptors are more
The role of Rh0 supported on the TiO2 surface is to reduce the reducible than molecular oxygen. Moreover, when accepting an
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 553

Table 4.3
Photodecomposition of various organic compounds on TiO2 film [400].
Samples Stoichiometric formula Melting point (◦ C) Light intensity (mW cm−2 ) Rate constants (µg cm−2 h−1 )

Octadecane C18 H38 28.28 0.8 0.70


Stearic acid C18 H36 O2 69.3 0.8 0.75
Glycerol trioleate C57 H104 O6 −5 1.1 6.6
Glycerol C3 H8 O3 17.8 1.1 31
Liquid paraffina – −12.5 1.1 7.6
Salad oilb – −5 to −10 1.1 6.2
PEG (MW 600 D) (C2 H6 O2 )n 15–25 1.1 17
PEG (MW 500kD) (C2 H6 O2 )n 57–65 1.1 12
a
Including 15–20-carbon hydrocarbons, alkyl naphthalene hydrocarbons.
b
Including oleic, linoleic, linolenic, palmitic acids and stearic acid.

Fig. 4.29. (a) Pots of weight changes vs. illumination time for the photodecomposition of octadecane under 0.8 mW cm−2 UV illumination. (b) Plots of weight changes vs.
illumination time for the photodecomposition of glycerol trioleate (•), glycerol (), liquid paraffin (+), and salad oil (×) under 1.1 mW cm−2 UV illumination [400]. ©2000,
Elsevier Sciences B.V.

electron, these chemicals will dissociate into highly reactive The decomposition rates of oily compounds were found to
radicals in subsequent reactions, which can also participate in the be dependent on the square root of light intensity [400]. Even
photocatalytic reactions [382,390,391]: in a dry atmosphere, the organic compounds were effectively
decomposed, suggesting that the water that is generated during
H2 O2 + e− → OH− + •OH (4.22) the decomposition can be effectively used. The decomposition
O3 + e + H− +
→ O2 + •OH (4.23) was believed to occur through a mechanism in which tetraoxide
intermediates are involved (Fig. 4.18). The hydrocarbon chains can
S2 O28− +e →−
SO24− + SO4•− . (4.24) be attacked at any point along their length, with C–C bond scission
and subsequent decomposition via the resulting aldehydes [235].
The decomposition process proceeds at significant rates, with the
4.2.3. Decomposition of liquid and solid films
evolution of CO2 as the only detectable gas-phase product. The
The photocatalytic decomposition of monolayers, multi-layers, quantum yield for stearic acid decomposition could be over 20%,
and thin films of organic substances has been well studied for as estimated by Mills et al. [404].
the purpose of development of self-cleaning surfaces [5,9,10, Lee et al. studied the solid-phase photocatalytic reaction on
392–404]. Minabe et al. compared the decomposition of several the carbon black soot/TiO2 interface [353]. They observed that
liquid and solid organic films (∼500 nm thick) on TiO2 surfaces, the bulk of the soot layer with ∼2 µm thickness was completely
under about 1 mW cm−2 UV illumination (Table 4.3) [400]. All oxidized to CO2 over a 30-h irradiation. The presence of O2 was
of the compounds exhibited pseudo-zero-order kinetics in the essential in the photocatalytic soot oxidation. In another study,
initial stage of the reactions. This is reasonable, since the effective Cho et al. studied the solid-phase photocatalytic degradation of
concentration of reactant is high and can be approximated by a polyvinylchloride–TiO2 polymer composites [355]. Irradiating the
constant value. Carbon dioxide was the only detectable gaseous composite film for 300 h in air reduced its average molecular
product during the experiments. The weight change vs. time weight by two-thirds and weight by 27%. The photocatalytic
behavior during the initial stages for octadecane and stearic degradation of the composite film was accompanied by an increase
acid yielded similar rates, i.e., thinning rates of 9–11 nm h−1 in the FT-IR carbonyl peak intensity, the evolution of CO2 and
(Fig. 4.29). The decomposition rate (ca. 70 nm h−1 ) for glycerol H2 O, and whitening due to visible light scattering from growing
trioleate, however, was much higher than those for octadecane cavities. SEM images of the irradiated composite films showed the
and stearic acid. Other liquid compounds, including liquid paraffin, development of cavities around the imbedded TiO2 particles and
salad oil, glycerol, and even polyethylene oxide, etc., exhibited indicated that active oxygen species which were photogenerated
degradation rates similar to that for glycerol trioleate. Thus, the on TiO2 surface desorbed and diffused across a few micrometers to
degradation rates of organic films are dependent on their physical react with the polymer matrix.
states, i.e., solid or liquid film. The higher rates for liquid organic
compounds are reasonable, since organic compounds in the liquid 4.2.4. Photocatalytic sterilization
state can remain in more intimate physical contact with the TiO2 TiO2 photocatalysts have been found to kill cancer cells,
surface than those in the solid state. bacteria, viruses, and algae under UV illumination [10,405–410].
554 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 4.30. AFM images of E. coli cells on a TiO2 film: (a) no illumination; (b) illumination for 1 day; (c) illumination for 6 days; light intensity was 1.0 mW cm−2 . The outermost
layer, labeled as ‘‘A’’ in (a), disappeared after 24 h of illumination, with no layer visible in image (b) at the designated positions [414]. ©2003, Elsevier Science S.A.

Fig. 4.31. Schematic illustration of the three stages in the process of E. coli photokilling on a TiO2 film. In the lower row, part of the cell envelope is magnified [414]. ©2003,
Elsevier Science S.A.

This results in important applications in the disinfection of of bacteria on the illuminated TiO2 surface could be divided into
air, water, and surfaces with TiO2 photocatalysis. Kikuchi et al. three stages (Fig. 4.31): (1) disordering of the outer membrane of
studied the killing of Escherichia coli on TiO2 thin films [347]. A bacteria cells by reactive species (•OH, H2 O2 , O•−
2 ); (2) disordering
typical experiment involves placing 150 microlitres of an E. coli of the inner membrane (the cytoplasmic membrane) and killing of
suspension, containing 3 × 104 cells, on an illuminated TiO2 - the cell; and (3) decomposition of the dead cell. In the first stage,
coated glass plate (1 mW cm−2 UV light). Under these conditions, the outer membranes of E. coli cells were decomposed partially by
there were no surviving cells after only 1 h of illumination. Even the reactive species produced by the TiO2 photocatalyst. During
when the cells were separated from the TiO2 film by a 50-µm this stage, cell viability was not lost very efficiently. The partial
PTFE membrane, with 0.4 µm pores, the cells were still killed decomposition of the outer membrane, however, changes the
remotely: after 4 h, there were none remaining. By contrast, after permeability to reactive species. Consequently, reactive species
4 h under UV illumination without a TiO2 film, only 50% of the easily reach and attack the inner membrane, leading to the
cells were killed. Further studies showed that various bacteria, peroxidation of the membrane lipid. The structural and functional
E. coli, Staphylococcus aureus, and Pseudomonas aeruginosa, etc., disordering of the cytoplasmic membrane due to lipid peroxidation
were killed rapidly on TiO2 surfaces under UVA illumination led to the loss of cell viability and cell death. If the illumination
(320–380 nm, ∼1 mW cm−2 ) [9]. Kühn et al. observed that the continued for a sufficiently long time, the dead cells were found to
killing rates of bacteria were dependent on the thickness and be decomposed completely.
structure of cell walls [411]. Bacteria having thin cell walls, such as Ordinary TiO2 surfaces do not work well in killing bacteria
E. coli, and P. aeruginosa were killed much faster than bacteria such under low-intensity UV illumination. However, when there are
as Candida albicans that have a thick eukaryotic cell wall [411]. Cu particles dispersed on the TiO2 surfaces, the hybrid surfaces
Sunada et al. observed that not only bacteria are killed on the exhibit excellent self-sterilizing properties even under room light
TiO2 surface by photocatalytic action, but also the toxic ingredient that contains only a very small amount of UV light (∼1 µW cm−2 )
of bacteria can be decomposed [412]. If the UV illumination (Fig. 4.32) [416]. Although metallic Cu is a kind of antibacterial
continued for a sufficiently long time, the bacteria were found material that works even in the dark, it is not the only reason for
to be mineralized completely into CO2 , H2 O and other mineral the observed antibacterial effect of the hybrid films, since some Cu-
substances [413–415]. This is a unique property of photocatalytic resistant bacteria can be effectively killed on the TiO2 –Cu hybrid
sterilization compared to that of other antibacterial agents. surface under indoor illumination [416]. Comparative studies
Besides bacteria, viruses and fungi can also be killed and totally showed that metallic Ag also enhanced the antibacterial effect of
mineralized by photocatalytic action [408,413], but the killing of TiO2 film, but metallic Pt did not, even though Pt could improve the
fungi sometimes is much slower than that of bacteria because of photocatalytic performance of TiO2 . Based on these observations,
their chemical stability [413]. Sunada et al. proposed a mechanism for the enhanced antibacterial
Sunada et al. studied the photokilling process of E. coli by performance of TiO2 –Cu hybrid film that emphasizes the effect
means of AFM (Fig. 4.30) [414]. They found that the photokilling of cuprous ions [416]. Under UV illumination, cuprous ions may
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 555

such as N [434–464], C [221,465–474], S [475–481], B [482], P


[483], and F [484–487], were reported to shift the absorption
threshold of TiO2 into the visible range and provide the TiO2
material with photocatalytic activity under visible illumination.
Various mechanisms have been postulated to explain the shift of
absorption threshold and the photocatalytic activity of non-metal-
doped TiO2 photocatalysts [435,449,450,452,459]. These will be
summarized in the sections below.

4.3.1. Non-metal doping


N-doped TiO2 . In a paper published in 1986, Sato described
that impurity NH4 Cl shifted the absorption threshold of TiO2
into the visible light region upon calcination [434]. The as-
calcined material, assigned to NOx -doped TiO2 , showed higher
photocatalytic activity for oxygen isotope equilibration and the
oxidation of carbon monoxide and ethane than pristine TiO2 in
the visible light region. This work, however, was not noticed for
quite a long time. Fifteen years after Sato’s paper, Asahi et al.
reported visible light photocatalysis with N-doped TiO2 [435].
Fig. 4.32. Changes in survival of copper-resistant E. coli cells on a Cu/TiO2 They prepared the photocatalysts by two methods: one was by
film under dark condition (•) and under UV illumination at light intensities of sputtering a TiO2 target in an N2 /Ar gas mixture; the other was
7 µW cm−2 () and 1 µW cm−2 (N) [416]. by calcining TiO2 powder in an NH3 /Ar atmosphere. Both methods
© 2003, American Chemical Society.
produced yellowish TiO2 photocatalysts that noticeably absorbed
light at less than 500 nm and showed activity under visible
exist on the TiO2 surface through the oxidation of metallic copper illumination for decomposition of acetaldehyde and decoloration
with photogenerated holes. The cuprous ion may play two roles of methylene blue solution (Fig. 4.33). Nitrogen atoms were
in killing bacteria: (1) it increases the yield of hydroxyl radical believed to substitute for lattice oxygen atoms, as evidenced by the
through a photo-Fenton process by reaction with photogenerated peak at 396 eV in the N 1s XPS spectra.
H2 O2 ; and (2) it diffuses into the cytoplasmic membrane of the Asahi’s work is considered as a breakthrough in visible light
bacteria and accelerates the lethal effect after the outer membrane photocatalysis. After that, researchers started to consider the
of the bacteria is destroyed by oxidizing species. doping TiO2 with non-metal atoms in order to obtain visible light
photoactivity, with N-doped TiO2 being the most studied system.
4.3. Visible-light-induced photocatalysis There have been a great volume of publications that deal with
the preparation of N-doped TiO2 by physical or chemical methods,
Due to their wide band-gaps, pristine anatase and rutile TiO2 including sol-gel [437,442,445,448,449,452,461], sputtering [440,
photocatalysts mainly absorb ultraviolet photons. However, solar 453], ion implantation [455,456], and plasma-enhanced chemical
light only contains a small amount of ultraviolet photons (about vapor deposition method [454]. Either substitutional or interstitial
5%), and room light lamps emit mainly visible photons. Therefore, nitrogen is assigned to the observed photocatalytic activity
starting one decade ago, efforts have been devoted to extending under visible illumination. Various spectroscopic techniques [449,
the spectral response of pure TiO2 material (mainly anatase TiO2 ) 451,452,457,462,463], surface analysis techniques [455,456], as
to visible light. These efforts have included doping TiO2 with well as theoretical calculations [435,447,449,450,456] have been
metal impurities [417–425], coupling TiO2 with narrow band-gap applied to N-doped TiO2 in order to elucidate the origin of the
semiconductors [426,427], preparing oxygen-deficient TiO2 [428– visible light photoactivity and to develop visible-light-responsive
433], and doping TiO2 with non-metal atoms (anion doping) [434– photocatalysts with higher activity.
462,221,463–487]. Especially in recent years, non-metal doping C-doped TiO2 . Khan et al. reported in 2002 that the controlled
has become the most attractive; various non-metallic elements combustion of Ti metal in a natural gas flame could produce

Fig. 4.33. (a) Optical absorption spectra of TiO2−x Nx and TiO2 films. (b) N 1s XPS spectra of TiO2−x Nx and TiO2 films. (c) CO2 evolution as a function of irradiation time (light
at zero) during the photodegradation of acetaldehyde gas under UV irradiation and visible irradiation [435].
© 2001, The American Association for the Advancement of Science.
556 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

C-doped TiO2 with an absorption threshold that was extended of 400–500 nm, and were able to oxidize nitrogen oxides
into the visible spectral range [465]. The preparation method under visible illumination. Justicia et al. prepared TiO2−x films
was similar to that first used by Fujishima and Honda in the by metal-organic chemical vapor deposition, using titanium
early 1970s for the photoelectrochemical splitting of water [54]. tetraisopropoxide as a Ti precursor and N2 as the carrier gas [430].
Khan et al. claimed that carbon substituted for some of the Their preparation condition was quite similar to several chemical
lattice oxygen atoms in their sample. Besides the lower band- preparations of non-metal-doped TiO2 materials, i.e., a reducing
gap energy than rutile (2.32 vs. 3.00 eV) estimated from the atmosphere was involved in the preparation [470,471].
absorption spectrum, the C-doped TiO2 material was reported
to photoelectrochemically split water with the unbelievably high 4.3.2. Origin of visible light photoactivity
photoconversion efficiency of 8.35% under Xe lamp illumination Although there has been much published work on non-metal-
[465]. Their work was critically commented on by several groups doped TiO2 materials, the origin of visible light photoactivity
[221,466–468]; the main criticism was the high photoconversion of these materials is still in debate [435,449,450,452,459]. The
efficiency. It should be noted that, in their paper Khan et al. controversies are focused on two issues: (1) the origin of
did not provide any evidence of carbon doping or photoactivity the absorption in the visible light region; and (2) in what
under visible illumination. A later study by Bard and co-workers state the non-metal atoms are doped in the TiO2 lattice. As a
showed that TiO2 nanotube arrays annealed under controlled CO representative, N-doped TiO2 has been examined intensively by
gas flow could split water under visible illumination, but they did various experimental and theoretical methods. In particular, recent
not try to estimate the photoconversion efficiency [470]. Recently, works that combine density functional theory calculations with
Grimes and co-workers have actively investigated carbon-doped surface analysis techniques and spectroscopic characterization
TiO2 nanotubes for photoelectrochemical water splitting and have have shed light on the origin of the visible light activity of N-doped
reported efficiencies as high as 6.8% [488–490]. TiO2 [449,452,456]. These also aid the understanding of other non-
Irie et al. reported the preparation of carbon-doped TiO2 metal-doped TiO2 materials.
by oxidative heating of TiC powder [469]. Either anatase or Origin of absorption in the visible light region. In their paper in
rutile type C-doped TiO2 was prepared through control of the 2001, Asahi et al. suggested that the N 2p level could mix with
heating temperature. The prepared materials were yellowish with the valence band of TiO2 , which is mainly made up of O 2p, and
noticeable shifts of the absorption threshold to the visible spectral this resulted in the narrowing of the band-gap and photocatalytic
region. The carbon dopant showed a peak at 281.8 eV in the activity in the visible light region (Fig. 4.33) [435]. They calculated
C 1s XPS spectra, which was assigned to substituted carbon. A
the densities of states of the substitutional doping of C, N, F, P, or
similar peak was also observed on the C-doped TiO2 samples
S or O in the anatase TiO2 crystal by the full-potential linearized
prepared by reactive sputtering [473] and ion-assisted electron-
augmented plane wave (FLAPW) formalism in the framework of
beam deposition methods [474]. The photocatalytic activity of C-
the local density approximation (LDA). Their calculation supported
doped TiO2 under visible illumination was quite weak, as evaluated
the idea of band-gap narrowing with N doping.
by the decomposition of gaseous 2-propanol [469].
The model of band-gap narrowing was soon challenged by
S-doped TiO2 . Umebayashi et al. reported that sulfur doping was several experimental and theoretical studies on N-doped TiO2
unable to extend the absorption of TiO2 into the visible region materials [437–441,443,447,459]. As commented by Serpone, it
[475–477]. They prepared S-doped TiO2 either by oxidative heating is hard to believe that a low level of doping (<2 atomic %) can
of TiS2 powder [475,476] or by sulfur-ion implantation [477]. In rigidly shift up the valence band [459]. Irie et al. reported that
either case, a red-shift of the absorption threshold was observed the quantum yields of gaseous 2-propanol decomposition on N-
compared to pristine TiO2 , and the prepared materials were able doped TiO2 powders under visible illumination were several times
to generate photocurrent and to decolor methylene blue solution lower than those under UV illumination [438]. This observation
under visible illumination. Umebayashi et al. explained the could not be explained by the band-gap narrowing model. Instead,
observed photoactivity in the visible spectrum to the substitution it suggested that nitrogen doping led to the formation of localized
of oxygen atoms with sulfur, which caused the narrowing of the midgap states above the valence-band edge (Fig. 4.34). Torres
band-gap [477]. et al. studied the photoelectrochemical behavior of N-doped TiO2
Ohno et al. found that S-doped TiO2 could be easily prepared by for water oxidation [440]. They observed that UV light gives a
a chemically modified sol-gel process [478–480]. They prepared higher probability than visible light for the electrons and holes to
the precursor powder by mixing titanium isopropoxide with leave the electrode and contribute to a photocurrent and oxidizing
thiourea; heating the powder produced yellow-colored S-doped water, which is consistent with the conclusion of Irie et al. [438].
TiO2 . The oxidation states of sulfur in Ohno’s samples, however, Nakamura et al. observed that an N-doped TiO2 photoelectrode
were S4+ and/or S6+ , inconsistent with the results of Umebayashi could oxidize SCN− and Br− under UV illumination; however,
et al. [479]. Sakthivel and Kisch followed the experiments of Ohno no oxidation was observed under visible illumination [439]. This
et al.; however, they did not find any evidence of sulfur doping. study also points out the possibility that the doped nitrogen species
Instead, they claimed the formation of N-doped TiO2 [442]. This gives rise to a midgap level slightly above the top of the valence
disagreement suggests the complexity of the chemistry involved band. A similar conclusion was drawn by Tachikawa et al. in
in the preparation, and the possible misreading of characterization transient spectroscopy studies on N-doped TiO2 [457], as well as
results. C- and S-doped TiO2 [481].
Others. There are several reports that deal with doping TiO2 with Lin et al. used the spin-polarized plane-wave pseudopotential
B [482], P [483], or F [484–487] atoms by the sol-gel method. method, based on density functional theory, to calculate the
The doped TiO2 materials are all yellowish, similar to N-, C-, or electronic-band structures and the optical absorption spectra of
S-doped TiO2 , and have been reported to exhibit activities for nitrogen-doped TiO2 [450]. According to their calculation results,
decomposition of organic substances. substitutional nitrogen should give rise to localized N 2p acceptor
TiO2−x , the oxygen-deficient form of TiO2 , has also been states above the valence band, and absorption in the region
studied extensively as a visible-light-responsive photocatalyst between 400 and 500 nm in the calculated spectrum. Moreover,
[428–433]. Nakamura et al. reported the preparation of TiO2−x the calculation reveals that even for high nitrogen concentration
through the H2 plasma treatment of TiO2 powders [428]. The (12.5% doping), the N 2p states are still localized, lying slightly
TiO2−x photocatalysts showed weak absorption in the region above the top of the O 2p valence band. Their conclusion is in
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 557

later synchrotron-based photoemission studies, Nambu et al. also


reached the same conclusion on the N-doped rutile (110) surface
[456]. They compared the valence-band spectra of Ar+ -reduced
and N-implanted rutile TiO2 (110), and found that both spectra
contained new states just below the Fermi level EF that could be
assigned to the occupied 3d states of Ti3+ , but for the N-implanted
TiO2 , they observed extra features above the top of the O 2p band,
which shifted the edge by ∼0.5 V towards EF (Fig. 4.35) [456].
Considering that the conduction-band edge is little influenced by
nitrogen doping [442], the N-derived states should result in a red-
Fig. 4.34. Various schemes illustrating the possible band-gap electronic structure
and excitation processes of visible-light-responsive TiO2 materials: (a) pure TiO2 ;
shift of the absorption edge of N-doped TiO2 materials.
(b) band-gap narrowing model for non-metal-doped TiO2 ; (c) oxygen-deficient The doping of nitrogen facilitates the formation of oxygen
TiO2 ; (d) localized midgap level model for non-metal-doped TiO2 ; (e) oxygen vacancies in TiO2 materials, which is observed as midgap
vacancy levels and non-metal-doped midgap levels are considered together. states below the conduction band [440,449,451,452,455,456]. This
explains why N doping does not change the n-type semiconducting
TiO2 to p-type. In the case of the rutile (110) surface, N doping led
to a 1 × 2 type surface reconstruction, triggered by the increased
formation of oxygen vacancies [455]. EPR and UPS studies, and
DFT calculations revealed the existence of attractive interactions
between the dopants and the O vacancies [449,452,456]. The
presence of nitrogen dopants was found to reduce the formation
energy of oxygen vacancies [451,452]. At the same time, the
existence of O vacancies stabilized the N impurities [449,452,456].
When oxygen vacancies and N impurities are together, there is an
electron transfer from the higher energy 3d band of Ti3+ to the
lower energy 2p band of the nitrogen impurities [449,452,456].
Oxygen vacancies also contribute to the absorption in the
visible light region (Fig. 4.36) [450]. This point was often neglected
in studies on non-metal-doped TiO2 [459,460]. Lin et al. calculated
the absorption spectra of oxygen-deficient TiO2 and found that the
entire absorption edge of oxygen-deficient TiO2 is red-shifted by
about 20 nm with the dominant visible light absorption occurring
above 500 nm (Fig. 4.36) [450].
Doping states of nitrogen atoms. As the direct evidence of
nitrogen doping, XPS studies on N-doped TiO2 materials often give
two kinds of N 1s peaks, one at about 396 eV, and the other one at
about 400 eV [452]. Sato observed a peak at 400 eV for TiO2 powder
heated in the presence of ammonium chloride and assigned the
peak to an NOx species [434]. Asahi et al., however, observed an
additional peak at 396 eV for a TiO2 film prepared by the sputtering
method in addition to the peak at 400 eV [435]. They assigned
the 396 eV peak to substitutional N doping, and argued that
Fig. 4.35. Valence-band spectra for Ar+ -reduced and N-implanted rutile TiO2 only substitutional N is responsible for the observed photoactivity
(110), acquired with a photon energy of 325 eV. Dashed lines denote the reference under visible illumination, since the 400 eV peak is also observable
spectrum from stoichiometric TiO2 (110) [456]. for pristine TiO2 powders. In later studies, the peak at about 400
© 2006, American Institute of Physics.
eV was frequently observed, in particular for the samples prepared
by chemical methods [438,442,444,452,461]. In some cases, only
contrast to the spin-restricted FLAPW calculation of Asahi et al. the peak at 400 eV was observed for N-doped TiO2 materials
[435], but is consistent with spin-polarized calculations [447] and that exhibited photocatalytic activity under visible illumination
several experimental results [437–441]. They also found that at [442,452,461]. It was thus believed that at least two kinds of
least 20% N doping is needed to mix the N 2p state with the O 2p nitrogen doping are responsible for the visible light photoactivity.
valence band, but in practice such a high doping level would result The 396 eV peak is generally assigned to substitutional nitrogen
in the formation of TiN [448,462]. doping; while the 400 eV peak is believed to related to interstitial
The formation of localized midgap states is further proven nitrogen doping.
by recent UV-photoelectron spectroscopy (UPS) studies on N- Di Valentin et al. studied N-doped TiO2 powder prepared by the
implanted TiO2 single crystals [455,456]. Batzill et al. doped sol-gel method with EPR combined with DFT calculations [449].
rutile and anatase single crystal surfaces by low energy nitrogen They observed the presence of two slightly different nitrogen
ion implantation (N. 3 at.%); the single N 1s peak at a binding species in the EPR spectra. On the basis of comparison with the DFT
energy of 396 eV suggests that the doping is of a substitutional calculations, one of the two species was assigned to substitutional
type [455]. Ultraviolet photoemission studies show that N doping nitrogen atoms, and the other was assigned to interstitial nitrogen
induces localized N 2p states within the band-gap just above in the form of NO (Fig. 4.37). In another study, Reyes-Garcia
the valence band, but no evidence of band-gap narrowing is et al. studied N-doped TiO2 photocatalysts with 15 N solid state
observed in this study. For the rutile (110) surface, N doping shifts NMR and EPR characterization [461]. Their samples were prepared
the edge up by 0.4 eV; for the anatase surface, the N-derived by high temperature nitridation of TiO2 nanoparticles in the
states span the region in the band-gap from the valence band presence of urea, or by a sol-gel preparation with various N-
maximum up to the higher lying Ti 3d defect states [455]. In containing chemicals as nitrogen sources. Interstitial nitrogen
558 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Table 4.4
Preparation and characteristics of N-doped TiO2 films.
Author (Ref.) Preparation method Description Photocatalytic activity

Asahi et al. [435] Sputtering deposition in N2 /Ar, Or Absorption onset: 500 nm; N1s peaks Decompose methylene blue and acetaldehyde in
annealing TiO2 powders in NH3 at at 396 eV and 400 eV; nitrogen visible irradiation, with similar activity in UV
873K concentration: 1–1.4 atomic %. irradiation as non-doped TiO2 .
Burda et al. [446] Nitriding TiO2 nanoparticles with Absorption onset: 600 nm, N1s peak at Decompose methylene blue in solution by visible
triethyl amine. 401.3 eV laser.
Sakthivel Sakthivel et al. [443] Hydrolyzing TiCl4 in water with Absorption onset: 520 nm; a weak N1s Degradation of 4-chlorophenol, benzene and
addition of ammonia, post-annealing peak at 404 eV acetaldehyde with visible light (λ > 455 nm),
at 673 K.
Irie et al. [438] Annealing TiO2 powders in NH3 N-substituted for oxygen: 0.5–1.9% Degradation of gaseous 2-propanol under visible
light irradiation; UV light activity was weaker for
doped samples than non-doped. Visible light
activity was several times lower than UV light
activity.
Mrowetz et al. [441] One sample (a) was prepared by Increased absorption in the visible Both samples A and H showed no apparent
Burda’s method, the other sample (H) light region for both sample A and activity for degradation of HCOO− and NH3 · H2 O
was prepared similarly to Asahi’s and sample H. in visible irradiation; sample A was bleached
Irie’s works. with UV irradiation.
Diwald et al. [444] Annealing rutile (110) in NH3 at 873 K. N1s peaks at 399.6 and 396.7 eV; Photodeposition of Ag from AgNO3 solution in
increased absorption in the range of visible irradiation.
2.4–3.0 eV.
Maeda et al. [454] PECVD in NH3 /Ar using Ti(Oi Pr)4 as N1s peaks at 396 eV and 399.3 eV, Degradation of stearic acid molecules under
precursor; post-annealing over 673 K. Area ratio of the two peaks is 4:1. visible irradiation.
Kitano et al. [453] RF Magnetron sputtering in N2 /Ar N-substituted for oxygen: 2–16.5%; Degradation of 2-propanol solution with visible
photo-response up to 550 nm. irradiation; 6% N-substitution is the best.
Yates et al. [491] Chemical vapor deposition in NH3 /N2 , Nitrogen atomic concentration: 1.5–5% No visible light activity for degradation of stearic
using TiCl4 and ethyl acetate as acid molecules; weak UV-activity.
precursor.

atoms were observed in their studies, and these were assigned to nitrogen species in parallel with oxygen vacancies could be favored
nitrates (NOx ). [449].
Di Valentin et al. calculated the electronic-band structures for
substitutional and interstitial N-doped anatase TiO2 [449]. For both 4.3.3. Activity and stability of N-doped TiO2 photocatalysts
of these N-doped model systems, the formation of localized states Asahi reported that nitrogen doping offered visible light (VL)-
in the band-gap was predicted (Fig. 4.37). Substitutional nitrogen photocatalytic activity to TiO2 , without any loss of UV-activity, for
states lie just above the valence band, while interstitial nitrogen the decoloration of methylene blue solution and degradation of
states lie higher in the gap [449]. The localized nature of the gaseous acetaldehyde [435]. However, many studies after Asahi’s
N-induced states has the consequence that the hole generated work showed that nitrogen doping only offered modest VL-activity
by visible irradiation is less mobile than that generated by UV to TiO2 , but at the same time lowered the UV-activity (Table 4.4)
irradiation. In particular, the interstitial N impurities, which give [438,442,443,463,464]. Some studies even showed almost no VL-
rise to the higher energy states in the gap, might behave as stronger activity for N-doped TiO2 (Table 4.4) [441,491]. It thus seems that
hole trapping sites, reducing the direct oxidation power of the the preparation history, which dictates the types and level of
sample in the photocatalytic process [449]. The calculations of nitrogen doping and the concentration of oxygen vacancies, greatly
Di Valentin et al. also suggested that there is a cost to induce influences the photocatalytic activity.
substitutional nitrogen formation starting from the interstitial one. Mrowetz argued that the often-used probe, methylene blue,
In an excess of oxygen and nitrogen, interstitial nitrogen doping is is not a good indicator for photocatalytic activity [441]. Methy-
definitely preferred. However, under highly reducing conditions, lene blue solution could be decolored via several pathways, in-
as is the case after annealing at high temperature, substitutional cluding oxidation with holes or hydroxyl radicals, reduction with

Fig. 4.36. (a) Calculated optical absorption spectra for various N concentrations in the polycrystalline TiO2 : (I) undoped TiO2 ; (II) 12.5% nitrogen doped; (III) 6.2% nitrogen
doped; and (IV) 3.1% nitrogen doped. (b) Calculated optical absorption spectra of polycrystalline TiO2 with different O vacancy contents: (I) undoped TiO2 ; (II) 12.5% oxygen
vacancies; (III) 6.2% oxygen vacancies; (IV) 3.1% oxygen vacancies [450].
© 2005, American Chemical Society.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 559

types of Nb• centers give rise to localized states in the band-


gap of the oxide. The diamagnetic Nb− species are expected to
be more abundant than the paramagnetic Nb• ones, since they
are energetically favored. Irradiation with visible light at 437 nm
promotes electrons from these states to the conduction band and
thus leads to the increase of paramagnetic Nb• centers, as shown
in Fig. 4.38. When the irradiation at 437 nm is performed in an
oxygen atmosphere, the increase of the Nb• signal is accompanied
by the simultaneous appearance of new EPR lines corresponding to
a surface superoxide O•−2 radical species (Fig. 4.38), suggesting that
the photoexcited electrons are available for chemical interactions
at the surface [452]. The holes, formed by charge separation,
remain localized on the Nb• centers, which limits the VL-activity of
N-doped TiO2 [452].
Emeline et al. measured the QYs of phenol degradation over a
commercial VL-responsive TiO2 photocatalyst [458]. They found,
under optimum conditions, where recombination losses were
minimal, that the QYs at 365 nm and 436 nm were 12% and
Fig. 4.37. Schematic sketch and electronic-band structure for (a) substitutional and 8%, respectively. The QY of phenol degradation over Degussa P25
(b) interstitial N-doped anatase TiO2 . The calculated value of the band-gap is also was 14% at 365 nm measured by the same experimental system
reported [449].
© 2005, American Chemical Society.
[458]. Emeline’s work was done under ‘‘light-limited’’ conditions,
which is by far different from the ‘‘light-rich’’ conditions of actual
photocatalytic systems. However, this suggested that the VL-
conduction-band electrons, sensitized photodegradation, and ad-
activity of N-doped TiO2 could be comparable to the UV-activity,
sorption on catalyst surfaces [492–495]. Thus, they studied the provided that the recombinative losses introduced by oxygen
VL-activity of N-doped TiO2 by the oxidation of HCOO− anions in vacancies were suppressed and the absorption of visible light were
water [441]. They used two samples: the first, designated ‘‘A’’, was improved.
prepared by Burda’s method, that is, nitriding TiO2 nanoparticles Intense efforts have been devoted to the development of prepa-
with triethyl amine [445,446]; the second, designated ‘‘H’’, was ration methods for highly efficient N-doped TiO2 materials in terms
prepared by annealing TiO2 powders in NH3 with a post-annealing of increasing both the VL-absorption and the activity. Maeda and
in air. They found that sample A suffered bleaching under UV irra- Watanabe developed a plasma-enhanced chemical vapor deposi-
diation. They inferred that the reaction between the TiO2 nanopar- tion method to prepare N-doped TiO2 [454], which ensured the
ticles and triethyl amine at room temperature led to the formation substitutional doping of nitrogen, in contrast to the sol-gel method
of surface organotitanium complexes, rather than to substitutional or annealing in NH3 , which often produced predominantly inter-
N-doped TiO2 , the former being readily degraded by UV light [441]. stitially doped nitrogen [434,441,452,461]. Their film photocata-
Sample H, although doped with nitrogen atoms, did not show ap- lysts were able to decompose stearic acid molecules under visible
preciable VL-activity for the degradation of HCCO− and NH+ 4 ions irradiation, suggesting promising applications in VL-activated self-
[441]. Moreover, no appreciable hydroxyl radicals were detected cleaning materials [454].
with ESR for H under visible irradiation. Their results are contrast Kitano et al. developed a RF magnetron sputtering technique
to those of Asahi [435] and Irie [438], who used similar methods to prepare N-doped TiO2 films [453]. By their method, nitrogen
to prepare N-doped TiO2 materials. One possible reason is that the atoms were able to substitute oxygen in the range of 2–16.5%.
post-annealing in Mrowetz’s work might re-oxidize the surface of Such a high-level doping was not achieved previously and resulted
sample H, which has a detrimental influence on VL-activity. in a steep absorption edge in the visible light region [453]. A
Livraghi et al. studied the VL-activity of N-doped TiO2 by means film photocatalyst with a nitrogen concentration of 6% (strong
of EPR spectroscopy [452]. They found that nitrogen dopants form absorption up to 500 nm) exhibited the highest reactivity for
either diamagnetic (Nb− ) or paramagnetic (Nb• ) bulk centers. Both the photocatalytic oxidation of 2-propanol in water under visible

Fig. 4.38. (a) EPR signal intensity of Nb• (substitutional and interstitial N) and O•− •
2 in various conditions. I0 is the Nb intensity in the non-irradiated sample. ‘‘Blue’’ indicates
437-nm irradiation, while ‘‘Green’’ indicates 500-nm irradiation. (b) Sketch of the proposed mechanism for the processes induced by visible light irradiation of the N-doped
sample in O2 atmosphere [452]. Nb− and Nb• are nitrogen dopants forming paramagnetic and diamagnetic bulk centers, respectively.
© 2006, American Chemical Society.
560 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

irradiation (λ > 450 nm). Higher nitrogen concentration would


introduce large amounts of oxygen vacancies in the films and result
in lower VL-activity [453].
The stability of N-doped TiO2 is also an important issue
concerning actual applications. It was reported that N-doped TiO2
exhibited better thermal stability than oxygen-deficient TiO2 , as a
result of the interaction between the nitrogen dopants and oxygen
vacancies [437]. The latter lost VL-absorption due to oxidation
upon heating at 473 K [437]. However, Kitano et al. observed
that N-doped TiO2 lost VL-absorption gradually upon heating
at 573 K, accompanied by a decrease in nitrogen concentration
[453]. Kitano et al. also examined the stability of N-doped TiO2
film under irradiation [453]. They found that the film after
photoelectrolysis of water under visible irradiation showed a
decreased N concentration at the surface layer, indicating that the
surface of the film was oxidized during photoelectrolysis. These are
challenges for the further development of these types of materials.

5. Fundamentals of the photo-induced hydrophilic (PIH) effect

5.1. Overview

Our group first reported, along with co-workers from the TOTO
Corp., in 1997 on the phenomenon that we termed the ‘‘light- Fig. 5.1. (a) Friction force microscopic (FFM) image (5 × 5 µm) for a rutile (110)
induced amphiphilic surface’’ [496]. When a titania film was surface before UV illumination; (b) FFM image (5 × 5 µm) of the same surface after
illuminated with UV light, the contact angle for water decreased illumination; (c) a medium scale FFM image (1000 nm × 1000 nm) of the framed
to near 0◦ , and the same occurred also with organic liquids. We area in (b); (d) a higher-magnification topographic image (240 × 240 nm) of the
framed area in (c). For (d), the sample stage was rotated 45◦ from the position in (c)
expected that there would be several applications for this new
(taken from Wang et al. [514]).
effect, including self-cleaning surfaces and anti-fogging mirrors. © 1998, John Wiley & Sons.
With friction force microscopy, it was observed that there was
a spontaneous formation of a pattern of two different types of
(see later), and thus it was of interest to examine the influence of
areas with hydrophobic and hydrophilic properties (Fig. 5.1). We
the crystal face exposed. Our study with single crystals of rutile
hypothesized that the presence of two types of surfaces at the tens
found that the (110) and (100) faces had similar characteristics,
of nanometer scale was responsible for the high affinity for both
i.e., both became hydrophilic at about the same rate under UV
water and organic liquids, hence the term ‘‘amphiphilic surface’’.
illumination [499,500]. However, the (001) surface was found
The following year, we published a more detailed article
to be much slower in becoming hydrophilic. Under a pure O2
and included a proposed concept for the mechanism, based
atmosphere, it was found that, for all three surfaces, (110), (100)
in part on the observed infrared spectrum, which showed the
and (001), the effect was partly suppressed; indeed, with the
reversible growth and decay of a peak (3695 cm−1 ) assigned to the
(001) surface, the effect was almost completely suppressed. X-
formation of hydroxyl groups, presumably at sites where there had
been dissociative adsorption of water at oxygen vacancies [514]. ray photoelectron spectra showed that the (110) and (100) faces
Even though the ideas have become much more detailed in the acquired a small peak in the Ti 2p spectrum that was attributable
meantime, the same basic idea is still valid. The large number to Ti3+ sites, while the (001) face did not. Thus, the idea of UV
of studies that have been described in Section 3.4 have led to induction of oxygen vacancies was supported. However, other
a consensus that water is adsorbed molecularly under neutral possible explanations exist, as discussed later.
pH conditions on the stoichiometric rutile (110) surface. Thus, The photo-induced hydrophilic (PIH) effect was initially
the interaction with water is not particularly strong, and the thought to involve simply the photocatalytic removal of hydro-
introduction of hydroxyl groups would be expected to lead to a carbon contamination films, but this was found to be an inade-
surface with a stronger interaction with water. quate explanation. One of the experiments that underscored this
The effect of enhanced hydrophilicity was found to slowly involved the comparison of TiO2 and SrTiO3 (501). These were both
revert to the normal, hydrophobic state during storage of the film prepared as polycrystalline films on glass. Both were found to pho-
in the dark, this process requiring varying amounts of time. It was tocatalytically decompose methylene blue, but the SrTiO3 film did
found that ultrasonication of the film accelerated the reconversion not become hydrophilic under illumination (Fig. 5.2).
to the initial state; with low-intensity agitation (0.6 mW cm−2 ), In an effort to continue improving the characteristics of the PIH
the process was found to take ca. 140 min, while, with high- effect, variations were examined. One of the most promising ones
intensity agitation, it required only 20 min [497]. Subsequent UV involved the deposition of a thin film of WO3 on the TiO2 surface
illumination served to make the surface hydrophilic again. [427]. An amorphous film was found to be more effective than a
Most of the experiments that were carried out involved pure crystalline film. With the former, a PIH conversion was observed
anatase TiO2 films, but it was found that the addition of 10–30 wt% to reach completion after 100 min at the surprisingly low UV light
SiO2 to a TiO2 film improved the hydrophilic properties [498]. The intensity of 1 µW cm−2 .
presence of SiO2 helped the films to retain higher surface area and The amphiphilic effect is unique for TiO2 : no other oxide that
to suppress the thermal conversion of anatase to rutile during film has been examined thus far exhibits this behavior. In further
heat treatment. exploration of this effect, it was found that, the oleophilic behavior
One of the first mechanisms proposed included a step in which of TiO2 films was actually reconverted to oleophobic after about
bridging oxygens, such as those on the rutile (110) surface, were 10 h of continuous UV illumination at 1 mW cm−2 [502], while
thought to be removed from the surface during UV illumination the hydrophilic character remained. It was speculated that this
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 561

Fig. 5.2. Comparison of the photocatalytic and PIH behavior of TiO2 and SrTiO3 : (left panel) time dependence of absorbance under UV irradiation of methylene blue adsorbed
on TiO2 (solid diamonds), on SrTiO3 (open squares), and on an uncoated glass substrate (open triangles); (right panel) time dependence of water contact angle under UV
irradiation of TiO2 (solid diamonds), and SrTiO3 (open squares). UV irradiation was 1 mW cm−2 under ambient conditions (295 K; relative humidity, 60%, air) (taken from
[501].

behavior resulted from a redistribution of the nanometer-scale


hydrophobic (oleophilic) domains with time.
The effect of repeated cycles of UV illumination followed
by ultrasonication was studied for rutile (110) and (001) films
[503]. Interestingly, although the (001) surface was found to be
quite slow compared to the (110) surface in reaching hydrophilic
behavior, taking about three times as long, it exhibited an
increasingly rapid effect with cycling; by the third cycle, it reached
a fully hydrophilic state after less than 50 min under a UV light
intensity of 40 mW cm−2 . This was thought to be the result of
restructuring of the surface to expose more bridging oxygens.
The influence of electrochemical potential was examined, and
it was found that more positive potentials were effective in
accelerating the PIH effect [504]. This effect reached a plateau in
the 0.4–0.8 V vs. Ag/AgCl range in aqueous pH 6.86 buffer. As the
potential was held at increasingly negative potentials, starting at
about −0.2 V, the PIH effect began to be suppressed, with the
suppression increasing with negative potential. This phenomenon
led to a modification of the mechanism; it was concluded that
photogenerated holes are trapped at bridging oxygens, causing Fig. 5.3. Trace of water contact angle (WCA) vs. time for an anatase film sample,
them to be ejected from the surface, thus allowing a water showing reversible lowering of WCA under UV illumination and increasing WCA
molecule to dissociate at the resulting oxygen vacancy. This under visible light illumination (taken from [506]).
mechanism will be discussed in more detail below. It should © 2002, Elsevier Science.
be noted that the type of film studied in this work begins to
absorb light at longer wavelengths (monitored at 700 nm) as the generated over a wide range of potential under illumination at
film is maintained at potentials below about −0.8 V vs. Ag/AgCl, 360 nm and an intensity of 50 mW cm−2 , whereas the anatase
which is an indication that electrons are building up, either in the film generated about 1.5 mW cm−2 photocurrent at 0.8 V. The
conduction band or in electronic states close to the conduction hydrophilic effect for the amorphous film was thought to involve
band. In this region, it becomes more difficult to separate the Eq. (3.2), thus generating surface hydroxyl groups, which should be
photogenerated electrons and holes, since the film behaves more more hydrophilic. This is still not a complete explanation, however,
like a metal and less like a semiconductor. because, if the effect just involves the surface, the anatase film
The electrochemical potential was also found to have an should also produce nearly the same degree of hydrophilicity.
effect even in the dark, but this effect was rather weak for There is clearly some other factor involved, for example, roughness.
conventional polycrystalline anatase films [505]. However, for It was discovered that visible light illumination, without
amorphous titania films, the effect was much more obvious. In the UV, is actually effective in reconverting anatase films from a
potential region below −0.6 V vs. Ag/AgCl, the amorphous film hydrophilic state back to a semi-hydrophobic state (Fig. 5.3)
began to absorb light (monitored at 700 nm). At this potential, [506]. This effect was found to result from the visible-light-
the contact angle also began to decrease in the dark for the induced heating of the film. The highly reversible alternation of
amorphous film. At more negative potentials, the effect was more the hydrophilic and hydrophobic character is one of the most
pronounced: at −1.0 V, the contact angle approached 0◦ after convincing demonstrations that this effect is not simply due to
20 min. In contrast, for the polycrystalline film, even at a potential the build-up and destruction of contamination layers. These results
of −1.2 V, the contact angle only approached 12◦ after 20 min. were also confirmed by Stevens et al. [507].
The photo-response of the amorphous film to UV light was found The PIH effect has also been examined on other metal oxides,
to be very small, with only a small µA cm−2 photocurrent being for example, ZnO [508]. This oxide also exhibits a pronounced
562 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

effect. Initially, the ZnO surface is much more hydrophobic than texturing was carried out by first forming a layer of polystyrene
the TiO2 surface, exhibiting a water contact angle of ca. 110◦ . The spheres (250 nm), followed by intrusion of a suspension of 15-
contact angle for these films also increases during storage in the nm TiO2 particles, with the polystyrene spheres finally being
dark in a manner similar to that for TiO2 . It was proposed that the removed by heat treatment in air (Fig. 5.4). Other recent work,
ZnO surface can also be induced by UV light to generate oxygen including that aimed at enhancing the visible light sensitivity, will
vacancies. This point will be discussed later. As with TiO2 , ZnO also be discussed later.
reconverts to a hydrophobic state at different rates according to the
type of atmosphere. For example, the conversion is fastest in pure 5.2. Mechanisms of the PIH effect
O2 . This is thought to be due to a competition between O2 and H2 O
for reaction with oxygen vacancies. Some indications have already been given regarding the pos-
A number of other metal oxides were also examined for their sible mechanisms for the PIH effect. These will now be discussed
PIH activity and photocatalytic activity [509]. The following oxides in more detail: they include the original mechanism, which essen-
were found to exhibit the photocatalytic effect: TiO2 , SrTiO3 and tially says that the PIH effect and photocatalysis are identical; the
ZnO, with SnO2 exhibiting only a slight effect. The following were reductive mechanism, which was originally proposed, the later ox-
found to exhibit a PIH effect: TiO2 , SnO2 , ZnO, WO3 and V2 O5 . idative mechanism, and finally, a new combined redox mechanism,
Other oxides studied were: CeO2 , CuO, MoO3 , Fe2 O3 , Cr2 O3 and which is again similar to photocatalysis but includes both.
In2 O3 . All of the latter group exhibited neither effect. These results
are also one of the most convincing with regard to distinguishing 5.2.1. Decomposition of organic films
the two effects. As already noted, SrTiO3 is the clearest example The initial thought that has occurred to almost everyone
of an oxide that exhibits the photocatalytic effect but not PIH; when learning about the PIH effect is that it is a simple
in this work, there are two examples of oxides that exhibit PIH process of photocatalytic removal of organic contamination films,
but not photocatalysis: WO3 and V2 O5 . Similar to TiO2 , these which are notorious and ubiquitous. Our group first assumed
oxides are able to lose oxygen, creating oxygen vacancies to the this mechanism, but later evidence, as outlined above, made it
point where the structure can change, as in the Magneli phases. increasingly clear that the two effects, while very closely related
However, an exception to this correlation is MoO3 , which is also indeed, are separate.
known to be able to convert to Magneli phases. This explanation Several groups during the past several years have revisited this
requires modification, because the loss of substantial amounts of issue. White et al. used the rutile (110) surface and found that the
oxygen can also lead to metallic conduction, which suppresses presence of oxygen vacancies, produced by heating the crystal, did
photoeffects that normally occur in semiconductors. not produce a noticeable effect on the TPD behavior for water [516].
Both the stoichiometric and the slightly reduced surfaces were
In an effort to make the PIH measurements more precise, Sakai
highly hydrophilic. After the surfaces are covered with a monolayer
et al. used the reciprocal of the water contact angle to quantify
of trimethylacetic acid, they become highly hydrophobic, as
the conversion rates from hydrophobic to hydrophilic and back
evidenced by TPD curves that indicate the presence of liquid rather
[510]. In this work, a correlation was found between strongly
than adsorbed water. After these surfaces are illuminated with UV
adsorbed water, as observed as a component in the O 1s XPS
light for 300 s in the presence of 5 × 10−6 Torr O2 , they became
spectra, and the reciprocal of the contact angle. Temperature-
much more hydrophilic and again exhibited the behavior of the
programmed desorption was also used to measure the amounts
clean surface, i.e., with first-layer, second-layer and multi-layer
and types of adsorbed water. The reconversion of the surface to
water TPD peaks.
hydrophobic was found to exhibit a clear transition between a fast
The main conclusions of this work were that (1) the original
and a slow process, the latter being thought to involve a structural TiO2 surface is highly hydrophobic, without UV illumination,
reconstruction process. because it is clean; (2) the presence or absence of oxygen vacancies
The influence of tensile and compressive stress on the PIH effect has no discernable effect on the apparent interaction with water;
is surprising [511]. Tensile stress applied to a TiO2 film enhances and (3) the photocatalytic removal of an organic monolayer leads
the effect, while compressive stress suppresses the effect. This to a recovery of hydrophilic properties.
interesting phenomenon was explained in terms of the intrinsic A second major work that advocated the contamination layer
stress created when the surface undergoes the reconstruction mechanism was that of Zubkov et al. [517]. This paper shows
discussed above. Related to this is a study in which it was observed convincingly the effect of additions of hexane to the gas phase
that a TiO2 film became significantly rougher under illumination above the sample. For example, with 120 ppm hexane present in a 1
and then reverted to its smooth texture in the dark, as reported by atm O2 atmosphere, there is an induction time of about 150 s; up to
Watanabe and coworkers [521]. that time, the contact angle is relatively constant at 20◦ , and then
Another related study, carried out with in situ grazing angle it drops suddenly to the range in which it is difficult to measure
incidence X-ray diffraction, showed that illumination of a rutile the contact angle (0◦ –7◦ ), as shown in Fig. 5.5. However, it is also
single crystal in the presence of various alcohols led to a reversible significant that, in the absence of hexane, the clean crystal initially
expansion of the lattice [512]. One possible explanation is the shows a contact angle of ca. 25◦ (see curve for 0 ppm hexane). After
electrochemical intercalation of protons Eq. (3.8), leading to the the initiation of UV illumination, it requires only ca. 12 s to reach
type of lattice expansion shown in Fig. 3.15. A more recent study the 0◦ –7◦ range. This result is further substantiated by results that
carried out with the X-ray truncation rod technique in aqueous are shown for the effect of the presence vs. absence of O2 (Fig. 5.6).
solution without an organic compound present to trap holes, led There is no doubt that contamination films cause the surface to
to a disordering of the single crystal surface [513]. This result was be hydrophobic; this issue is important and must be kept in mind
considered to be consistent with the type of domain formation in the proper design of experimental measurements in this area.
reported by Wang et al. in our early papers on the PIH effect [496, In very recent work, it has become increasingly evident that there
514]. are actually two distinct effects, the first being the destruction of
More recent work has focused on new ways to enhance the PIH organic contaminants, which lowers the contact angle from higher
effect. For example, Gu et al. found that nanotexturing of the TiO2 ranges (e.g., 60◦ ) to the 20◦ –35◦ range and a separate effect that
surface can lead to an extremely long-lived hydrophilic character, lowers the value down to the 0◦ –5◦ range [518]. Consistent with
lasting more than one year, in comparison to the conventional film, this is the result already cited by Zubkov et al., as well as the recent
which becomes completely hydrophobic within 50 days [515]. The work of Hennessy et al. [513].
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 563

Fig. 5.4. Nanostructured TiO2 surface (a) microscopic image and (b) WCA vs. time for the conventional TiO2 (upper curve) and nanotextured (lower curve) surfaces.
© 2004, American Institute of Physics.

pronounced and is clearly not the dominant one in the PIH effect.
Thus, it is clear that the mechanism needs to be modified.

5.2.3. Oxidative mechanism


The report that positive polarization has a strong effect on the
attainment of a hydrophilic surface demonstrates the importance
of photogenerated holes in any broadly useful mechanism for
the PIH effect [504]. The mechanism that was proposed in that
paper again invoked the production of oxygen vacancies, and this
indeed appears to be a weakness, for the same reasons already
discussed. In this case, we cannot argue that the direct product
of the reaction of a vacancy with water could be the same as
that obtained electrochemically, but there is almost certainly a
structure that is produced photochemically that could be similar
to the one obtained electrochemically. This point also needs to be
examined in further research. The recent work of Nakato and co-
workers on the photoproduction of peroxo species on the titania
Fig. 5.5. Traces of WCA vs. time under various atmospheres: 1 atm O2 containing
surface is important in this regard [219,220,519,520].
added hexane (taken from Zubkov et al. [517]).
© 2005, American Chemical Society.
5.2.4. Combined redox mechanism
5.2.2. Reductive mechanism The mechanism already given for the general photocatalytic
In our papers prior to 2001, we focused on what could be called processes (see Section 4.1.1 and Fig. 4.5) is also proposed for the PIH
a reductive mechanism, in which UV light was thought to produce effect, because it combines both reductive and oxidative elements,
oxygen vacancies, at which water then dissociated, creating two both of which are important. Specifically, Fig. 4.5a, which involves
hydroxyl groups. There are several papers in which it is reported water as an electron donor, without an organic present, may be
that UV light can create oxygen defects, but, as already mentioned, useful. Panel A includes water as an electron acceptor, while panel
Mezhenny et al. showed the lack of an effect of reasonable light B includes O2 as an electron acceptor. Either may be appropriate
intensities on the appearance of rutile (110) surfaces with STM under specific conditions. From the result of Zubkov et al. (Fig. 5.6),
[139]. Thus, it seems clear that this mechanism would have to be however, it can be seen that O2 is necessary to accelerate the
modified. One modification that is sensible is simply to recognize PIH process. It is important to keep in mind, however, that O2
that an electron with the reducing power of hydrogen, or even accelerates the recovery of the original hydrophobic behavior in
slightly less, can directly reduce a bridging oxygen on rutile (110) the dark.
to produce a single OH group (reaction (3.2)). This product is As mentioned earlier, this type of mechanism might be
very similar, if not identical to that produced via reaction of consistent with a report in which the effect of tensile stress
a thermally produced oxygen vacancy with water. The degree was discussed [511]. The presence of localized areas in which
of sameness or difference should in fact be carefully examined, hydrogen is inserted electrochemically is consistent with the idea
but, to a first approximation, it should be identical. Our report of volume expansion, as discussed theoretically by Koudriachova
on results obtained with cathodic polarization enhancement of et al. (see Fig. 3.15a, b) [160]; volume expansion can certainly
hydrophilicity [505] seems to support this, but further work needs lead to compressive stress, which would be alleviated by tensile
to be carried out. stress. Volume expansion was also indicated in the work of
As White et al. point out, even in the case in which there are Matsushige and co-workers [512], in which in situ X-ray diffraction
indeed oxygen vacancies present on the surface, they do not have measurements were made on a rutile single crystal under UV
a measurable effect on the hydrophilic properties [516]. Intuitively, illumination in the presence of organic hole trapping agents. This
we expect that the presence of bridging hydroxyls should lead idea is also consistent with the report of Watanabe and co-workers
to an enhanced interaction with bulk water, and this supposition on increased surface roughness during illumination [521]. More
appears to be valid from the experiments just cited with cathodic recently, an in situ X-ray truncation rod study showed an increase
polarization [505]. However, the effect with anatase film is not in apparent roughness during UV illumination in the absence of a
564 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Table 5.1
Possible unitary reactions that might be involved in the PIH effect.

Fig. 5.6. Traces of WCA vs. time under inert atmosphere (no O2 ) and under 10% O2
(taken from Zubkov et al. [517]).
© 2005, American Chemical Society.

hole trapping agent [513]. This was suggested to be consistent with


the original reports of Wang et al., in which surface nanostructures
were observed on a rutile single crystal during UV illumination
(Fig. 5.1). It is likely that the presence of hole trapping agent in the
earlier X-ray study [512] allowed the regions of proton injection to
increase in size and thus to be observed easily.
The development of a general mechanism for photocatalysis
and a specific one for the PIH effect is an ongoing endeavor. At
this point, we would like to recommend that a focused effort be
undertaken to understand the possible reactions occurring on the
TiO2 rutile (110) surface, both experimentally and theoretically,
that might lead to PIH effects. As already mentioned, there are
several candidates for surface entities that might enhance the
hydrophilic properties, including bridging OH groups, as in the
reductive mechanism [496,505,514], and surface-trapped holes
of various types, as in the oxidative mechanism [504]. On the
oxidative side, the intermediates involved with the O2 evolution
reaction could also be important [111,218–220]. The attack of a
surface-trapped hole on adsorbed water can in principle produce
an adsorbed •OH radical in an electrochemical reaction, or can
lead to a bond breakage, which might be more likely in the case
of a bridging oxygen. We have attempted to assemble as many as
possible of the reasonable unitary reactions, i.e., a single reaction
event, including electron transfer, proton transfer, bond breaking
and bond making (Fig. 5.7, Table 5.1.). Such unitary reactions might
also be useful in building more complicated mechanisms. We
present these as a possible starting point for further guidelines in
stimulating both experimental and theoretical studies that would
try to finally be directed at achieving a deep understanding of the
most studied oxide surface, rutile (110). If good agreement can
be achieved between experiment and theory on such a surface, it
should also be a great impetus toward the achievement of similar
understanding of other important rutile and anatase surfaces.

5.2.5. Visible-light-induced PIH effect


Several reports have appeared concerning the development of
visible-light-induced hydrophilicity, encouraged by the boom in
studies on visible light photocatalysis [454,473,474,522]. These
studies focusing on N, or C-doped TiO2 films and showed quite The locations of each of the moieties are indicated by the letters in the titles for each
scattered results on the rate of hydrophilic conversion due to reaction, referring to Fig. 5.7. The listed reactions are a sampling of the possible ones
for the TiO2 rutile (110) surface.
differences in the preparation methods of the films. Some studies
stated that the hydrophilic conversion rates were slow and the
superhydrophilic state were not obtainable even in reasonably studies stated that the superhydrophilic state could be obtained for
long period (over 100 h) of visible irradiation [473,522], but some N-, or C-doped TiO2 films within hours of visible irradiation [454,
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 565

Fig. 6.1. Schematic diagram of the decontamination process occurring on the


superhydrophilic self-cleaning surface [15]. ©2006, Elsevier Science Ltd.

flow to the superhydrophilic property of TiO2 surface, i.e., water


penetrated the molecular-level space between the stain and
the superhydrophilic TiO2 surface [14,15]. In other words, this
phenomenon has effectively removed the limitation of the self-
cleaning function of TiO2 photocatalysis set by the number of
incident photons. Even though the number of photons may
be insufficient to decompose the adsorbed stain, the surface
Fig. 5.7. Schematic top view of the rutile (110) surface, showing the structures that
are used in Table 5.1 regarding the possible reactions involved in the PIH effect. The
is maintained clean when water is supplied there. Thus, they
red rectangle corresponds to the designation ‘a’ in Table 5.1, the blue rectangle to suggested that the best use of self-cleaning TiO2 surfaces should
‘b’, and the green rectangle to ‘c’. (For interpretation of the references to colour in be exterior construction materials, since these materials could be
this figure legend, the reader is referred to the web version of this article.) exposed to abundant sunlight and natural rainfall [9,14,15]. Such
materials, including tiles, glass, aluminium siding, plastic films,
474]. However, all of the studies have reached the same conclusion, tent materials, cement, etc., have already been commercialized in
Japan since the late 1990s and in other countries in recent years
i.e., the rate of hydrophilic conversion is much slower under
[15,524]. Some examples are shown in Fig. 6.2.
visible irradiation than UV irradiation even when the absorbed
As estimated by TOTO, Ltd., the pioneer of self-cleaning
photons are in same number for both cases, the same as what
technology, a building in a Japanese city covered with ordinary
was observed in visible light photocatalysis. Also, it appears that
tiles should be cleaned at least every five years to maintain
films that are more photocatalytically active undergo, hydrophilic
a good appearance, while that covered with self-cleaning tiles
conversion more easily and rapidly under the visible irradiation
should remain clean over a span of twenty years without any
[454,473,474,522].
maintenance. Therefore, self-cleaning technology can lead to large
decreases in maintenance costs. We have been examining the
6. Brief review of applications efficacy of self-cleaning technology at the Photocatalyst Museum
of our Institute. Fig. 6.3 shows the results of one test on a self-
6.1. Self-cleaning surfaces cleaning tent material produced by Taiyo Kogyo Co. The tent
material, made from PVC film, was coated with TiO2 photocatalyst
The TiO2 surface can decompose organic contamination with material on the left half. The test was started on July 22nd, 2004;
the aid of ultraviolet light. This observation suggests the applica- on that day, both the non-coated and TiO2 -coated halves had
tion of TiO2 photocatalysis to a novel ‘‘self-cleaning’’ technique, the same white appearance, as shown in Fig. 6.3a. The photo
i.e., a surface coated with TiO2 can maintain itself clean under shown in Fig. 6.3b was taken on April 23rd, 2007. The TiO2 -
ultraviolet illumination (Fig. 6.1) [5,9,10,392]. This technique is coated part remains white in the photo, whereas the non-coated
obviously of great value, since it can utilize freely available so- part, however, is dark grey. Note that the tent was placed in a
lar light or waste ultraviolet emission from fluorescent lamps, location surrounded by tall buildings, where little direct sunlight
save maintenance costs, and reduce the use of detergents. Watan- illumination could reach the surface. Nevertheless, even under
abe, Hashimoto and Fujishima first demonstrated this self-cleaning such conditions, the self-cleaning effect is obvious. On clear days,
concept on a titania-coated ceramic tile in 1992 [392]. A simi- contaminants accumulate on the whole surface, but during a rain,
lar idea was conceived independently by Heller [5]. One of the these are washed off the TiO2 -coated part of the surface.
first commercialized products using this technique was the self- In Japan, several thousand tall buildings have been covered with
cleaning cover glass for highway tunnel lamps [9,14,523]. This type TOTO’s self-cleaning tiles. The Central Japan International Airport,
of lamp, which is often a sodium lamp in Japan, emits UV light of which opened in February, 2005 near Nagoya, Japan, used over
about 3 mW cm−2 at the position of the cover glass. This UV light 20 000 m2 of self-cleaning glass manufactured by the Nippon Sheet
is of no use for lighting, but it is sufficient to decompose the con- Glass Co. Self-cleaning tent materials have been widely applied for
tamination from exhaust compounds. As a result, the cover glass storage structures, business facilities, bus and train stations, sports
can maintain transparency for long-term use. centers, sunshades in parks and at the seaside. The PanaHome
The efficacy of self-cleaning surfaces was found to be dependent Company, one of the major house manufacturers in Japan, has
on the relative rates of decontamination vs. contamination. marketed ‘‘eco-life’’-type houses since 2003; self-cleaning tiles and
The TiO2 photocatalyst can maintain the surface clean only windows, in addition to a solar cell-covered roof-top, are utilized.
when the photocatalytic decontamination rate is greater than New applications of self-cleaning technology are also under
that of contamination. Wang et al., however, observed that the examination. Zhang et al. recently reported on antireflective
self-cleaning effect of TiO2 surfaces could be enhanced when TiO2 –SiO2 self-cleaning coatings that may find application for
water flow, such as natural rainfall, was applied to the surface the cover glass of solar cells [525,526]. Conventional TiO2 -based
[514]. They attributed this enhancing phenomenon of water coatings are highly reflective due to the large refractive index of the
566 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 6.2. Applications of self-cleaning exterior building materials. (a) Picture of the MM Towers, in Yokohama, coated with self-cleaning tiles (courtesy of TOTO). (b) Picture
of the Matsushita Denso building covered with self-cleaning glass (courtesy of Nippon Sheet Glass). (c) Picture of the self-cleaning sound-proof wall (courtesy of Sekisui).
(d) Eco-life-type houses using self-cleaning tiles and glass (courtesy of PanaHome). (e) Self-cleaning roof of a train station in Motosumiyoshi (courtesy of Taiyo Kogyo).

Fig. 6.3. Outdoor exposure test for a PVC tent material manufactured by Taiyo Kogyo done in the Photocatalyst Museum, KAST. The left half part of the tent material was
coated with TiO2 . (a) Picture taken in July 22, 2004. (b) Picture taken in April 23, 2007.

material, and thus the loss of transmittance by TiO2 can be larger the coating enhanced the transmission of the glass to exceed 97%
than that caused by contamination layers. Zhang et al. prepared for visible light, greater than the 90–92% usual for glass (Fig. 6.4)
nanoporous TiO2 –SiO2 coatings that have a low refractive index; [525,526].
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 567

[537–540]. These chemicals, including natural hormones, dioxins,


biphenol-A, etc., in the aquatic environment have been implicated
as health hazards for both humans and wildlife. They are assumed
to disrupt normal endocrine functions through interaction with
steroid hormone receptors, even at very low concentrations.
Conventional biological methods to remove these EDCs require
long periods of time, and chemical oxidation methods are in
general not economical because of the low concentrations of the
EDCs. Nakashima et al. designed a photocatalytic reactor using
TiO2 -modified-PTFE mesh sheets as photocatalysts and utilized
this to treat the water discharged from the Kitano sewage-
treatment plant, which is located next to the Tama River near
Tokyo [540]. Concentrations of estron in the discharged water were
140 ng L−1 . Under UV illumination (1.2 mW cm−2 ), about 90%
of the initial estron was decomposed in a very short time, with
good reproducibility (Fig. 6.5). Thus, TiO2 photocatalysis can be
Fig. 6.4. Transmission spectra of a bare glass and a glass coated with TiO2 –SiO2
composite film. The maximum transmittance for the coated glass could be over 99% applied in the treatment of sewage effluent as a safe method for
[525]. removing natural and synthetic estrogens. This approach should
© 2005, American Chemical Society. also be suitable for the removal of other low-level EDCs in the
aquatic environment.
TiO2 -coated textiles are another promising application of self- Hashimoto and co-workers recently reported the application
cleaning technology. Recent studies by Kiwi and co-workers have of photocatalytic technology to the hydroponic cultivation of
shown that coffee and wine stains on TiO2 -treated cotton or tomatoes [14]. In conventional hydroponics, tomato plants are
synthetic textiles can be decolorized under sunlight [527,528]. planted in an organic culture medium. The culture solution,
containing nitric acid, phosphoric acid, potash, etc., flows through
6.2. Water purification the culture medium in a circulated mode and provides nutrients
to the plants. However, organic substances are released into
As noted in our historical overview, the earliest reports of the culture medium during circulation, and as a result, plant
the cleaning properties of TiO2 were those of Frank and Bard in pathogens can be propagated in the system and can cause diseases
1977, in which they found that suspensions of TiO2 powder in in the plants. The researchers connected a photocatalytic-water-
contaminated water were able to photocatalyze the conversion of treatment tank to the circulation system and used sunlight to
cyanide to cyanate, thus detoxifying the water [56,57]. decompose organic substances in solution. The method is effective,
One of the advantages of TiO2 photocatalysis for water as evidenced by the apparent decrease in the total organic carbon
decontamination is that only the TiO2 photocatalyst (immobilized in solution and the increased yields of tomatoes.
or suspended) and UV light, either from solar light or artificial Other important water-purification applications of TiO2 pho-
light sources, are needed and thus its cost can be lower than other tocatalysis include water disinfection [406,408,541–545], re-
kinds of advanced oxidation techniques (UV/O3 , UV/H2 O2 , photo- mediation of metal contamination [546,547], and oxidation of
Fenton). Moreover, no toxic intermediate products are generated arsenite [548]. For developing countries, solar photocatalytic dis-
in the photocatalytic decontamination process; this makes it very infection appears to be a promising process to produce drinking
attractive for cleaning the water environment, even for cleaning water, which could help in improving public health. Rincón et al.
drinking water [379,380,382,529]. reported that real lake water contaminated with E. coli K12 at a
However, it is generally accepted that TiO2 photocatalysis level of 106 CFU mL−1 could be disinfected completely within 3 h
is feasible only for the treatment of wastewater that contains on a clear day in summer, using a CPC reactor and a TiO2 slurry
contaminants at low to medium pollutant concentrations, because catalyst [549]. In comparison, solar light alone is unable to effec-
of its relatively low efficiency and the limited flux of ultraviolet tively disinfect the contaminated water. The mechanism for photo-
photons [382,530]. Malato and co-workers studied the solar
catalytic disinfection is still under debate (see Section 4.2.4) [408].
decontamination of industrial [390], agricultural [531,532] (3,4),
Various active oxygen species, such as hydroxyl radicals, superox-
and municipal wastewater [533] at the pilot-plant scale, using
ide, hydrogen peroxide, singlet oxygen, etc., may participate in the
compound parabolic collector (CPC) reactors and TiO2 slurry
disinfection process. In addition, the self-repair action of microbes
photocatalysts. They observed the complete mineralization of
complicates the disinfection process [408].
organic matter at concentrations of ca. 50 mg L−1 within hours
There have been several types of water-purification facilities on
under sunlight. However, the decontamination rate was generally
the market for various purposes. One of these is designed to re-
lower for TiO2 photocatalysis than for the photo-Fenton process
move volatile organic compounds (VOC), such as trichloroethylene,
in their studies [534,535]. Herrmann and co-workers reported
a solar photo-reactor design based on the multi-step cascade tetrachloroethylene, 1,3-dichlorobenzene, and dichloromethane,
falling-film principle to ensure good exposure to sunlight and good etc., from underground water or soil. These substances are widely
oxygenation of the effluent to be treated [536]. These workers used used as extracting solvents in industrial processes and in dry clean-
an immobilized TiO2 photocatalyst that was prepared by coating ing and are released into the environment in large quantities. The
non-woven paper with anatase paste using aqueous SiO2 colloid facility basically involves suctioning up underground water from
as a binder material in order to avoid the filtration procedure a deep well dug in the polluted area, spraying the water to evap-
required in slurry photo-reactor. This film photo-reactor was orate the VOCs into the gas phase, and decomposing the gaseous
found to be as efficient as the CPC slurry photo-reactor for 4- VOCs with a photocatalytic air-cleaning unit. VOCs at concentra-
chlorophenol, formetanate, and a mixture of pesticides in terms of tion levels ranging from several hundreds to several thousands of
total degradation under identical solar exposure. parts per million can be decomposed almost completely to carbon
One interesting application of TiO2 photocatalysis is to remove dioxide and hydrochloric acid with such a facility, with a maximum
endocrine disruptor chemicals (EDC) in the aqueous environment treatment capability of 5 kg VOCs per day for a 15-kW facility.
568 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 6.5. (a) Concentration change of estron (E1) in water discharged from a sewage-treatment plant during photocatalysis. The measurement was repeated three times
(UV light intensity, 1.2 mW cm−2 ; water temperature, 15 ◦ C) [540]. ©2003, Elsevier Science Ltd. (b) Picture of a photocatalytic reactor for treatment of sewage water. The
reactor consists of two 34-L columns, 100 pieces of TiO2 -coated PTFE mesh sheet, and 12 black-light (40 W) lamps which gave a light intensity of about 1.2 mW cm−2 .

Fig. 6.6. (a) Photographic images of a honeycomb-type air-cleaning filter (left) and a three-dimensional porous ceramic air-cleaning filter (right); (b) Schematic diagram of
the microscopic structure of the air-cleaning filter; the substrate materials (ceramics, paper, fabrics) are coated with a composite of TiO2 and adsorbent.

6.3. Air purification type construction or a three-dimensional porous structure for


minimum pressure drop. TiO2 nanoparticles are coated on the body
One of the most important applications of TiO2 photocatal- of the filter with active carbon, zeolites, etc., as co-adsorbents.
ysis is to decontaminate, deodorize, and disinfect indoor air. The co-adsorbents facilitate the adsorption of VOCs on the filter,
Low-concentration volatile organic compounds (VOCs), such as and ensure that no intermediate gaseous compounds diffuse into
formaldehyde, and toluene, emitted from interior furnishings and the ambient atmosphere. The adsorbed substances diffuse on the
construction materials, may lead to the ‘‘sick building syndrome’’ surface of the adsorbent until they reach the TiO2 and then undergo
and other diseases. Besides, the indoor air of public facilities, which photodecomposition. The oxidizing species on the photocatalyst
is normally contaminated with bacteria, fungi, etc., threatens the can also diffuse to the surface of the adsorbent and participate
health of users. Conventional air-purification systems often adapt
in oxidation reactions there. UV light, illuminated on the filter
filter-type components for the cleaning of polluted air. Pollutants
surface from black-light lamps, is generally at the level of several
are accumulated in the active carbon filters, and the filter becomes
mW cm−2 ; this ensures the rapid mineralization of VOCs that
saturated with adsorbed substances and loses its function after a
typically exist in concentrations of about 1 ppm or even lower.
certain period of use. Treatment of the used air filters may cause
the risk of secondary pollution. Photocatalytic air-cleaning filters, After long-term use, the filter may be poisoned by HNO3 or H2 SO4
however, can decompose the organic substances instead of accu- formed during removal of ammonia or hydrogen sulfide. In such
mulating them and, as a result, exhibit better performance than a case, the poisoned filter can be regenerated by simply washing
conventional ones [9,15,373,550–552]. Moreover, the photocat- with water.
alytic filter can kill the bacteria floating in indoor air, which is also The available air-cleaning devices based on photocatalytic
important for indoor air purification [553,554]. filters include air cleaners, air conditioners, air-cleaning units
Fig. 6.6 shows images of two types of photocatalytic filters used for refrigerators, etc. Among these application groups, the air
in air-purification devices. The filters feature either a honeycomb- cleaner is the largest. There are more than thirty companies
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 569

Fig. 6.7. Usage of TiO2 -based photocatalytic material on roadway surfaces to convert nitrogen oxides (NOx ) to nitrate: (left) application of the coating; and (right) finished
roadway, with the coated surface showing a lighter color (courtesy of Fujita Road Construction Co., Ltd.).

manufacturing photocatalyst-type air cleaners in Japan, with over 6.4. Self-sterilizing surfaces
one hundred product types available on the market. The consumers
of photocatalyst-type air cleaners include hospitals, institutions for TiO2 photocatalysts can kill bacteria on their surfaces, and
the elderly, pet stores, offices, smoking rooms, sports clubs, hotels,
therefore self-sterilizing surfaces can be prepared [9,10,530]. Such
schools, restaurants, food manufacturers, or in any location where
materials are particularly useful for hospitals and elder care
indoor air purification is necessary. In addition to air cleaners,
facilities, where the control of surface and airborne bacteria and
deodorizing sheets that are composed of TiO2 photocatalyst and
viruses is important. In cooperation with TOTO Ltd., Fujishima
active carbon adsorbents are also available on the market. The
et al. developed a type of antibacterial tile made by covering
sheet is able to remove malodorous substances in kitchens,
refrigerators, vehicles, and shoes through the adsorption function ordinary tiles with a TiO2 –Cu composite coating [9,530]. In a test
of the activated carbon. Placing the used sheet in sunlight readily experiment, the bacteria on these tiles were killed completely in
decomposes the adsorbed pollutants by the action of the TiO2 one hour under illumination equal to the brightness of the surface
photocatalyst and regenerates the sheet. of a study desk (UV light: ∼4 µW cm−2 ). On the basis of these
Nitrogen oxides (NOx ) exhausted from automobiles have be- results, such tiles were tested on the floor and walls of a hospital
come a serious source of air pollution in urban areas. Photocatalytic operating room where sterile conditions are crucial [9,530]. After
TiO2 has the capability of removing nitrogen oxides [555,556]. It installing the tiles, the bacterial counts on the walls decreased to
has been suggested that the nitrogen oxide (NO) in the air is oxi- negligible levels in a period of one hour. Surprisingly, the bacterial
dized when the catalyst is exposed to light. Then, through the in- counts in the surrounding air also decreased significantly. TOTO
termediate stage of nitrogen dioxide (NO2 ), it is finally converted Ltd. has marketed the antibacterial tiles in Japan. Its users include
to nitric acid, which must then be stored. At the nitrogen dioxide hospitals, hotels, and restaurants, among others.
stage, part of the gas may escape from the photocatalyst surface, One advantage of photocatalytic sterilizing surfaces is that
but, with an adsorbent like activated carbon mixed with the cata- they operate in a passive fashion, that is, without the need for
lyst, this gas may be effectively trapped [9]. electrical power or chemical reagents, only light and oxygen being
The range of concentrations of air pollutants that can be required. Unlike chemical antibacterial agents, TiO2 surfaces are
efficiently removed is from 0.01 ppm to 10 ppm. These concentra- non-poisonous and will not result in environmental pollution.
tions correspond to the range from those in the ordinary environ- It is thus expected in the near future that self-sterilizing TiO2
mental atmosphere to those in highway tunnels. Some companies
materials may have many medical applications. At least, hospital
in Japan are now considering covering roads with TiO2 photocata-
room walls, medical instruments, and hospital uniforms can all
lyst, and removing the NOx exhausted from automobiles with sun-
be coated with self-sterilizing TiO2 materials. In a recent study,
light. TiO2 -coated road bricks, which are used to cover pavement,
Ohko et al. coated flexible silicone catheters with a TiO2 thin
have become available on the market. Interestingly, TiO2 coatings
film [558]. The film, as shown in Fig. 6.8, strongly adhered to
for roadways are being prepared by mixing colloidal TiO2 solutions
with cement. Nitric acid, formed during the oxidation of nitrogen the body of the catheter: it was able to sustain repeated bending
oxides, reacts with the cement to form calcium nitrate, which can without cracking and peeling observed macroscopically; it was
be washed off by rainwater. This method, which has been termed able to pass an adhesive tape test, and it was resistant to scratching
‘‘photo-road’’ technology, has been tested in more than fourteen with a 3H pencil. The TiO2 -coated silicone catheter exhibited good
places in Japan over the past nearly ten years (see Fig. 6.7). One self-sterilizing properties under 1 mW cm−2 UV illumination, as
of the places tested is on the 7th belt highway in Tokyo, with shown in Fig. 6.9, and was shown to be safe in an experiment
an area of about 300 m2 . Nitrogen oxide removed from this test- using cultured cells and animal experiments [559]. Clinical studies
ing area was estimated to be 50–60 mg per day, corresponding showed that the antibacterial capability of TiO2 -coated catheters
to the amount of nitrogen oxide discharged by 1000 automobiles was superior to that of conventional catheters, and suggested
passing by. the promising clinical use of the self-sterilizing catheter as an
More recently, this idea has been developed to a great extent intermittent (self-)catheterization catheter for neurogenic bladder
in Europe. In particular, the Italian cement company Italcementi patients [559]. The self-sterilizing catheter can be disinfected in
has created several demonstration projects, for example, in Rome a specially designed portable sterilizer box, or by direct sunlight.
and Paris, to check the actual photocatalytic effect of a TiO2 -coated The intra-corporeal use of TiO2 -coated catheters is also under
highway in the conversion of NOx and SOx to more environmentally examination. In this case, metallic Ag is deposited on the TiO2
benign forms, i.e., NO− 2−
3 and SO4 [557]. surface to enhance the antibacterial action under dark conditions.
570 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 6.8. Scanning electron micrographs of (a) the surface and (b) the cross-section of the TiO2 -coated silicone catheter. There is a thin intermediate SiO2 layer between TiO2
and catheter substrate [558].
© 2000, John Wiley & Sons, Inc.

application of this phenomenon has been for automobile side-view


mirrors. The TOTO Ltd., in Japan currently is supplying TiO2 -treated
glass for such mirrors to major automobile manufacturers [9,14].
In addition, it has marketed adhesive TiO2 -coated plastic films or
TiO2 -containing sprays for the conversion of existing mirrors.
The uses of anti-fogging technology, however, are not limited to
mirrors and glass. One example in our daily life is air conditioners.
When an air conditioner is operating to cool during the summer,
air that is cooled on the heat exchanger of the air conditioner
often has moisture condensing out of it, except in an extremely
dry climate. For an ordinary plate heat exchanger, the condensed
Fig. 6.9. (a) Survival rate of E. coli inside several kinds of catheters under water droplets fog the fins and large water droplets often form,
ultraviolet-A irradiation. The catheters were filled with a 0.1 mL aliquot of E. coli
often bridging between the adjacent fins; this can reduce the heat
suspension, as follows: N, control 1, original silicone catheters, no UVA irradiation;
, control 2, TiO2 -coated catheters, no UVA irradiation; , original silicone catheters, exchange efficiency and increase the resistance to air flow. Coating
with UVA irradiation (1 mW cm−2 ); ×, TiO2 -coated catheters, with UVA irradiation the heat exchanger surface with superhydrophilic TiO2 , however,
(1 mW cm−2 ) [559]. can hinder the fogging and the formation of water bridging, and
© 2007, The Japanese Urological Association. remove the condensed water more effectively. This has been
proven by the recent studies of Takata et al. [561]. They observed
6.5. Anti-fogging surfaces that the volume of water flowing off superhydrophilic TiO2 -coated
fins was 20% greater than that from ordinary fins, when humid
The fogging of the surfaces of mirrors and glass occurs when air (RH 90%, 40 ◦ C) was passed through the cooled fins (surface
moist air cools down on these surfaces, forming many water temperature: 15 ◦ C).
droplets. These droplets, ranging from micrometer to millimeter As the properties of superhydrophilic films continue to be
size, tend either to scatter light or simply to reflect or refract it improved, there will be many other possible applications of
randomly. In both cases, visual clarity is impaired drastically. In the anti-fogging phenomenon. Research activities have been
a serendipitous discovery, Watanabe and co-workers found that focusing on: (1) increasing the rate of superhydrophilic conversion
a TiO2 –SiO2 surface could become extremely hydrophilic under [427,515]; (2) prolonging the superhydrophilic states of the
UV illumination [500,560]. The result of this property is that films [515,526,562]; and (3) activating the films at longer
water spreads evenly across the surface. If the amount of water is wavelengths [473,522]. A rough TiO2 surface morphology has
relatively small, the water layer becomes very thin and evaporates the advantages of fast superhydrophilic conversion and long
quickly. If the amount of water is larger, it forms a sheet-like layer persistence of the superhydrophilic state, but it can do harm to
that also has high visual clarity (Fig. 6.10). The first commercial film transparency. A mixed film of TiO2 and SiO2 (or Al2 O3 ) has

Fig. 6.10. Photographs of fogged surfaces of usual glass (a) and TiO2 -coated glass after sufficient UV illumination (b), typically > 150 min at an intensity of 1 mW cm−2
[500]. ©1999, Elsevier Science Ltd.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 571

the advantages of long persistence of the superhydrophilic state,


but an excessive SiO2 –TiO2 ratio can have a negative effect on
the rate of superhydrophilic conversion. To develop a desirable
anti-fogging film, one obviously should strike a balance between
film composition and surface roughness. Extending the response
wavelength of TiO2 materials into the visible light region is
considered important for the design and actual applications of
superhydrophilic anti-fogging materials, but until now, only very
few reports have appeared on the superhydrophilic conversion
of visible-light-responsive TiO2 films (N-, or C-doped TiO2 ) under
visible illumination (see Section 5.2.5).

6.6. Heat transfer and heat dissipation

Superhydrophilic surfaces may improve the heat transfer


accompanied by liquid–vapor-phase transition. Takata et al.
studied the heat transfer properties of superhydrophilic TiO2
surfaces. They found that a TiO2 -coated Cu cylinder exhibited
excellent heat transfer characteristics in the nucleated boiling Fig. 6.11. Energy-saving system using solar light and stored rainwater [14].
regime, and its critical heat flux (CHF), a condition where liquid
cannot rewet the heater wall because of the rate of vapor 6.7. Anticorrosion applications
production impeding the liquid flow back to the hot surface,
was about two times higher than that of the uncoated surface In general, metals are protected from corrosion in several
[561,563]. They also compared the falling-film evaporation on ways, for example, by a corresponding passive metal compound
a superhydrophilic surface with that on an ordinary surface layer, by a paint layer, or by a sacrificial metal coating (with a
[564]. In this work, water was sprayed on the heated surface less positive corrosion potential). Of these, the passive layer and
through nozzles to form a water film. They found that on the the paint layer are effective only when the coating covers the
superhydrophilic surface, a stable water film was formed for metal surface perfectly. As for the sacrificial metal coating, its
much smaller water flow rates than those on an ordinary surface. lifetime is limited, because the metal coating dissolves gradually.
Since the heat transfer rate increases for thinner water films, the Recently, the application of photocatalytic technology in corrosion
prevention for metals has been examined [566–574].
superhydrophilic surface is expected to improve the performance
In cooperation with the Koyo Electrical Construction Co.,
of falling-film evaporators, which are widely used in seawater
Ohko et al. studied the anticorrosion effects of a TiO2 coating
desalination units and milk and juice concentrators. They observed
for Type 304 stainless steel [566]. Under UV irradiation, TiO2
that the superhydrophilic surface performed 40 times better
injected electrons into the steel and as a result protected it from
than the ordinary surface in low water flow and low heat flux. corrosion, while the photogenerated holes decomposed organic
For high water flow, a thick water film formed on both the contaminants to provide a self-cleaning function. Interestingly,
superhydrophilic and the ordinary surface, and thus no difference the coating remained effective even if perforated with pinholes.
in heat transfer rate was observed. These workers suggested that Later, the same group observed that TiO2 , when coupled with
the superhydrophilic surface could be an ideal heat transfer surface WO3 , maintained an anticorrosion effect even in the dark for
and would be applicable to various heat transfer phenomena that a period of time after UV irradiation ceased [567,570,571]. This
are affected by surface wettability [561,563,564]. phenomenon can be explained by the energy storage ability of
Hashimoto et al. reported recently the cooling of the exterior WO3 , as illustrated in Fig. 6.12. Under UV irradiation, electrons in
walls of buildings with a falling-water film, with the intent to the valence band of TiO2 are excited to the conduction band. The
help resolve the urban heat island problem for major cities and to excited electrons are injected into the metal so as to maintain its
reduce the electricity consumption by air conditioners in summer potential more negative than the corrosion potential. The excess
[14,565]. In their experiments, water was continuously sprinkled electrons can be accumulated in the ‘‘electron pool’’ of WO3 due
onto the surfaces of buildings that were covered with TiO2 films. to the lower conduction-band edge of WO3 compared to TiO2 . The
Due to the superhydrophilic property of the material, water was reduced WO3 then reacts with protons adsorbed on the surface
able to spread on the surfaces in the form of a thin water film or sodium cations in the electrolyte solution to form metastable
(∼0.1 mm). The evaporation of the water film can effectively cool tungsten bronze (Mx WO3 , M: H+ , Na+ ; x ≤ 1) so that the reductive
the building surface and also the surrounding air. They observed energy generated at the irradiated semiconductor can be stored
as in a battery. After the UV light is turned off, the metastable
on a clear day in the middle of summer that the temperature
tungsten bronze gradually releases electrons into the metal via a
drops were 15 ◦ C on window glass and 40–50 ◦ C on black-roof-tile
self-discharge process, and as a result the metal remains protected
surfaces. The cooling of the building surfaces is expected to result
from corrosion in the dark until the discharge process is completed.
in the reduction of electricity consumption by air conditioners.
Similarly, MoO3 also exhibited an electron pool function when
As estimated by Hashimoto et al., this reduction could be ten composited with TiO2 [572].
to several tens of percentage for a real house. Therefore, TiO2 - Research activities are continuing in order to extend the
coated building materials could contribute much to energy-saving anticorrosion period of TiO2 -based coatings in the dark [573,
technologies in the future (Fig. 6.11). The water film is so thin that 574]. Recent studies have shown that phosphotungstic acid (PWA)
only a small amount of water is required to cover the building exhibits a longer self-discharge time than WO3 [574]. The TiO2 -
surfaces. Moreover, natural rainwater can be collected and stored PWA coating can maintain an anticorrosion effect over 12 h in
in specially designed reservoirs for this purpose so that the cost of electrolyte solution, after being illuminated with 1 mW cm−2 UV
the water can be reduced. The water film also helps in maintaining light for 1 h. This is a very promising result for the development of
the building surface clean by the self-cleaning effect. practical semiconductor coatings for metal corrosion protection.
572 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 6.12. Schematic diagram of an energy storage-type photocatalytic anticorro-


sion system, which can protect metal from corrosion even in the dark [567].
© 2001, American Chemical Society.

6.8. Environmentally friendly surface treatment

In conventional technology for manufacturing printed circuit


boards (PCBs), insulating resin materials are first etched with
strong oxidants such as permanganate and hexavalent chromium
to prepare rough surfaces followed by electroless plating of a
copper layer on the whole surface and removal of the unwanted
copper after applying a temporary mask. The surface treatment
procedure is of vital importance for the strong adherence of the
copper layer to the resin substrate. However, the use of those
oxidants can cause serious environmental problems. Moreover,
the roughness of the etched surfaces is often of the order
of micrometers, which can lead to technical problems in the Fig. 6.13. SEM mapping of 10 µm/10 µm L/S copper wiring. (a)–(c) PCB substrates
fabrication of fine copper wires. Honma and co-workers recently were treated by TiO2 photocatalyst. (d)–(f) PCB substrates were etched with
reported a novel environmentally friendly surface treatment permanganate etching process. (c) and (f) are reflection mode images. As shown, the
photocatalytic treatment did not change the surface morphology of PCB much, and
technique for PCB manufacture by treatment of resin surfaces could be used to prepare finer copper wires than those obtained by the conventional
with TiO2 photocatalyst, replacing conventional etchants [575]. etching method [575].
They found that an aqueous TiO2 suspension was able to oxidize © 2006, Surface Finishing Society of Japan.
an epoxy resin surface within a depth of 30–50 nm under UV
illumination but did not affect surface roughness appreciably. composition/wettability contrast. This novel method would have
Palladium and tin catalysts for the electroless deposition of copper the advantages of simplicity and low cost: it does not require
penetrated into the surface-modified layer and electroless copper expensive photo-resist materials and special equipment.
plating was initiated at the bottom of this layer. As a result, the Tatsuma and co-workers first examined the concept of
copper layer showed excellent adhesion (1.15 kN m−1 ) to the photocatalytic lithography based on the remote oxidation effect of
epoxy surface, even better than that obtained by conventional TiO2 [360]. They placed a TiO2 -coated quartz substrate in contact
technology. By using this new surface treatment technique, Honma with a glass plate modified with an ultrathin organic layer, or
et al. succeeded in fabricating fine patterns of copper with lines and silicon, copper, or silver plate, separated by a small gap. The TiO2
spaces of 10 µm/10 µm, as shown in Fig. 6.13. was irradiated with UV light in air through a photomask. As a
Honma and co-workers also applied this new surface treatment result, two-dimensional images corresponding to the photomask
technique to modify the ABS resin surface, which is often etched were obtained; those images were based on the contrast of non-
with hexavalent chromium in conventional technology [576]. The oxidized to oxidized surfaces. In a following study, Kubo et al.
1, 2-polybutadiene distributed in the ABS matrix was preferentially examined the resolution of their novel lithography technique and
reformed by photocatalytic oxidation and induced hydrophilicity found that the resolution of remote oxidation-based photocatalytic
on the reformed surface. The ABS resin surface was reformed down lithography deteriorated slightly due to the lateral diffusion of
to the depth of a few µm and the catalyst (Pd and Sn) penetrated active oxygen species in the air [353] They obtained a resolution of
into the surface to the extent of 30–40 nm. Metal layers deposited 10 µm by optimizing the TiO2 film/photomask assembly (Fig. 6.14)
and suggested that the theoretical resolution should approach the
on the modified ABS surface showed adhesive strengths of about
scale of the light wavelength, namely the submicrometer level,
1.0 kN m−1 . They concluded that the superior adhesion strength
provided that a highly collimated light beam and a highly uniform,
was derived from a nano-anchor effect by the deposition of metal
non-light scattering TiO2 film were available [353].
inside the 30–40 nm reformed resin layer.
Lee and Sung combined photocatalytic lithography and selec-
tive atom layer deposition (ALD) to prepare patterned ZrO2 thin
6.9. Photocatalytic lithography films on Si substrates [577]. In their study, a patterned TiO2 thin
film (line width, 400 nm; line space, 580 nm) on quartz was used
The active oxygen species such as the hydroxyl radical or singlet as the lithographic tool. The TiO2 film was placed on an octadecyl-
O2 generated during the photocatalytic process are so active that siloxane self-assembled monolayer (SAM)-modified Si substrate,
they can oxidize not only organic materials but also inorganic and the assembly was irradiated with 254-nm UV light. The SAM
materials such as copper, silicon, carbon, and SiC etc. [353,359, was selectively removed, retaining the dimensions of the TiO2 pat-
360]. The oxidation can occur either on the TiO2 surface or at a tern without noticeable line spreading. By means of the ALD tech-
distance away from the surface. Therefore, it is possible to develop nique, a ZrO2 thin film was selectively deposited in the region with-
a novel photocatalytic lithographic method based on the above- out the SAM, with a 430-nm linewidth, through reaction with the
mentioned oxidation reactions to fabricate surface patterns with silanol groups of the Si substrate.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 573

Fig. 6.14. (a) Micrograph of the Cu plate surfaces patterned by the photocatalytic lithography. (b) Micrographs of masked TiO2 film (upper) and heptyldecafluorode-
cyltrimethoxysilane (HDFS)-coated glass plates patterned by photocatalytic lithography (below) [360].
© 2004, American Chemical Society.

Fig. 6.15. Micrographs of a drop of polystyrene microsphere suspension pipetted on a preheated superhydrophobic–superhydrophilic pattern, which were taken at (a) 5 s,
(b) 5 min, (c) 8 min, and (d) 9 min after pipetting the drop. Polystyrene microspheres were selectively deposited in the superhydrophilic stripes. The scale bar in pictures is
400 µm [585].
© 2007, American Chemical Society.

Photocatalytic lithography has also been applied to prepare and to site-selectively immobilize semiconductor quantum dots
chemical affinity/wettability patterns on surfaces [578–580]. [582]. There have been several studies concerning the pattern-
Spencer and co-workers irradiated an alkanethiol monolayer on an ing of superhydrophobic TiO2 films by photocatalytic lithogra-
Au substrate through a TiO2 film and a filter with a transmittance phy [583,584]. The resulting superhydrophobic–superhydrophilic
gradient [580]. Thus, these workers were able to prepare a surface patterns were shown to guide the evaporation of a la-
concentration gradient of alkanethiol on the Au surface, since tex suspension (Fig. 6.15) or to be filled site-selectively with
the oxidized alkanethiol molecule was removed from the surface. electroless-plated metal or polymer hydrogel, due to the extremely
Adsorption of other types of alkanethiol molecules led to the large wettability contrast in the pattern [585,586] In these cases,
formation of chemical affinity/wettability gradients on the Au the diffusion of holes in TiO2 and the surface diffusion of active oxy-
surface. gen species should restrict the lithographic resolution to the mi-
In addition to remote oxidation-based photocatalytic lithogra- crometer scale, as suggested by the studies of Haick and Paz [357,
phy, there have also been some efforts to prepare chemical affin- 358] A recent study by Kobayashi et al., however, showed that a
ity/wettability patterns directly on TiO2 surfaces. The patterned feature with 30-nm linewidth was able to be produced in a stearic
TiO2 surfaces were used to assemble colloidal crystal films [581] acid monolayer deposited on a TiO2 substrate by photocatalytic
574 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

Fig. 6.16. (a) Photochromism of the Ag–TiO2 film. Absorption spectra of Ag deposited on the TiO2 film by ultraviolet light irradiation (300–400 nm, 1 mW cm−2 , 15 min) and
after visible light irradiation (>400 nm, 50 mW cm−2 , 1 h). Corresponding photographs are also shown. (b) Multicolor photochromism of the Ag–TiO2 film. A–D, Differential
absorption spectra of the Ag–TiO2 film after visible light irradiation (10 mW cm−2 , 5 min) using a xenon lamp with bandpass filters [588].
© 2003, Nature Publishing Group.

lithography, aided by a local electrical field applied by a conduc-


tive AFM cantilever [587]. The positive bias (10 V) applied to TiO2
substrate lowered the local charge recombination, which led to the
rapid decomposition of the stearic acid molecules at the site to
which bias was applied.

6.10. Photochromism

Photochromic materials, which change their colors reversibly


in response to light, can be applied to smart windows, displays Fig. 6.17. A thick multicolored Ag–TiO2 film. Each spot (6 mm diameter) was
and memories. Conventional photochromic materials respond irradiated successively with a blue, green, red, or white light.
© 2004, American Chemical Society [591].
in a monochromatic way, so that multicolor photochromism
has required several different materials or filters combined
also ZnO films were able to work as the substrate material for Ag-
appropriately. If multicolor photochromism could be achieved
based multicolor photochromism [589]. SiO2 and ITO, however, did
with a simple material, photochromic devices would find a greater
not work. These workers suggested that both TiO2 or ZnO play
number of applications, including a rewritable color copy paper
an important role in the charge separation between the excited
or electronic paper and a high-density multi-wavelength optical
electrons and the Ag+ . Actually, a fraction of the excited electrons
memory. Ohko et al. recently reported the interesting multicolor on Ag were transferred to oxygen molecules via TiO2 or ZnO and
photochromism of TiO2 films loaded with silver nanoparticles non-excited Ag.
[588–591]. The film color, initially brownish-grey, changed under The color image is bleached gradually in air by ambient white
monochromatic visible light to almost the same color as that of light, which is detrimental to practical applications. Naoi et al.
the light (Figs. 6.16 and 6.17); the apparently uniform Ag–TiO2 showed that a stable color image can be obtained by either
film can take on almost any color. The color reverted to brownish- modifying the Ag nanoparticles with thiols [592], or washing the
grey under ultraviolet light, and these processes were found to be photogenerated Ag+ ions off the film [591]. Interestingly, the color
repeatable. image fixed by washing off the Ag+ ions could be maintained as a
Ohko et al. proposed the following model for the multicoloring ‘‘latent image’’, that is, the image can be recovered after a cycle of
process [588]: on irradiation with monochromatic visible light, white light bleaching and UV coloration due to the ‘‘molding effect’’
the corresponding Ag nanoparticles adsorb the light, and the of the nanoporous TiO2 film [591].
electrons thus excited are accepted by O2 , resulting in oxidation
of the Ag nanoparticles to Ag+ and an absorption decrease 6.11. Microchemical systems
appears at the corresponding wavelength, i.e., the color of the
excitation wavelength. The most important role of the TiO2 film Microchemical systems using microchannels as reaction fields
is the repeatable generation of Ag nanoparticles with diverse provide many technical advantages over conventional technolo-
sizes and shapes due to the ‘‘molding effect’’ of the nanoporous gies for chemical synthesis as well as analysis. Coating the in-
TiO2 film; this resulted in different plasmon resonance absorption ner walls of microchannels with TiO2 provides photocatalytic
wavelengths of the Ag particles in the whole visible light spectrum. reactivity to microchemical systems. Possible applications in-
Later studies by Kawahara et al. showed that not only TiO2 but clude the photocatalytic degradation of organic pollutants [590,
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 575

593], photocatalytic synthesis of organic compounds [594], site-


selective introduction of functional metal nanoparticles [595],
photo-controllable wettability of channels [596,597], and self-
cleaning microchannels, among others.
Takei et al. employed a TiO2 -modified microchannel chip for
the photocatalytic synthesis of L-pipecolinic acid from L-lysine
[594]. They observed that the conversion rate in the chip was 70
times larger than that in a cuvette with almost the same selectivity
and enantiomeric excess [594]. They suggested that the large
surface area of the microchannels, together with the short diffusion
distance, should improve the efficiency of photocatalytic reactions
within the channels.
Castellana et al. reported a simple method to introduce
noble metal nanoparticles inside sealed microfluidic channels
that were coated with a TiO2 photocatalyst [595]. Under UV
illumination, metal ions, for example, Ag+ , were reduced from
solution to the area covered with TiO2 to form an Ag nanoparticle
film. In cases in which photomasks were used, a patterned Ag
film was readily prepared. The Ag nanoparticle film could be
biofunctionalized to capture proteins with good discrimination
between addressed locations and background. Their method
appears to be quite general, and films of Au, Pd and Cu could
also be patterned. The ability to address individual ligands atop
nanoparticle films inside microfluidic devices can be combined
with such technologies as transmission surface plasmon resonance Fig. 6.18. (a) Irradiation time for preparation of the 4-step Laplace valve in one
channel and (b) fluorescence microscope images of fluorescein aqueous solution in
spectroscopy or surface-enhanced fluorescence. This allows for
the photopatterned and tuned channel [596].
the development of powerful lab-on-a-chip devices with label- © 2007, The Royal Society of Chemistry.
free detection or fluorescence detection with enhanced sensitivity.
Moreover, the metal nanoparticle films can be used to immobilize review or in a pure surface science-oriented review, for example,
patches of oriented enzymes inside microfluidic devices, which electrochemistry and photoelectrochemistry.
have promising applications in enzymatic microreactors. The early work in the field can be inspirational and still teach
Kim and co-workers applied photocatalytic reactions to control us much. For example, the work of Baur in the 1920s shows us the
the wettability of microchannels, by which pressure barriers, so- power of imagination in the development of science. His idea of the
called Laplace valves, were introduced to microchannels at desired pairing of oxidation and reduction reactions is still a central one
positions [596]. They modified the inner walls of the microchan- [23,24,598]. The work of Goodeve and Kitchener in 1938 could be
nels with TiO2 nanoparticles; afterwards they modified the TiO2 published today with only slight modification [26], but the fact that
nanoparticles with a hydrophobic self-assembled monolayer. The it was done at that time should inspire us to even greater efforts.
walls of the microchannel became superhydrophobic (water con- Fundamental studies in spectroscopy have led to a much better
tact angle > 150◦ ) after these procedures. When UV illumination understanding of the processes occurring in TiO2 in its various
was applied to the superhydrophobic microchannel, TiO2 nanopar- forms, including single crystals, powders and films. There is so
ticles decomposed the hydrophobic monolayer, which led to the much information available now that it is a challenge to make
gradual decrease of the water contact angle to 0◦ , the so-called sense of it all. One of the points that we have learned is that the
superhydrophilic state. Since the water contact angle of the TiO2 - nature of the sample itself, i.e., its preparation, handling, history,
modified microchannel wall was adjustable between 150◦ and 0◦ , and chemical and physical characteristics must be specified in
the pressure barrier was found to be tunable over a wide range; the as great detail as possible, because all of these are important in
latter was determined by the Laplace pressure P = −2γ cos θ /r, determining the observed behavior.
where γ is the surface tension of the liquid, r the channel radius, The dual use of experimental studies with single crystal stud-
and θ the contact angle. They demonstrated the preparation of ies and theoretical studies on the same surfaces is becoming
4-step wettability-based Laplace valves having different pressure an extremely powerful approach. It is hoped that this approach
barriers in one channel (Fig. 6.18), and also the construction of a will be extended and deepened, with increasingly sophisticated
batch operation system consisting of two sub-nL dispensers and a techniques, until a convergence can be reached even for photocat-
reaction chamber. alytically related reactions on the rutile (110) surface, which we
Thus, given the number of promising approaches already venture to say is still not completely understood. As a convergence
reported, it appears likely that photocatalysis will find increasing is achieved for this surface, we can become more confident that our
application in microchemical and microfluidic devices. experimental and theoretical techniques can be applied to other
surfaces. Of course, this is not to say that we should wait to study
7. Summary the other surfaces, but the point is that we cannot be truly confi-
dent until we can obtain convergence for this most studied system.
We have tried to provide an overview of the field of The work of Henderson has provided an excellent example [113,
photocatalysis from its very beginning in 1921 [17] through its 114,130,141–143,280,516,599–608].
developments in fundamental studies, both experimental and The recent STM work on reactions of water and oxygen with the
theoretical, which have been strongly tied to applications. Of rutile (110) surface is truly impressive and provides an example
course, it is impossible to do justice to this vast field in a review of what can be learned at the atomic scale [129,133,137]. The
of even this length, and new work is emerging every day. This realization that trace water, even in ultra-high vacuum, can react so
is the nature of the field. We have also tried to focus on some readily with oxygen vacancies to produce OH groups should cause
of the aspects of the field that would not be treated in a general us to examine carefully any work in which vacancies are discussed,
576 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

at least on the TiO2 surface. Similar approaches can no doubt lead [5] A. Heller, Acc. Chem. Res. 28 (1995) 503–508.
to further fundamental insight into the basic phenomena that are [6] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995)
69–96.
involved with photocatalysis. [7] A. Mills, S. LeHunte, J. Photochem. Photobiol. A: Chem. 108 (1997) 1–35.
As important as the surface is, we must not forget about the bulk [8] J. Peral, X. Domènech, D.F. Ollis, J. Chem. Tech. Biotech. 70 (1997) 117–140.
material, even in the form of a nanoparticle. This is even a more [9] A. Fujishima, K. Hashimoto, T. Watanabe, TiO2 Photocatalysis: Fundamentals
and Applications, BKC, Inc., Tokyo, 1999.
challenging aspect, since we do not have a real-space technique
[10] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1–21.
like STM that can probe beneath the surface at the atomic scale. [11] D.F. Ollis, C. R. Acad. Sci. Paris, Serie IIC, Chim. 3 (2000) 405–411.
It will be necessary to develop ingenious new techniques to try to [12] D.A. Tryk, A. Fujishima, K. Honda, Electrochim. Acta 45 (2000) 2363–2376.
do this. The roles of sub-surface oxygen vacancies, before and after [13] A. Fujishima, D.A. Tryk, in: A.J. Bard, M. Stratmann, S. Licht (Eds.),
Encyclopedia of Electrochemistry, Vol. 6: Semiconductor Electrodes and
reacting with water, as well as other bulk structures will need to Photoelectrochemistry, Weinheim, 2002.
be elucidated fully in order to understand in detail the trapping of [14] K. Hashimoto, H. Irie, A. Fujishima, Japan. J. Appl. Phys. 44 (2005) 8269–8285.
electrons and holes. [15] A. Fujishima, X. Zhang, C. R. Chimie 9 (2006).
[16] A. Fujishima, X. Zhang, D.A. Tryk, Internat. J. Hydrogen Energy 32 (2007)
After examining carefully the controversial aspects of the
2664–2672.
photo-induced hydrophilic effect, we have reached the conclusion [17] C. Renz, Helv. Chim. Acta 4 (1921) 961–968.
that there is an aspect of this effect that does not involve simply [18] A.E. Jacobsen, Ind. Eng. Chem. 41 (1949) 523–526.
the cleaning of the surface, as shown by recent work [513,518]. The [19] A. Fujishima, K. Honda, Nature 238 (1972) 37–38.
[20] T. Watanabe, A. Fujishima, K. Honda, in: T. Ohta (Ed.), Solar-Hydrogen Energy
precise nature of the effect has not been elucidated even now, but Systems, Oxford, 1979, pp. 137–169.
we propose that the surface species are basically the same ones [21] A. Fujishima, Innovative Hydrogen Production from Water (UNESCO), 2000.
involved with conventional photocatalysis. [22] M.A. Fox, M.T. Dulay, Chem. Rev. 93 (1993) 341–357.
[23] E. Baur, A. Perret, Helv. Chim. Acta 7 (1924) 910–915.
We have indicated some of the fascinating new applications of
[24] E. Baur, C. Neuweiler, Helv. Chim. Acta 10 (1927) 901–907.
photocatalysis. These are not to be considered so very separate [25] C. Renz, Helv. Chim. Acta 15 (1932) 1077–1084.
from the fundamental work. Both tend to reinforce each other. [26] C.F. Goodeve, J.A. Kitchener, Trans. Faraday Soc. 34 (1938) 570–579.
We believe that people working in both fundamental and applied [27] C.F. Goodeve, J.A. Kitchener, Trans. Faraday Soc. 34 (1938) 902–908.
[28] M.C. Markham, K.J. Laidler, J. Phys. Chem. 57 (1953) 363–369.
aspects should try very hard to understand what the others are [29] T.R. Rubin, J.G. Calvert, G.T. Rankin, W. MacNevin, J. Am. Chem. Soc. 75 (1953)
doing. This will undoubtedly lead to advances in both areas. 2850–2853.
[30] M.C. Markham, M.C. Hannan, S.W. Evans, J. Am. Chem. Soc. 76 (1954)
Appendix. TiO2 film preparation methods 820–823.
[31] M.C. Markham, M.C. Hannan, R.M. Paternostro, C.B. Rose, J. Am. Chem. Soc. 80
(1958) 5394–5397.
For reference, a list of photocatalytic film preparation methods [32] R.E. Stephens, B. Ke, D. Trivich, J. Phys. Chem. 59 (1955) 966–969.
has been assembled in Table A.1. [33] J.G. Calvert, K. Theurer, G.T. Rankin, W.M. MacNevin, J. Am. Chem. Soc. 76
(1954) 2575–2578.
[34] C.B. Vail, J.P. Holmquist, L. White, J. Am. Chem. Soc. 76 (1954) 624–625.
Table A.1
[35] M.G. Markham, J. Barry, M. Iava, J. Haddad, J. Phys. Chem. 61 (1957)
Preparation of TiO2 active surface layers. 1665–1668.
Method Author Year Ref. [36] J.C. Kuriacose, M.C. Markham, J. Phys. Chem. 65 (1961) 2232–2236.
[37] M.C. Markham, J.C. Kuriacose, J. DeMarco, C. Giaquinto, J. Phys. Chem. 66
Sol-gel Negishi et al. 1995 [76] (1962) 932–936.
Paz and Heller 1997 [395] [38] A. Nergararian, M.C. Markham, J. Phys. Chem. 67 (1963) 2679–2682.
Tada et al. 1997 [610] [39] M.C. Markham, M.C. Upreti, J. Catal. 4 (1965) 229–238.
Spray-pyrolysis Wang et al. 1998 [611] [40] D.R. Kennedy, M. Ritchie, J. Mackenzie, Trans. Faraday Soc. 54 (1958).
Ohko et al. 2001 [566] [41] A. Terenin, Y. Solonitzin, Discuss. Faraday Soc. 28 (1959) 28–35.
Electrophoretic deposition Fernandez et al. 1995 [612] [42] V.N. Filimonov, Doklady Akademii Nauk SSSR 154 (1964) 922–925.
Yanagida et al. 2005 [613] [43] S. Kato, F. Mashio, Kogyo Kagaku Zasshi (J. Chem. Soc. Japan., Indust. Chem.
Chemical bath deposition Gao et al. 2003 [614] Sect.) 67 (1964) 1136–1140.
CVD Goossens 1998 [615] [44] I.S. McLintock, M. Ritchie, Trans. Faraday Soc. 61 (1965) 1007–1016.
Ding et al. 2001 [616] [45] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, Environ. Sci. Technol. 22
(1988) 798–806.
O’Neill 2003 [617]
[46] F. Lohmann, Ber. Bunsenge. Phys. Chem. 70 (1966) 87–92.
RF Magnetron sputtering Zeman et al. 2002 [618]
[47] C. Wagner, W. Traud, Z. Elektrochem. 44 (1938) 391–454.
Song et al. 2003 [619]
[48] C. Wagner, W. Traud, F. Mansfeld, Corrosion 62 (2006) 843–855.
Kitano et al. 2005 [620] [49] S.R. Morrison, T. Freund, J. Chem. Phys. 47 (1967) 1543–1551.
Song et al. 2006 [621] [50] A. Fujishima, K. Honda, S. Kikuchi, Kogyo Kagaku Zasshi (J. Chem. Soc. Japan.,
Takeuchi et al. 2006 [622] Indust. Chem. Sect.) 72 (1969) 108–113.
Kitano et al. 2006 [453] [51] A. Fujishima, K. Honda, J. Chem. Soc. Japan. 74 (1971) 355 (in Japanese).
Electron-beam evaporation Yang et al. 2004 [623] [52] T. Inoue, T. Watanabe, A. Fujishima, K. Honda, Chem. Lett. (1977) 1073–1076.
Yang et al. 2004 [624] [53] A. Fujishima, T. Inoue, K. Honda, J. Am. Chem. Soc. 101 (1979) 5582–5588.
Vacuum evaporation Miyata et al. 2006 [625] [54] A. Fujishima, K. Kohayakawa, K. Honda, J. Electrochem. Soc. 122 (1975)
Plasma spray Berger-Keller et al. 2003 [626] 1487–1489.
Lee et al. 2003 [627] [55] A. Fujishima, K. Kohayakawa, K. Honda, Bull. Chem. Soc. Jpn. 48 (1975)
Anodic oxidation Mor et al. 2006 [628] 1041–1042.
Macak et al. 2007 [629] [56] S.N. Frank, A.J. Bard, J. Am. Chem. Soc. 99 (1977) 303–304.
Thermal oxidation Fujishima et al. 1975 [54] [57] S.N. Frank, A.J. Bard, J. Phys. Chem. 81 (1977) 1484–1488.
Inorganic–organic graded film Takami et al. 2002 [630] [58] S.N. Frank, A.J. Bard, J. Am. Chem. Soc. 99 (1977) 4667–4675.
Silica gel substrate Kobayakawa et al. 1998 [631] [59] B. Kraeutler, A.J. Bard, J. Am. Chem. Soc. 100 (1978) 2239–2240.
[60] H. Reiche, W.W. Dunn, A.J. Bard, J. Phys. Chem. 83 (1979) 2248–2251.
[61] E. Pelizzetti, M. Borgarello, E. Borgarello, N. Serpone, Chemosphere 17 (1988)
499–510.
[62] R.I. Bickley, T. Gonzalez-Carreno, J.S. Lees, L. Palmisano, R.J.D. Tilley, J. Solid
References State Chem. 92 (1991) 178–190.
[63] R.W. Matthews, in: D.F. Ollis, H. Al-Ekabi (Eds.), Photocatalytic Purfication
[1] E. Pelizzetti, N. Serpone (Eds.), Homogeneous and Heterogeneous Photocatal- and Treatment of Water and Air, Amsterdam, 1993, pp. 121–138.
ysis, D. Reidel Publishing Company, Dordrecht, 1986. [64] A. Mills, R.H. Davies, D. Worsley, Chem. Soc. Rev. 22 (1993) 417.
[2] N. Serpone, E. Pelizzetti (Eds.), Photocatalysis — Fundamentals and Applica- [65] L.A. Dibble, G.B. Raupp, Environ. Sci. Technol. 26 (1992) 492–495.
tions, John Wiley & Sons, New York, 1989. [66] J. Peral, D.F. Ollis, J. Catal. 136 (1992) 554–565.
[3] P.V. Kamat, Chem. Rev. 93 (1993) 267. [67] L.A. Phillips, G.B. Raupp, J. Molec. Catal. 77 (1992) 297–311.
[4] D.F. Ollis, H. Al-Ekabi (Eds.), Photocatalytic Purification and Treatment of [68] G.B. Raupp, C.T. Junio, Appl. Surf. Sci. 72 (1993) 321–327.
Water and Air, Elsevier, Amsterdam, 1993. [69] T.N. Obee, R.T. Brown, Environ. Sci. Technol. 29 (1995) 1223–1231.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 577

[70] D.M. Blake, Bibliography of Work on the Photocatalytic Removal of [124] J. Goniakowski, M.J. Gillan, Surf. Sci. 350 (1996) 145–158.
Hazardous Compounds from Water and Air, Golden, Colorado, USA, National [125] P.J.D. Lindan, C. Zhang, Phys. Rev. B 72 (2005) 075439-7.
Renewable Energy Laboratory, 1994. [126] P.J.D. Lindan, N.M. Harrison, M.J. Gillan, Phys. Rev. Lett. 80 (1998) 762.
[71] D.M. Blake, Bibliography of Work on the Photocatalytic Removal of [127] C. Zhang, P.J.D. Lindan, J. Chem. Phys. 118 (2003) 4620–4630.
Hazardous Compounds from Water and Air Update Number 1 to June, 1995, [128] P.J.D. Lindan, C. Zhang, Phys. Rev. Lett. 95 (2005) 029601-1.
Golden, Colorado, USA, National Renewable Energy Laboratory, 1995. [129] S. Wendt, et al., Surf. Sci. 598 (2005) 226–245.
[72] D.M. Blake, Bibliography of Work on the Photocatalytic Removal of [130] M.A. Henderson, Langmuir 12 (1996) 5093–5098.
Hazardous Compounds from Water and Air Update Number 2 to October, [131] U. Diebold, J. Lehman, T. Mahmoud, M. Kuhn, G. Leonardelli, W. Hebenstreit,
1996, Golden, Colorado, USA, National Renewable Energy Laboratory, 1996. M. Schmid, P. Varga, Surf. Sci. 411 (1998) 137–153.
[73] I. Sopyan, S. Murasawa, K. Hashimoto, A. Fujishima, Chem. Lett. (1994) 1994. [132] S. Suzuki, K.-I. Fukui, H. Onishi, Y. Iwasawa, Phys. Rev. Lett. 84 (2000) 2156.
[74] A.L. Linsebigler, G. Lu, J.J.T. Yates, Chem. Rev. 95 (1995) 735–758. [133] S. Wendt, J. Matthiesen, R. Schaub, E.K. Vestergaard, E. Laegsgaard,
[75] H. Matsubara, M. Takada, S. Koyama, K. Hashimoto, A. Fujishima, Chem. Lett. F. Besenbacher, B. Hammer, Phys. Rev. Lett. 96 (2006) 066107-4.
(1995) 767–768. [134] S. Wendt, et al. http://www.phys.au.dk/spm/movies/hydroxylation.mpg,
[76] N. Negishi, T. Iyoda, K. Hashimoto, A. Fujishima, Chem. Lett. (1995) 841–842. 2005.
[77] H. Zhang, J.F. Banfield, J. Phys. Chem. B 104 (2000) 3481–3487. [135] S. Wendt, et al. http://www.phys.au.dk/spm/movies/Double_OHbr_splitting.
[78] M.R. Ranade, A. Navrotsky, H.Z. Zhang, J.F. Banfield, S.H. Elder, A. Zaban, P.H. mpg, 2005.
Borse, S.K. Kulkarni, G.S. Doran, H.J. Whitfield, Proc. Nat. Acad. Sciences 99 [136] R. Schaub, E. Wahlstrom, A. Rønnau, E. Lægsgaard, I. Stensgaard,
(2002) 6476–6481. F. Besenbacher, Science 299 (2003) 377–379.
[79] V. Shklover, M.K. Nazeeruddin, S.M. Zakeeruddin, C. Barbe, A. Kay, T. Haibach, [137] O. Bikondoa, C.L. Pang, R. Ithnin, C.A. Muryn, H. Onishi, G. Thornton, Nature
W. Steurer, R. Hermann, H.U. Nissen, M. Grätzel, Chem. Mater. 9 (1997) Mater. 5 (2006) 189–192.
430–439. [138] Z. Zhang, O. Bondarchuk, B.D. Kay, J.M. White, Z. Dohnalek, J. Phys. Chem. B
[80] S.D. Burnside, V. Shklover, C. Barbe, P. Comte, F. Arendse, K. Brooks, M. Grätzel, 110 (2006) 21840–21845.
Chem. Mater. 10 (1998) 2419–2425. [139] S. Mezhenny, P. Maksymovych, T.L. Thompson, O. Diwald, D. Stahl, S.D. Walck,
[81] M. Ramamoorthy, D. Vanderbilt, R.D. King-Smith, Phys. Rev. B 49 (1994) J.T. Yates, Chem. Phys. Lett. 369 (2003) 152–158.
16721. [140] A.N. Shultz, W. Jang, W.M. Hetherington, D.R. Baer, L.-Q. Wang,
[82] R. Hengerer, B. Bolliger, M. Erbudak, M. Grätzel, Surf. Sci. 460 (2000) 162–169. M.H. Engelhard, Surf. Sci. 339 (1995) 114–124.
[83] Y. Liang, S. Gan, S.A. Chambers, E.I. Altman, Phys. Rev. B 63 (2001) 235402. [141] W.S. Epling, C.H.F. Peden, M.A. Henderson, U. Diebold, Surf. Sci. 412–413
[84] N. Ruzycki, G.S. Herman, L.A. Boatner, U. Diebold, Surf. Sci. 529 (2003) (1998) 333–343.
L239–L244. [142] M.A. Henderson, W.S. Epling, C.L. Perkins, C.H.F. Peden, U. Diebold, J. Phys.
[85] A. Beltran, L. Gracia, J. Andres, J. Phys. Chem. B 110 (2006) 23417–23423. Chem. B 103 (1999) 5328–5337.
[86] L.S. Dubrovinsky, N.A. Dubrovinskaia, V. Swamy, J. Muscat, N.M. Harrison, [143] M.A. Henderson, W.S. Epling, C.H.F. Peden, C.L. Perkins, J. Phys. Chem. B 107
R. Ahuja, B. Holm, B. Johansson, Nature 410 (2001) 653–654. (2003) 534–545.
[87] M. Mattesini, J.S. de Almeida, L. Dubrovinsky, N. Dubrovinskaia, B. Johansson, [144] A. Rothschild, A. Levakov, Y. Shapira, N. Ashkenasy, Y. Komem, Surf. Sci.
R. Ahuja, Phys. Rev. B 70 (2004) 115101-9. 532–535 (2003) 456–460.
[88] S. Andersson, A. Magnéli, Naturwissenschaften 43 (1956) 495–496. [145] T.L. Thompson, J.T. Yates, J. Phys. Chem. B 109 (2005) 18230–18236.
[89] L.A. Bursill, B.G. Hyde, Prog. Solid State Chem. 7 (1972) 177–253. [146] T.L. Thompson, J.T. Yates, J. Phys. Chem. B 110 (2006) 7431–7435.
[90] Y. Le Page, P. Strobel, J. Solid State Chem. 47 (1983) 6–15. [147] B.H. Soffer, J. Chem. Phys. 35 (1961) 940–945.
[91] Y. Le Page, P. Strobel, J. Solid State Chem. 44 (1982) 273–281. [148] G.J. Hill, J. Phys. D: Appl. Phys. (1968) 1151.
[92] J.B. Goodenough, Phys. Rev. 117 (1960) 1442. [149] O.W. Johnson, W.D. Ohlsen, P.I. Kingsbury, Phys. Rev. 175 (1968) 1102.
[93] M. Hirasawa, T. Seto, T. Orii, N. Aya, H. Shimura, Appl. Surf. Sci. 197–198 [150] O.W. Johnson, J. DeFord, J.W. Shaner, J. Appl. Phys. 44 (1973) 3008–3012.
(2002) 661–665. [151] J.W. DeFord, O.W. Johnson, J. Appl. Phys. 44 (1973) 3001–3007.
[94] M.D. Earle, Phys. Rev. 61 (1942) 56–62. [152] O.W. Johnson, S.H. Paek, J.W. DeFord, J. Appl. Phys. 46 (1975) 1026–1033.
[95] D.C. Cronemeyer, M.A. Gilleo, Phys. Rev. 82 (1951) 975. [153] J.B. Bates, R.A. Perkins, Phys. Rev. B 16 (1977) 3713.
[96] D.C. Cronemeyer, Phys. Rev. 87 (1952) 876. [154] J.B. Bates, J.C. Wang, R.A. Perkins, Phys. Rev. B 19 (1979) 4130.
[97] R.G. Breckenridge, W.R. Hosler, Phys. Rev. 91 (1953) 793. [155] S. Klauer, M. Wöhlecke, Europhys. Lett. 20 (1992) 439–443.
[98] H. Tang, K. Prasad, R. Sanjines, P.E. Schmid, F. Levy, J. Appl. Phys. 75 (1994) [156] J.V. Cathcart, R.A. Perkins, J.B. Bates, L.C. Manley, J. Appl. Phys. 50 (1979)
2042–2047. 4110–4119.
[99] L. Forro, O. Chauvet, D. Emin, L. Zuppiroli, H. Berger, F. Lévy, J. Appl. Phys. 75 [157] J. Robertson, P.W. Peacock, Thin Solid Films 445 (2003) 155–160.
(1994) 633. [158] P.W. Peacock, J. Robertson, Appl. Phys. Lett. 83 (2003) 2025–2027.
[100] E. Hendry, F. Wang, J. Shan, T.F. Heinz, M. Bonn, Phys. Rev. B 69 (2004) [159] D.J. Park, H.H. Nahm, C.H. Park, J. Korean Phys. Soc. 49 (2006) S473; S476.
081101-4. [160] M.V. Koudriachova, S.W. de Leeuw, N.M. Harrison, Phys. Rev. B 70 (2004)
[101] R.P. Feynman, Phys. Rev. 97 (1955) 660–665. 165421-5.
[102] B.O. Aduda, P. Ravirajan, K.L. Choy, J. Nelson, Int. J. Photoenergy 6 (2004) [161] J. Leconte, A. Markovits, M.K. Skalli, C. Minot, A. Belmajdoub, Surf. Sci. 497
141–147. (2002) 194–204.
[103] O.K. Varghese, D. Gong, M. Paulose, K.G. Ong, C.A. Grimes, Sens. Actuators B [162] S.H. Lim, J. Luo, Z. Zhong, W. Ji, J. Lin, Inorg. Chem. 44 (2005) 4124–4126.
93 (2003) 338–344. [163] G. Bromiley, A. Shiryaev, Phys. Chem. Minerals 33 (2006) 426–434.
[104] G.K. Mor, M.A. Carvalho, O.K. Varghese, M.V. Pishko, C.A. Grimes, J. Mater. Res. [164] D.A. Panayotov, J.J.T. Yates, Chem. Phys. Lett. 436 (2007) 204–208.
19 (2004) 628–634. [165] P.J. Boddy, J. Electrochem. Soc. 115 (1968) 199–203.
[105] Abu S. Zuruzi, Noel C. MacDonald, M. Moskovits, A. Kolmakov, Angew. Chem. [166] A. Fujishima, K. Honda, S. Kikuchi, J. Chem. Soc. Japan 72 (1969) 108 (in
Internat. Ed. 46 (2007) 4298–4301. Japanese).
[106] U. Diebold, Surf. Sci. Rep. 48 (2003) 53–229. [167] A. Fujishima, A. Sakamoto, K. Honda, Seisan Kenkyu 21 (1969) 450.
[107] D. Vogtenhuber, R. Podloucky, A. Neckel, S.G. Steinemann, A.J. Freeman, Phys. [168] L. Kavan, M. Graetzel, S.E. Gilbert, C. Klemenz, H.J. Scheel, J. Am. Chem. Soc.
Rev. B 49 (1994) 2099. 118 (1996) 6716–6723.
[108] R. Lindsay, A. Wander, A. Ernst, B. Montanari, G. Thornton, N.M. Harrison, [169] E.C. Dutoit, F. Cardon, W.P. Gomes, Ber. Bunsenge. Phys. Chem. 80 (1976)
Phys. Rev. Lett. 94 (2005) 246102-4. 475–481.
[109] V. Swamy, J. Muscat, J.D. Gale, N.M. Harrison, Surf. Sci. 504 (2002) 115–124. [170] R. De Gryse, W.P. Gomes, F. Cardon, J. Vennik, J. Electrochem. Soc. 122 (1975)
[110] H. Uetsuka, A. Sasahara, H. Onishi, Langmuir 20 (2004) 4782–4783. 711–712.
[111] R. Nakamura, N. Ohashi, A. Imanishi, T. Osawa, Y. Matsumoto, H. Koinuma, [171] M. Tomkiewicz, J. Electrochem. Soc. 126 (1979) 1505–1510.
Y. Nakato, J. Phys. Chem. B 109 (2005) 1648–1651. [172] D.L. Ullman, J. Electrochem. Soc. 127 (1980) 1321–1322.
[112] Z. Zhang, P. Fenter, N.C. Sturchio, M.J. Bedzyk, M.L. Machesky, [173] J. Schoonman, K. Vos, G. Blasse, J. Electrochem. Soc. 128 (1981) 1154–1157.
D.J. Wesolowski, Surf. Sci. 601 (2007) 1129–1143. [174] E.H. Rhoderick, R.H. Williams, Metal-Semiconductor Contacts, Clarendon
[113] M.A. Henderson, Surf. Sci. Rep. 46 (2002) 1–308. Press, Oxford, 1988.
[114] M.A. Henderson, Surf. Sci. 355 (1996) 151–166. [175] H.O. Finklea, J. Electrochem. Soc. 129 (1982) 2003–2008.
[115] M.S. Shultz, S. Baldelli, C. Schnitzer, D. Simonelli, J. Phys. Chem. B 106 (2002) [176] A. Goossens, M. Vazquez, D.D. Macdonald, Electrochim. Acta 41 (1996) 35–45.
5313–5324. [177] D.S. Ginley, M.L. Knotek, J. Electrochem. Soc. 126 (1979) 2163–2166.
[116] E.V. Stefanovich, T.N. Truong, Chem. Phys. Lett. 299 (1999) 623–629. [178] L.A. Harris, R. Schumacher, J. Electrochem. Soc. 127 (1980) 1186–1188.
[117] R. Schaub, P. Thostrup, N. Lopez, E. Laegsgaard, I. Stensgaard, J.K. Nørskov, [179] P. Clechet, C. Martelet, R. Olier, J.P. Thomas, M. Fallavier, J. Electrochem. Soc.
F. Besenbacher, Phys. Rev. Lett. 87 (2001) 266104. 130 (1983) 1795–1796.
[118] W. Langel, Surf. Sci. 496 (2002) 141–150. [180] H. Pelouchova, P. Janda, J. Weber, L. Kavan, J. Electroanalyt. Chem. 566 (2004)
[119] L.A. Harris, A.A. Quong, Phys. Rev. Lett. 93 (2004) 086105-4. 73–83.
[120] A.V. Bandura, D.G. Sykes, V. Shapovalov, T.N. Troung, J.D. Kubicki, [181] A. Ghicov, H. Tsuchiya, R. Hahn, J.M. Macak, A.G. Munoz, P. Schmuki,
R.A. Evarestov, J. Phys. Chem. B 108 (2004) 7844–7853. Electrochem. Commun. 8 (2006) 528–532.
[121] L.A. Harris, A.A. Quong, Phys. Rev. Lett. 95 (2005) 029602-1. [182] G. Nogami, J. Electrochem. Soc. 132 (1985) 76–81.
[122] H. Perron, J. Vandenborre, C. Domain, R. Drot, J. Roques, E. Simoni, [183] J.C. Wang, J. Electrochem. Soc. 134 (1987) 1915–1920.
J.J. Ehrhardt, H. Catalette, Surf. Sci. 601 (2007) 518–527. [184] J.C. Wang, J.B. Bates, Solid State Ion. 18&19 (1986) 224–228.
[123] P.J.D. Lindan, N.M. Harrison, J.M. Holender, M.J. Gillan, Chem. Phys. Lett. 261 [185] B.D. Cahan, C.T. Chen, J. Electrochem. Soc. 129 (1982) 700–705.
(1996) 246–252. [186] B.D. Cahan, C.T. Chen, J. Electrochem. Soc. 129 (1982) 474–480.
578 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

[187] F. Cao, G. Oskam, P.C. Searson, J.M. Stipkala, T.A. Heimer, F. Farzad, G.J. Meyer, [241] A. Furube, T. Asahi, H. Masuhara, H. Yamashita, M. Anpo, Chem. Phys. Lett.
J. Phys. Chem. 99 (1995) 11974–11980. 336 (2001) 424–430.
[188] A. Muñoz, Q. Chen, P. Schmuki, J. Solid State Electrochem. 11 (2007) [242] A. Furube, T. Asahi, H. Masuhara, H. Yamashita, M. Anpo, J. Phys. Chem. B 103
1077–1084. (1999) 3120–3127.
[189] J.J. Kim, D.A. Tryk, T. Amemiya, K. Hashimoto, A. Fujishima, J. Electroanalyt. [243] P.V. Kamat, I. Bedja, S. Hotchandani, J. Phys. Chem. 98 (1994) 9137–9142.
Chem. 435 (1997) 31–38. [244] A. Henglein, Chem. Rev. 89 (1989) 1861–1873.
[190] J.J. Kim, D.A. Tryk, T. Amemiya, K. Hashimoto, A. Fujishima, J. Electroanalyt. [245] U. Kölle, J. Moser, M. Grätzel, Inorg. Chem. 24 (1985) 2253–2258.
Chem. 433 (1997) 9–17. [246] T. Yoshihara, R. Katoh, A. Furube, Y. Tamaki, M. Murai, K. Hara, S. Murata,
[191] B. Wiecek, U. Twardoch, J. Phys. Chem. Solids 65 (2004) 263–268. H. Arakawa, M. Tachiya, J. Phys. Chem. B 108 (2004) 3817–3823.
[192] B. Parkinson, F. Decker, J.F. Juliaõ, M. Abramovich, H.C. Chagas, Electrochim. [247] N. Serpone, D. Lawless, R. Khairutdinov, E. Pelizzetti, J. Phys. Chem. 99 (1995)
Acta 25 (1980) 521–525. 16655–16661.
[193] F. Fabregat-Santiago, I. Mora-Sero, G. Garcia-Belmonte, J. Bisquert, J. Phys. [248] G. Rothenberger, J. Moser, M. Graetzel, N. Serpone, D.K. Sharma, J. Am. Chem.
Chem. B 107 (2003) 758–768. Soc. 107 (1985) 8054–8059.
[194] L. Kavan, K. Kratochvilova, M. Grätzel, J. Electroanalyt. Chem. 394 (1995) [249] D. Lawless, N. Serpone, D. Meisel, J. Phys. Chem. 95 (1991) 5166–5170.
93–102. [250] D.P. Colombo Jr., R.M. Bowman, J. Phys. Chem. 100 (1996) 18445–18449.
[195] L. Kavan, M. Grätzel, J. Rathousky, A. Zukalb, J. Electrochem. Soc. 143 (1996) [251] D.P. Colombo Jr., R.M. Bowman, J. Phys. Chem. 99 (1995) 11752–11756.
394–400. [252] D.W. Bahnemann, M. Hilgendorff, R. Memming, J. Phys. Chem. B 101 (1997)
[196] B. O’Regan, M. Grätzel, D. Fitzmaurice, J. Phys. Chem. 95 (1991) 10525–10528. 4265–4275.
[197] G. Redmond, D. Fitzmaurice, M. Grätzel, J. Phys. Chem. 97 (1993) 6951–6954. [253] D. Bahnemann, A. Henglein, J. Lilie, L. Spanhel, J. Phys. Chem. 88 (1984)
[198] G. Redmond, D. Fitzmaurice, J. Phys. Chem. 97 (1993) 1426–1430. 709–711.
[199] G. Rothenberger, D. Fitzmaurice, M. Grätzel, J. Phys. Chem. 96 (1992) [254] C. Colbeau-Justin, M. Kunst, D. Huguenin, J. Mater. Sci. 38 (2003) 2429–2437.
5983–5986. [255] A. Yamakata, T.-A. Ishibashi, H. Onishi, Chem. Phys. Lett. 333 (2001) 271–277.
[200] G. Boschloo, D. Fitzmaurice, J. Phys. Chem. B 103 (1999) 7860–7868. [256] A. Yamakata, T. Ishibashi, H. Onishi, J. Phys. Chem. B 105 (2001) 7258–7262.
[201] G. Boschloo, D. Fitzmaurice, J. Phys. Chem. B 103 (1999) 2228–2231. [257] S.H. Szczepankiewicz, A.J. Colussi, M.R. Hoffmann, J. Phys. Chem. B 104 (2000)
[202] L.A. Lyon, J.T. Hupp, J. Phys. Chem. B 103 (1999) 4623–4628. 9842–9850.
[203] P. Clechet, C. Martelet, J.R. Martin, R. Olier, Electrochim. Acta 24 (1979) [258] A. Yamakata, T.-A. Ishibashi, K. Takeshita, H. Onishi, Topics Catal. V35 (2005)
457–461. 211–216.
[204] J.M. Kesselman, G.A. Shreve, M.R. Hoffmann, N.S. Lewis, J. Phys. Chem. 98 [259] S.T. Martin, H. Herrmann, M.R. Hoffmann, J. Chem. Soc. Faraday Trans. 90
(1994) 13385–13395. (1994) 3323–3330.
[205] D. Tafalla, P. Salvador, Ber. Bunsenges. Phys. Chem. 91 (1987) 475–479. [260] S.T. Martin, H. Herrmann, W. Choi, M.R. Hoffmann, J. Chem. Soc. Faraday
[206] A. Tsujiko, H. Itoh, T. Kisumi, A. Shiga, K. Murakoshi, Y. Nakato, J. Phys. Chem. Trans. 90 (1994) 3315–3322.
B 106 (2002) 5878–5885. [261] X. Yang, N. Tamai, Phys. Chem. Chem. Phys. 3 (2001) 3393–3398.
[207] M.R. Tarasevich, A. Sadkowski, E. Yeager, B.E. Conway, J.O’M. Bockris, [262] Y. Tamaki, A. Furube, M. Murai, K. Hara, R. Katoh, M. Tachiya, Phys. Chem.
E. Yeager, S.U.M. Khan, R.E. White (Eds.), Kinetics and Mechanisms of Chem. Phys. 9 (2007) 1453–1460.
Electrode Processes, New York, 1983, pp. 301–398. [263] Y. Tamaki, A. Furube, R. Katoh, M. Murai, K. Hara, H. Arakawa, M. Tachiya,
[208] Y.V. Pleskov, Y.Y. Gurevich, Semiconductor Photoelectrochemistry, Consul- C. R. Chimie 9 (2006) 268–274.
tants Bureau, New York, 1986. [264] Y. Tamaki, A. Furube, M. Murai, K. Hara, R. Katoh, M. Tachiya, J. Am. Chem.
[209] J.F. Dewald, J. Phys. Chem. Solids 14 (1960) 155–161. Soc. 128 (2006) 416–417.
[210] J.F. Dewald, Bell Syst. Tech. J. 39 (1960) 615–639.
[265] K. Iwata, T. Takaya, H. Hamaguchi, A. Yamakata, T. Ishibashi, H. Onishi,
[211] W.H. Brattain, P.J. Boddy, J. Electrochem. Soc. 109 (1962) 574–582.
H. Kuroda, J. Phys. Chem. B 108 (2004) 20233–20239.
[212] H. Gerischer, J. Electrochem. Soc. 113 (1966) 1174–1182.
[266] D.C. Hurum, A.G. Agrios, K.A. Gray, T. Rajh, M.C. Thurnauer, J. Phys. Chem. B
[213] A. Fujishima, K. Honda, Seisan Kenkyu 22 (1970) 478.
107 (2003) 4545–4549.
[214] A. Fujishima, K. Honda, Bull. Chem. Soc. Jpn. 44 (1971) 1148–1150.
[267] O. Micic, Y. Zhang, K.R. Cromack, A. Trifunac, M. Thurnauer, J. Phys. Chem. 97
[215] P. Salvador, C. Gutierrez, J. Phys. Chem. 88 (1984) 3696–3698.
(1993) 13284–13288.
[216] P. Salvador, J. Phys. Chem. 89 (1985) 3863–3869.
[268] T. Berger, M. Sterrer, O. Diwald, E. Knozinger, D. Panayotov, T.L. Thompson,
[217] Y. Nakato, H. Akanuma, Y. Magari, S. Yae, J.I. Shimizu, H. Mori, J. Phys. Chem.
J.T. Yates, J. Phys. Chem. B 109 (2005) 6061–6068.
B 101 (1997) 4934–4939.
[269] Y. Nakaoka, Y. Nosaka, J. Photochem. Photobiol. A: Chem. 110 (1997)
[218] T. Kisumi, A. Tsujiko, K. Murakoshi, Y. Nakato, J. Electroanalyt. Chem. 545
299–305.
(2003) 99–107.
[270] O.I. Micic, Y. Zhang, K.R. Cromack, A.D. Trifunac, M.C. Thurnauer, J. Phys.
[219] R. Nakamura, Y. Nakato, J. Am. Chem. Soc. 126 (2004) 1290–1298.
Chem. 97 (1993) 7277–7283.
[220] R. Nakamura, T. Okamura, N. Ohashi, A. Imanishi, Y. Nakato, J. Am. Chem. Soc.
127 (2005) 12975–12983. [271] R.F. Howe, M. Grätzel, J. Phys. Chem. 89 (1985) 4495–4499.
[221] B. Neumann, P. Bogdanoff, H. Tributsch, S. Sakthivel, H. Kisch, J. Phys. Chem. [272] R.B. Lauer, J.R.R. Addiss, A.K. Ghosh, J. Appl. Phys. 42 (1971) 3508–3512.
B 109 (2005) 16579–16586. [273] J.R.R. Addiss, A.K. Ghosh, F.G. Wakim, Appl. Phys. Lett. 12 (1968) 397–400.
[222] Z.w. Qu, G.J. Kroes, J. Phys. Chem. B 110 (2006) 23306–23314. [274] A.K. Ghosh, F.G. Wakim, R.R. Addiss, Phys. Rev. 184 (1969) 979.
[223] T.L. Villarreal, R. Gomez, M. Neumann-Spallart, N. Alonso-Vante, P. Salvador, [275] G.K. Boschloo, A. Goossens, J. Phys. Chem. 100 (1996) 19489–19494.
J. Phys. Chem. B 108 (2004) 15172–15181. [276] S. Leytner, J.T. Hupp, Chem. Phys. Lett. 330 (2000) 231–236.
[224] T. Inoue, C. Weber, A. Fujishima, K. Honda, Bull. Chem. Soc. Jpn. 53 (1980) 234. [277] N. Serpone, D. Lawless, R. Khairutdinov, J. Phys. Chem. 99 (1995)
[225] B. Kraeutler, A.J. Bard, J. Am. Chem. Soc. 100 (1978) 4317–4318. 16646–16654.
[226] C.-M. Wang, A. Heller, H. Gerischer, J. Am. Chem. Soc. 114 (1992) 5230–5234. [278] N.D. Abazovic, M.I. Comor, M.D. Dramicanin, D.J. Jovanovic, S.P. Ahrenkiel,
[227] I. Izumi, W.W. Dunn, K.O. Wilbourn, F.-R.F. Fan, A.J. Bard, J. Phys. Chem. 84 J.M. Nedeljkovic, J. Phys. Chem. B 110 (2006) 25366–25370.
(1980) 3207–3210. [279] D.A. Panayotov, J.J.T. Yates, Chem. Phys. Lett. 410 (2005) 11–17.
[228] H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, J. Electroanal. Chem. 379 (1994) [280] M.A. Henderson, J.M. White, H. Uetsuka, H. Onishi, J. Am. Chem. Soc. 125
199–205. (2003) 14974–14975.
[229] H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, A. Heller, J. Phys. Chem. 99 [281] C. Di Valentin, G. Pacchioni, A. Selloni, Phys. Rev. Lett. 97 (2006) 166803-4.
(1995) 11896–11900. [282] H.P. Boehm, Discuss. Faraday Soc. 52 (1971) 264–275.
[230] K. Ikeda, H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, J. Phys. Chem. B 101 [283] Y. Nosaka, T. Daimon, A.Y. Nosaka, Y. Murakami, Phys. Chem. Chem. Phys. 6
(1997) 2617–2620. (2004) 2917–2918.
[231] K. Ikeda, K. Hashimoto, A. Fujishima, J. Electroanal. Chem. 437 (1997) [284] T. Tachikawa, M. Fujitsuka, T. Majima, J. Phys. Chem. C 111 (2007) 5259–5275.
241–244. [285] W. Xu, D. Raftery, J. Phys. Chem. B 105 (2001) 4343–4349.
[232] H. Maeda, K. Ikeda, K. Hashimoto, K. Ajito, M. Morita, A. Fujishima, J. Phys. [286] S. Pilkenton, S.-J. Hwang, D. Raftery, J. Phys. Chem. B 103 (1999)
Chem. B 103 (1999) 3213–3217. 11152–11160.
[233] K. Ikeda, H. Sakai, R. Baba, K. Hashimoto, A. Fujishima, Chem. Lett. (1995) [287] I.A. Shkrob, M.C. Sauer, D. Gosztola, J. Phys. Chem. B 108 (2004) 12512–12517.
979–980. [288] I.A. Shkrob, M.C. Sauer, J. Phys. Chem. B 108 (2004) 12497–12511.
[234] H. Gerischer, A. Heller, J. Phys. Chem. 95 (1991) 5261–5267. [289] S. Tojo, T. Tachikawa, M. Fujitsuka, T. Majima, Phys. Chem. Chem. Phys. 6
[235] J. Schwitzgebel, J.G. Ekerdt, H. Gerischer, A. Heller, J. Phys. Chem. 99 (1995) (2004) 960–964.
5633–5638. [290] A.Y. Nosaka, T. Fujiwara, H. Yagi, H. Akutsu, Y. Nosaka, Langmuir 19 (2003)
[236] H. Gerischer, A. Heller, J. Electrochem. Soc. 139 (1992) 113–118. 1935–1937.
[237] G. Nicolis, I. Prigogine, Self-Organization in Nonequilibrium Systems: From [291] C.y. Wang, H. Groenzin, M.J. Shultz, J. Am. Chem. Soc. 126 (2004) 8094–8095.
Dissipative Structures to Order through Fluctuations, John Wiley & Sons, New [292] G. Riegel, J.R. Bolton, J. Phys. Chem. 99 (1995) 4215–4224.
York, 1977. [293] H. Noda, K. Oikawa, H. Ohya-Nishiguchi, H. Kamada, Bull. Chem. Soc. Jpn. 67
[238] A.M. Peiro, C. Colombo, G. Doyle, J. Nelson, A. Mills, J.R. Durrant, J. Phys. Chem. (1994) 2031–2037.
B 110 (2006) 23255–23263. [294] C.D. Jaeger, A.J. Bard, J. Phys. Chem. 83 (1979) 3146–3152.
[239] A.V. Emeline, V.K. Ryabchuk, N. Serpone, J. Phys. Chem. B 109 (2005) [295] M.A. Grela, M.E.J. Coronel, A.J. Colussi, J. Phys. Chem. B 100 (1996)
18515–18521. 16940–16946.
[240] Y. Wang, K. Hang, N.A. Anderson, T. Lian, J. Phys. Chem. B 107 (2003) [296] R.B. Draper, M.A. Fox, Langmuir 6 (1990) 1396–1402.
9434–9440. [297] R.B. Draper, M.A. Fox, J. Phys. Chem. 94 (1990) 4628–4634.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 579

[298] T. Tachikawa, S. Tojo, M. Fujitsuka, T. Majima, Langmuir 20 (2004) [352] T. Tatsuma, S.-i. Tachibana, T. Miwa, D.A. Tryk, A. Fujishima, J. Phys. Chem. B
2753–2759. 103 (1999) 8033–8035.
[299] S. Tojo, T. Tachikawa, M. Fujitsuka, T. Majima, Chem. Phys. Lett. 384 (2004) [353] M.C. Lee, W. Choi, J. Phys. Chem. B 106 (2002) 11818–11822.
312–316. [354] J.S. Park, W. Choi, Langmuir 20 (2004) 11523–11527.
[300] T. Tachikawa, S. Tojo, M. Fujitsuka, T. Majima, J. Phys. Chem. B 108 (2004) [355] S. Cho, W. Choi, J. Photochem. Photobiol. A: Chem. 143 (2001) 221–228.
5859–5866. [356] K. Kawahara, Y. Ohko, T.T.a.A. Fujishima, Phys. Chem. Chem. Phys. 5 (2003)
[301] K. Ishibashi, A. Fujishima, T. Watanabe, K. Hashimoto, J. Photochem. Photobiol 4764–4766.
A: Chem. 134 (2000) 139–142. [357] H. Haick, Y. Paz, Chem. Phys. Chem. 4 (2003) 617–620.
[302] Y. Nosaka, S. Komori, K. Yawata, T. Hirakawa, A.Y. Nosaka, Phys. Chem. Chem. [358] H. Haick, Y. Paz, J. Phys. Chem. B 105 (2001) 3045–3051.
Phys. (2003) 4731–4735. [359] Y. Ishikawa, Y. Matsumoto, Y. Nishida, S. Taniguchi, J. Watanabe, J. Am. Chem.
[303] P. Salvador, J. Electrochem. Soc. 128 (1981) 1895–1900. Soc. 125 (2003) 6558–6562.
[304] C. Gutierrez, P. Salvador, J. Electrochem. Soc. 133 (1986) 924–929. [360] T. Tatsuma, W. Kubo, A. Fujishima, Langmuir 18 (2002) 9632–9634.
[305] Y. Murakami, E. Kenji, A.Y. Nosaka, Y. Nosaka, J. Phys. Chem. B 110 (2006) [361] W. Kubo, T. Tatsuma, Analyt. Sci. 20 (2004) 591–593.
16808–16811. [362] W. Kubo, T. Tatsuma, J. Am. Chem. Soc. 128 (2006) 16034–16035.
[306] O. Legrini, E. Oliveros, A.M. Braun, Chem. Rev. 93 (1993) 671–698. [363] W. Kubo, T. Tatsuma, A. Fujishima, H. Kobayashi, J. Phys. Chem. B 108 (2004)
[307] S.J. Hwang, C. Petucci, D. Raftery, J. Am. Chem. Soc. 120 (1998) 4388–4397. 3005–3009.
[308] G. Munuera, A. Navio, V. Rives-Arnau, J. Chem. Soc., Faraday Trans. 1 77 (1981) [364] A.Y. Nosaka, E. Kojima, T. Fujiwara, H. Yagi, H. Akutsu, Y. Nosaka, J. Phys.
2747–2749. Chem. B 107 (2003) 12042–12044.
[309] T. Daimon, Y. Nosaka, J. Phys. Chem. C 111 (2007) 4420–4424. [365] M.L. Sauer, D.F. Ollis, J. Catal. 163 (1996) 215–217.
[310] K. Hirakawa, T. Hirano, Chem. Lett. 35 (2006) 832. [366] Y. Ohko, K. Hashimoto, A. Fujishima, J. Phys. Chem. A 101 (1997) 8057–8062.
[311] J. Janczyk, E. Krakowska, G. Stochel, W. Macyk, J. Am. Chem. Soc. 128 (2006) [367] Y. Ohko, D.A. Tryk, K. Hashimoto, A. Fujishima, J. Phys. Chem. B 102 (1998)
15574–15575. 2699–2704.
[312] T. Berger, M. Sterrer, O. Diwald, E. Knözinger, Chem. Phys. Chem. 6 (2005) [368] N.N. Lichtin, M. Avudaithai, Environ. Sci. Technol. 30 (1996) 2014–2020.
2104. [369] T.N. Obee, S.O. Hay, Environ. Sci. Technol. 31 (1997) 2034–2038.
[313] K.-I. Ishibashi, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys. Chem. B 104 [370] T. Sano, N. Negishi, K. Takeuchi, S. Matsuzawa, Solar Energy 77 (2004)
(2000) 4934–4938. 543–552.
[314] K. Ishibashi, Y. Nosaka, K. Hashimoto, A. Fujishima, J. Phys. Chem. B 102 (1998) [371] H. Einaga, S. Futamura, T. Ibusuki, Appl. Catal. B: Environ. 38 (2002) 215–225.
2117–2120. [372] M.C. Blount, J.L. Falconer, Appl. Catal. B: Environ. 39 (2002) 39–50.
[315] J.M. Coronado, A.J. Maira, J.C. Conesa, K.L. Yeung, V. Augugliaro, J. Soria, [373] W.-K. Jo, K.-H. Park, Chemosphere 57 (2004) 555–565.
Langmuir 17 (2001) 5368–5374. [374] J. Peral, D.F. Ollis, J. Molec, Catal. A: Chem. 115 (1997) 347–354.
[316] M. Anpo, N. Aikawa, Y. Kubokawa, M. Che, C. Louis, E. Gianello, J. Phys. Chem. [375] N. González-García, J.A. Ayllón, X. Doménech, J. Peral, Appl. Catal. B: Environ.
89 (1985) 5689–5694. 52 (2004) 69–77.
[317] Y. Nosaka, M. Nakamura, T. Hirakawa, Phys. Chem. Chem. Phys. 4 (2002) [376] R.-D. Sun, A. Nakajima, T. Watanabe, K. Hashimoto, J. Photochem. Photobiol.
1088–1092. A: Chem. 154 (2003) 203–209.
[318] T. Hirakawa, H. Kominami, B. Ohtani, Y. Nosaka, J. Phys. Chem. B 105 (2001) [377] J.A. Moss, S.H. Szczepankiewicz, E. Park, M.R. Hoffmann, J. Phys. Chem. B 109
6993–6999. (2005) 19779–19785.
[319] R. Nakamura, A. Imanishi, K. Murakoshi, Y. Nakato, J. Am. Chem. Soc. 125 [378] H. Einaga, T. Ibusuki, S. Futamura, Environ. Sci. Technol. 38 (2004) 285–289.
(2003) 7443–7450. [379] D.F. Ollis, E. Pelizzetti, N. Serpone, Environ. Sci. Technol. 25 (1991)
[320] A. Tilocca, C. DiValentin, A. Selloni, J. Phys. Chem. B 109 (2005) 20963–20967. 1522–1529.
[321] A.L. Attwood, D.M. Murphy, J.L. Edwards, T.A. Egerton, R.W. Harrison, Res. [380] D. Bahnemann, Solar Energy 77 (2004) 445–459.
Chem. Intermed. 29 (2003) 449–465. [381] J.M. Herrmann, Topics Catal. 34 (2005) 49–65.
[322] H. Goto, Y. Hanada, T. Ohno, M. Matsumura, J. Catal. 225 (2004) 223–229. [382] J. Blanco-Galvez, P. Fernandez-Ibanez, S. Malato-Rodriguez, J. Solar Energy
[323] X. Li, W.S. Jenks, J. Am. Chem. Soc. 122 (2000) 11864–11870. Eng. 129 (2007) 4–15.
[324] A.L. Attwood, J.L. Edwards, C.C. Rowlands, D.M. Murphy, J. Phys. Chem. A 107 [383] A. Salinaro, A.V. Emeline, J. Zhao, H. Hidaka, V.K. Ryabchuk, N. Serpone, Pure
(2003) 1779–1782. Appl. Chem. 71 (1999) 321–335.
[325] C.A. Jenkins, D.M. Murphy, J. Phys. Chem. B 103 (1999) 1019–1026. [384] N. Serpone, A. Salinaro, Pure Appl. Chem. 71 (1999) 303–320.
[326] A.R. Gonzalez-Elipe, M. Che, J. Chim. Phys. 79 (1982) 535. [385] N. Serpone, J. Photochem. Photobiol. A: Chem. 104 (1997) 1–12.
[327] E. Carter, A.F. Carley, D.M. Murphy, Chem. Phys. Chem. 8 (2007) 113–123. [386] R. Enríquez, A.G. Agrios, P. Pichat, Catal. Today 120 (2007) 196–202.
[328] P. Pichat, H. Courbon, R. Enriquez, T.T.Y. Tan, R. Amal, Res. Chem. Intermed. [387] A.G. Agrios, P. Pichat, J. Photochem. Photobiol. A: Chem. 180 (2006) 130–135.
33 (2007) 239–250. [388] R. Enriquez, P. Pichat, J. Environ. Science Health A 41 (2006) 955–966.
[329] D.S. Muggli, J.L. Falconer, J. Catal. 187 (1999) 230–237. [389] S. Sakthivel, M.V. Shankar, M. Palanichamy, B. Arabindoo, D.W. Bahnemann,
[330] D.S. Muggli, J.L. Falconer, J. Catal. 191 (2000) 318–325. V. Murugesan, Water Res. 38 (2004) 3001–3008.
[331] K. Yoshida, J. Yamasaki, N. Tanaka, Nanotechnol. (2004) S349. [390] M. Kositzi, A. Antoniadis, I. Poulios, I. Kiridis, S. Malato, Solar Energy 77 (2004)
[332] K. Yoshida, J. Yamasaki, N. Tanaka, Appl. Phys. Lett. 84 (2004) 2542–2544. 591–600.
[333] K. Yoshida, T. Nanbara, J. Yamasaki, N. Tanaka, J. Appl. Phys. 99 (2006) [391] T.E. Agustina, H.M. Ang, V.K. Vareek, J. Photochem. Photobiol. C: Photochem.
084908-8. Rev. 6 (2005) 264–273.
[334] M.T. Chen, C.F. Lien, L.F. Liao, J.L. Lin, J. Phys. Chem. B 107 (2003) 3837–3843. [392] T. Watanabe, K. Hashimoto, A. Fujishima, 1st International Conference on
[335] T.L. Thompson, D.A. Panayotov, J.T. Yates, I. Martyanov, K. Klabunde, J. Phys. TiO2 Photocatalytic Purification and Treatment of Water and Air, 1992.
Chem. B 108 (2004) 17857–17865. [393] S. Sitkiewitz, A. Heller, Nouv. J. Chem. 20 (1996) 233–241.
[336] J. Zhuang, C.N. Rusu, J.T. Yates, J. Phys. Chem. B 103 (1999) 6957–6967. [394] Y. Paz, Z. Luo, L. Rabenberg, A. Heller, J. Mater. Res. 10 (1996) 2842.
[337] S.A. Larson, J.A. Widegren, J.L. Falconer, J. Catal. 157 (1995) 611–625. [395] Y. Paz, A. Heller, J. Mater. Res. 12 (1997) 2759.
[338] P.A. Morris Hotsenpiller, J.D. Bolt, W.E. Farneth, J.B. Lowekamp, G.S. Rohrer, J. [396] P. Sawunyama, L. Jiang, A. Fujishima, K. Hashimoto, J. Phys. Chem. B 101
Phys. Chem. B 102 (1998) 3216–3226. (1997) 11000–11003.
[339] J.B. Lowekamp, G.S. Rohrer, P.A. Morris Hotsenpiller, J.D. Bolt, W.E. Farneth, J. [397] P. Sawunyama, L. Jiang, A. Fujishima, K. Hashimoto, Langmuir 15 (1999)
Phys. Chem. B 102 (1998) 7323–7327. 3551–3556.
[340] Y. Yamamoto, K. Nakajima, T. Ohsawa, Y. Matsumoto, H. Koinuma, Japan. J. [398] V. Romeas, P. Pichat, C. Guillard, T. Chopin, C. Lehaut, Ind. Eng. Chem. Res. 38
Appl. Phys. 44 (2005) L511–L514. (1999) 3878–3885.
[341] Y. Nakato, H. Akanuma, J.-i. Shimizu, Y. Magari, J. Electroanalyt. Chem. 396 [399] V. Roméas, P. Pichat, C. Guillard, T. Chopin, C. Lehaut, New J. Chem. 23 (1999)
(1995) 35–39. 365–374.
[342] A. Tsujiko, T. Kisumi, Y. Magari, K. Murakoshi, Y. Nakato, J. Phys. Chem. B 104 [400] T. Minabe, D.A. Tryk, P. Sawunyama, Y. Kikuchi, K. Hashimoto, A. Fujishima,
(2000) 4873–4879. J. Photochem. Photobiol. A: Chem. 137 (2000) 53–62.
[343] T. Sugiura, S. Itoh, T. Ooi, T. Yoshida, K. Kuroda, H. Minoura, J. Electroanalyt. [401] J.T. Remillard, J.R. McBride, K.E. Nietering, A.R. Drews, X. Zhang, J. Phys. Chem.
Chem. 473 (1999) 204–208. B 104 (2000) 4440–4447.
[344] T. Sugiura, T. Yoshida, H. Minoura, Electrochem. Solid-State Lett. 1 (1998) [402] E. Puzenat, P. Pichat, J. Photochem. Photobiol. A: Chem. 160 (2003) 127–133.
175–177. [403] A. Mills, A. Lepre, N. Elliott, S. Bhopal, I.P. Parkin, S.A. O’Neill, J. Photochem.
[345] T. Ohno, K. Sarukawa, M. Matsumura, New J. Chem. 26 (2002) 1167–1170. Photobiol. A: Chem. 160 (2003) 213–224.
[346] T. Taguchi, Y. Saito, K. Sarukawa, T. Ohno, M. Matsumura, New J. Chem. 27 [404] A. Mills, J. Wang, J. Photochem. Photobiol. A: Chem. 182 (2006) 181–186.
(2003). [405] J.R. Peller, R.L. Whitman, S. Griffith, P. Harris, C. Peller, J. Scalzitti, J.
[347] Y. Kikuchi, K. Sunada, T. Iyoda, K. Hashimoto, A. Fujishima, J. Photochem. Photochem. Photobiol. A: Chem. 186 (2007) 212–217.
Photobiol. A: Chem. 106 (1997) 51–56. [406] C. McCullagh, J.M.C. Robertson, D.W. Bahnemann, P.K.J. Robertson, Res. Chem.
[348] K. Naito, T. Tachikawa, S.C. Cui, A. Sugimoto, M. Fujitsuka, T. Majima, J. Am. Intermed. 33 (2007) 359–375.
Chem. Soc. 128 (2006) 16430–16431. [407] A. Fujishima, J. Ohtsuki, T. Yamashita, S. Hayakawa, Photochemistry and
[349] J.S. Park, W. Choi, Chem. Lett. (2005) 1630. Photobiology (Proceedings of the 8th Annual Meeting of the Japanese Society
[350] T. Tatsuma, S.-i. Tachibana, A. Fujishima, J. Phys. Chem. B 105 (2001) for Photomedicine and Photobiology) 1986, 45.
6987–6992. [408] T. Matsunaga, R. Tomoda, T. Nakajima, H. Wake, FEMS Microbiol. Lett. 29
[351] K. Naito, T. Tachikawa, M. Fujitsuka, T. Majima, J. Phys. Chem. B 109 (2005) (1985) 211–214.
23138–23140. [409] R. Cai, K. Hashimoto, Y. Kubota, A. Fujishima, Chem. Lett. (1992) 427–430.
580 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

[410] R. Cai, K. Hashimoto, K. Itoh, Y. Kubota, A. Fujishima, Bull. Chem. Soc. Jpn. 64 [461] E.A. Reyes-Garcia, Y. Sun, K. Reyes-Gil, D. Raftery, J. Phys. Chem. C 111 (2007)
(1991) 1268–1273. 2738–2748.
[411] K.P. Kühn, I.F. Chaberny, K. Massholder, M. Stickler, V.W. Benz, H.-G. Sonntag, [462] H. Chen, A. Nambu, W. Wen, J. Graciani, Z. Zhong, J.C. Hanson, E. Fujita,
L. Erdinger, Chemosphere 53 (2003) 71–77. J.A. Rodriguez, J. Phys. Chem. C 111 (2007) 1366–1372.
[412] K. Sunada, Y. Kikuchi, K. Hashimoto, A. Fujishima, Environ. Sci. Technol. 32 [463] C. Belver, R. Bellod, S.J. Stewart, F.G. Requejo, M. Fernández-García, Appl.
(1998) 726–728. Catal. B: Environ. 65 (2006) 309–314.
[413] E.J. Wolfrum, J. Huang, D.M. Blake, P.C. Maness, Z. Huang, J. Fiest, W.A. Jacoby, [464] S.-K. Joung, T. Amemiya, M. Murabayashi, K. Itoh, Chem. Eur. J. 12 (2006)
Environ. Sci. Technol. 36 (2002) 3412–3419. 5526–5534.
[414] K. Sunada, T. Watanabe, K. Hashimoto, J. Photochem. Photobiol. A: Chem. 156 [465] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Science 297 (2002) 2243–2245.
(2003) 227–233. [466] A. Fujishima, Science 301 (2003) 1673a.
[415] W.A. Jacoby, B.C. Maness, E.J. Wolfrum, D.M. Blake, J.A. Fennell, Environ. Sci. [467] C. Hägglund, M. Grätzel, B. Kasemo, Science 301 (2003) 1673b.
Technol. 32 (1998) 2650–2653. [468] K.S. Lackner, Science 301 (2003).
[416] K. Sunada, T. Watanabe, K. Hashimoto, Environ. Sci. Technol. 37 (2003) [469] H. Irie, Y. Watanabe, K. Hashimoto, Chem. Lett. 32 (2003) 772–773.
4785–4789. [470] J.H. Park, S. Kim, A.J. Bard, Nano Lett. 6 (2006) 24–28.
[417] J.-M. Hermann, M.N. Mozzanega, P. Pichat, J. Photochem. 22 (1983) 333. [471] S. Sakthivel, H. Kisch, Angew. Chem. Internat. Ed. 42 (2003) 4908–4911.
[418] J.-M. Herrmann, J. Disdier, P. Pichat, Chem. Phys. Lett. 108 (1984) 618–622. [472] C. DiValentin, G. Pacchioni, A. Selloni, Chem. Mater. 17 (2005) 6656–6665.
[419] P. Pichat, J.-M. Hermann, J. Disdier, M. Mozzanega, H. Courbon, Stud. Surf. Sci. [473] H. Irie, S. Washizuka, K. Hashimoto, Thin Solid Films 510 (2006) 21–25.
Catal. 19 (1984) 319. [474] S.-W. Hsu, T.-S. Yang, T.-K. Chen, M.-S. Wong, Thin Solid Films 515 (2007)
[420] N. Serpone, D. Lawless, J. Disdier, J.-M. Herrmann, Langmuir 10 (1994) 3521–3526.
643–652. [475] T. Umebayashi, T. Yamaki, S. Tanaka, K. Asai, Chem. Lett. 32 (2003) 330–331.
[421] M. Anpo, Y. Ichihashi, M. Takeuchi, H. Yamashita, Res. Chem. Intermed. 24 [476] T. Umebayashi, T. Yamaki, H. Itoh, K. Asai, Appl. Phys. Lett. 81 (2002) 454–456.
(1998) 143–149. [477] T. Umebayashi, T. Yamaki, S. Yamamoto, A. Miyashita, S. Tanaka, T. Sumita, K.
[422] S. Ikeda, N. Sugiyama, B. Pal, G. Marcí, L. Palmisano, H. Noguchi, K. Uosaki, B. Asai, J. Appl. Phys. 93 (2003) 5156–5160.
Ohtani, Phys. Chem. Chem. Phys. 3 (2001) 267–273. [478] T. Ohno, T. Mitsui, M. Matsumura, Chem. Lett. 32 (2003) 364–365.
[423] M. Anpo, M. Takeuchi, J. Catal. 216 (2003) 505–516. [479] T. Ohno, M. Akiyoshi, T. Umebayashi, K. Asai, T. Mitsui, M. Matsumura, Appl.
[424] S. Karvinen, P. Hirva, T.A. Pakkanen, J. Molec. Struct.: THEOCHEM 626 (2003) Catal. A: Gen. 265 (2004) 115–121.
271–277. [480] K. Takeshita, A. Yamakata, T.-A. Ishibashi, H. Onishi, K. Nishijima, T. Ohno,
[425] J. Zhu, Z. Deng, F. Chen, J. Zhang, H. Chen, M. Anpo, J. Huang, L. Zhang, Appl. J. Photochem. Photobiol. A: Chem. 177 (2006) 269–275.
Catal. B: Environ. 62 (2006) 329–335. [481] T. Tachikawa, S. Tojo, K. Kawai, M. Endo, M. Fujitsuka, T. Ohno, K. Nishijima,
[426] H. Tada, T. Mitsui, T. Kiyonaga, T. Akita, K. Tanaka, Nature Mater. 5 (2006) Z. Miyamoto, T. Majima, J. Phys. Chem. B 108 (2004) 19299–19306.
782–786. [482] W. Zhao, W. Ma, C. Chen, J. Zhao, Z. Shuai, J. Am. Chem. Soc. 126 (2004)
[427] M. Miyauchi, A. Nakajima, K. Hashimoto, T. Watanabe, Adv. Mater. 12 (2000) 4782–4783.
1923–1927. [483] L. Lin, W. Lin, Y. Zhu, B. Zhao, Y. Xie, Chem. Lett. 34 (2005) 284–285.
[428] I. Nakamura, N. Negishi, S. Kutsuna, T. Ihara, S. Sugihara, K. Takeuchi, J. Mol. [484] J.C. Yu, J.G. Yu, W.K. Ho, Z.T. Jiang, L.Z. Zhang, Chem. Mater. 14 (2002)
Catal. A: Chem. 162 (2000) 205–212. 3808–3816.
[429] I. Nakamura, S. Sugihara, K. Takeuchi, Chem. Lett. 29 (2000) 1276–1277. [485] K. Nukumizu, J. Nunoshige, T. Takata, J.N. Kondo, M. Hara, H. Kobayashi,
[430] I. Justicia, P. Ordejón, G. Canto, J.L. Mozos, J. Fraxedas, G.A. Battiston, R.
K. Domen, Chem. Lett. 32 (2003) 196–197.
Gerbasi, A. Figueras, Adv. Mater. 14 (2002) 1399–1402.
[486] D. Li, H. Haneda, N.K. Labhsetwar, S. Hishita, N. Ohashi, Chem. Phys. Lett. 401
[431] F.M. Hossain, G.E. Murch, L. Sheppard, J. Nowotny, Solid State Ion. 178 (2007)
(2005) 579–584.
319–325.
[487] D. Li, H. Haneda, S. Hishita, N. Ohashi, Chem. Mater. 17 (2005) 2596–2602.
[432] T. Ihara, M. Miyoshi, M. Ando, S. Sugihara, Y. Iriyama, J. Mater. Sci. V36 (2001)
[488] G.K. Mor, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, Nano Lett. 5
4201–4207.
(2005) 191–195.
[433] H. Kikuchi, M. Kitano, M. Takeuchi, M. Matsuoka, M. Anpo, P.V. Kamat, J. Phys.
[489] G.K. Mor, H.E. Prakasam, O.K. Varghese, K. Shankar, C.A. Grimes, Nano Lett. 7
Chem. B 110 (2006) 5537–5541.
(2007) 2356–2364.
[434] S. Sato, Chem. Phys. Lett. 123 (1986) 126–128.
[490] O.K. Varghese, C.A. Grimes, Solar Energy Mater. Solar Cells 92 (2008)
[435] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001)
374–384.
269–271.
[491] H.M. Yates, M.G. Nolan, D.W. Sheel, M.E. Pemble, J. Photochem. Photobiol. A:
[436] T. Morikawa, R. Asahi, T. Ohwaki, K. Aoki, Y. Taga, Japan. J. Appl. Phys. Part 2
Chem. 179 (2006) 213–223.
40 (2001) L561-–L563.
[492] S.-K. Lee, A. Mills, Chem. Commun. (2003) 2366–2367.
[437] T. Ihara, M. Miyoshi, Y. Iriyama, O. Matsumoto, S. Sugihara, Appl. Catal. B:
[493] A. Mills, J. Wang, J. Photochem. Photobiol. A: Chem. 127 (1999) 123–134.
Environ. 42 (2003) 403–409.
[438] H. Irie, Y. Watanabe, K. Hashimoto, J. Phys. Chem. B 107 (2003) 5483–5486. [494] K. Vinodgopal, D.E. Wynkoop, P.V. Kamat, Environ. Sci. Technol. 30 (1996)
[439] R. Nakamura, T. Tanaka, Y. Nakato, J. Phys. Chem. B 108 (2004) 10617–10620. 1660–1666.
[440] G.R. Torres, T. Lindgren, J. Lu, C.G. Granqvist, S.E. Lindquist, J. Phys. Chem. B [495] H. Kisch, W. Macyk, Chem. Phys. Chem. 3 (2002) 399–400.
108 (2004) 5995–6003. [496] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M.
[441] M. Mrowetz, W. Balcerski, A.J. Colussi, M.R. Hoffmann, J. Phys. Chem. B 108 Shimohigoshi, T. Watanabe, Nature 388 (1997) 431–433.
(2004) 17269–17273. [497] N. Sakai, R. Wang, A. Fujishima, T. Watanabe, K. Hashimoto, Langmuir 14
[442] S. Sakthivel, M. Janczarek, H. Kisch, J. Phys. Chem. B 108 (2004) 19384–19387. (1998) 5918–5920.
[443] S. Sakthivel, H. Kisch, Chem. Phys. Chem. 4 (2003) 487–490. [498] M. Machida, K. Norimoto, T. Watanabe, K. Hashimoto, A. Fujishima, J. Mater.
[444] O. Diwald, T.L. Thompson, T. Zubkov, E.G. Goralski, S.D. Walck, J.T. Yates, J. Sci. 34 (1999) 2569–2574.
Phys. Chem. B 108 (2004) 6004–6008. [499] R. Wang, N. Sakai, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys. Chem. B
[445] J.L. Gole, J.D. Stout, C. Burda, Y. Lou, X. Chen, J. Phys. Chem. B 108 (2004) 103 (1999) 2188–2194.
1230–1240. [500] T. Watanabe, A. Nakajima, R. Wang, M. Minabe, S. Koizumi, A. Fujishima,
[446] C. Burda, Y. Lou, X. Chen, A.C.S. Samia, J. Stout, J.L. Gole, Nano Lett. 3 (2003) K. Hashimoto, Thin Solid Films 351 (1999) 260–263.
1049–1051. [501] M. Miyauchi, A. Nakajima, A. Fujishima, K. Hashimoto, T. Watanabe, Chem.
[447] C. Di Valentin, G. Pacchioni, A. Selloni, Phys. Rev. B 70 (2004) 085116–4. Mater. 12 (2000) 3–5.
[448] H. Ozaki, S. Iwamoto, M. Inoue, Chem. Lett. 34 (2005) 1082–1083. [502] A. Nakajima, S. Koizumi, T. Watanabe, K. Hashimoto, Langmuir 16 (2000)
[449] C. Di Valentin, G. Pacchioni, A. Selloni, S. Livraghi, E. Giamello, J. Phys. Chem. 7048–7050.
B 109 (2005) 11414–11419. [503] A Nakajima, S.-I. Koizumi, T. Watanabe, K. Hashimoto, J. Photochem.
[450] Z. Lin, A. Orlov, R.M. Lambert, M.C. Payne, J. Phys. Chem. B 109 (2005) Photobiol. A: Chem. 146 (2001) 129–132.
20948–20952. [504] N. Sakai, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys. Chem. B 105 (2001)
[451] Y. Nakano, T. Morikawa, T. Ohwaki, Y. Taga, Appl. Phys. Lett. 86 (2005) 3023–3026.
132104–3. [505] N. Sakai, A. Fujishima, T. Watanabe, K. Hashimoto, J. Electrochem. Soc. 148
[452] S. Livraghi, M.C. Paganini, E. Giamello, A. Selloni, C.D. Valentin, G. Pacchioni, (2001) E395–E398.
J. Am. Chem. Soc. 128 (2006) 15666–15671. [506] M. Miyauchi, N. Kieda, S. Hishita, T. Mitsuhashi, A. Nakajima, T. Watanabe,
[453] M. Kitano, K. Funatsu, M. Matsuoka, M. Ueshima, M. Anpo, J. Phys. Chem. B K. Hashimoto, Surf. Sci. 511 (2002) 401–407.
110 (2006) 25266–25272. [507] N. Stevens, C.I. Priest, R. Sedev, J. Ralston, Langmuir 19 (2003) 3272–3275.
[454] M. Maeda, T. Watanabe, J. Electrochem. Soc. 153 (2006) C186–C189. [508] R.-D. Sun, A. Nakajima, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys.
[455] M. Batzill, E.H. Morales, U. Diebold, Phys. Rev. Lett. 96 (2006) 026103–4. Chem. B 105 (2001) 1984–1990.
[456] A. Nambu, J. Graciani, J.A. Rodriguez, Q. Wu, E. Fujita, J.F. Sanz, J. Chem. Phys. [509] M. Miyauchi, A. Nakajima, T. Watanabe, K. Hashimoto, Chem. Mater. 14
125 (2006) 094706–8. (2002) 2812–2816.
[457] T. Tachikawa, Y. Takai, S. Tojo, M. Fujitsuka, H. Irie, K. Hashimoto, T. Majima, [510] N. Sakai, A. Fujishima, T. Watanabe, K. Hashimoto, J. Phys. Chem. B 107 (2003)
J. Phys. Chem. B 110 (2006) 13158–13165. 1028–1035.
[458] A.V. Emeline, X. Zhang, M. Jin, T. Murakami, A. Fujishima, J. Phys. Chem. B 110 [511] T. Shibata, H. Irie, K. Hashimoto, J. Phys. Chem. B 107 (2003) 10696–10698.
(2006) 7409–7413. [512] T. Horiuchi, H. Ochi, K. Kaisei, K. Ishida, K. Matsushige, Materials Research
[459] N. Serpone, J. Phys. Chem. B 110 (2006) 24287–24293. Society, Structure-Property Relationships of Oxide Surfaces and Interfaces II
[460] V.N. Kuznetsov, N. Serpone, J. Phys. Chem. B 110 (2006) 25203–25209. (MRS Proceedings Volume 751) 2002 Z3.47.1-Z3.47-4.
A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582 581

[513] D.C. Hennessy, M. Pierce, K.-C. Chang, S. Takakusagi, H. You, K. Uosaki, [562] M. Miyauchi, H. Tokudome, J. Mater. Chem. 17 (2007) 2095–2100.
Electrochim. Acta 53 (2008) 6173–6177. [563] Y. Takata, S. Hidaka, M. Masuda, T. Ito, Internat. J. Energy Res. 27 (2003)
[514] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, 111–119.
M. Shimohigoshi, T. Watanabe, Adv. Mater. 10 (1998) 135–138. [564] Y. Takata, S. Hidaka, J.M. Cao, K. Tanaka, M. Masuda, T. Ito, T. Watanabe,
[515] Z.Z. Gu, A. Fujishima, O. Sato, Appl. Phys. Lett. 85 (2004) 5067–5069. M. Shimohigoshi, Thermal Science Eng. 8 (2000) 33–41.
[516] J.M. White, J. Szanyi, M.A. Henderson, J. Phys. Chem. B 107 (2003) 9029–9033. [565] H. Irie, K. Sunada, K. Hashimoto, Electrochemistry 72 (2004) 807–812.
[517] T. Zubkov, D. Stahl, T.L. Thompson, D. Panayotov, O. Diwald, J.T. Yates, J. Phys. [566] Y. Ohko, S. Saitoh, T. Tatsuma, A. Fujishima, J. Electrochem. Soc. 148 (2001)
Chem. B 109 (2005) 15454–15462. B24–B28.
[518] X. Yan, R. Abe, T. Ohno, M. Toyofuku, B. Ohtani, Thin Solid Films 516 (2008) [567] T. Tatsuma, S. Saitoh, Y. Ohko, A. Fujishima, Chem. Mater. 13 (2001)
5872–5876. 2838–2842.
[519] A. Imanishi, H. Suzuki, K. Murakoshi, Y. Nakato, J. Phys. Chem. B 110 (2006) [568] Y. Ohko, S. Saitoh, T. Tatsuma, A. Fujishima, Electrochemistry 70 (2002)
21050–21054. 460–462.
[520] A. Imanishi, T. Okamura, N. Ohashi, R. Nakamura, Y. Nakato, J. Am. Chem. Soc. [569] H. Park, K.Y. Kim, W. Choi, J. Phys. Chem. B 106 (2002) 4775–4781.
(2007). [570] P. Ngaotrakanwiwat, T. Tatsuma, S. Saitoh, Y. Ohko, A. Fujishima, Phys. Chem.
[521] K.-I. Katsumata, A. Nakajima, T. Shiota, N. Yoshida, T. Watanabe, Chem. Phys. 5 (2003) 3234–3237.
Y. Kameshima, K. Okada, J. Photochem. Photobiol. A: Chem. 180 (2006) [571] T. Tatsuma, S. Saitoh, P. Ngaotrakanwiwat, Y. Ohko, A. Fujishima, Langmuir
75–79. 18 (2002) 7777–7779.
[522] H. Irie, S. Washizuka, Y. Watanabe, T. Kako, K. Hashimoto, J. Electrochem. Soc. [572] Y. Takahashi, P. Ngaotrakanwiwat, T. Tatsuma, Electrochim. Acta 49 (2004)
152 (2005) E351–E356. 2025–2029.
[523] H. Honda, A. Ishizaki, R. Sama, K. Hashimoto, A. Fujishima, J. Illuminat. Eng. [573] P. Ngaotrakanwiwat, T. Tatsuma, J. Electroanalyt. Chem. 573 (2004)
Soc. Winter (1998) 1998. 263–269.
[524] L. Cassar, MRS Bulletin 29 (2004) 328–331. [574] P. Ngaotrakanwiwat, S. Saitoh, Y. Ohko, T. Tatsuma, A. Fujishima, J.
[525] X.T. Zhang, O. Sato, M. Taguchi, Y. Einaga, T. Murakami, A. Fujishima, Chem. Electrochem. Soc. 150 (2003) A1405–A1407.
Mater. 17 (2005) 696–700. [575] T. Bessho, K. Inoue, K. Ishikawa, I. Koiwa, H. Honma, J. Surface Finishing Soc.
[526] X. Zhang, A. Fujishima, M. Jin, A.V. Emeline, T. Murakami, J. Phys. Chem. B 110 Japan. 57 (2006) 157.
(2006) 25142–25148. [576] M. Sugimoto, K. Tashiro, T. Bessho, I. Koiwa, H. Honma, J. Surface Finishing
[527] A. Bozzi, T. Yuranova, I. Guasaquillo, D. Laub, J. Kiwi, J. Photochem. Photobiol. Soc. Japan. 57 (2006) 162.
A: Chem. 174 (2005) 156–164. [577] J.P. Lee, M.M. Sung, J. Am. Chem. Soc. 126 (2004) 28–29.
[528] A. Bozzi, T. Yuranova, J. Kiwi, J. Photochem. Photobiol. A: Chem. 172 (2005) [578] W. Kubo, T. Tatsuma, Appl. Surf. Sci. 243 (2005) 125–128.
27–34. [579] H. Notsu, W. Kubo, I. Shitanda, T. Tatsuma, J. Mater. Chem. (2005) 1523–1527.
[529] J.M. Herrmann, Top. Catalysis 34 (2005) 49–65. [580] N. Blondiaux, S. Zurcher, M. Liley, N.D. Spencer, Langmuir 23 (2007)
[530] T.N. Rao, A. Fujishima, D.A. Tryk, in: A.J. Bard, M. Stratmann, S. Licht 3489–3494.
(Eds.), Encyclopedia of Electrochemistry, in: Semiconductor Electrodes and [581] Z.-Z. Gu, A. Fujishima, O. Sato, Angew. Chem. Internat. Ed. 114 (2002)
Photoelectrochemistry, vol. 6, Weinheim, 2002. 2171–2174.
[531] I. Oller, W. Gernjak, M.I. Maldonado, L.A. Pérez-Estrada, J.A. Sánchez-Pérez, [582] R.S. Dibbell, G.R. Soja, R.M. Hoth, D.F. Watson, Langmuir 23 (2007)
S. Malato, J. Hazard. Mater. 138 (2006) 507–517. 3432–3439.
[532] M.H. Pérez, G. Peñuela, M.I. Maldonado, O. Malato, P. Fernández-Ibáñez, [583] K. Tadanaga, J. Morinaga, A. Matsuda, T. Minami, Chem. Mater. 12 (2000)
I. Oller, W. Gernjak, S. Malato, Appl. Catal. B: Environ. 64 (2006) 272–281. 590–592.
[533] M. Kositzi, I. Poulios, S. Malato, J. Caceres, A. Campos, Water Res. 38 (2004) [584] X.T. Zhang, O. Sato, A. Fujishima, Langmuir 20 (2004) 6065–6067.
1147–1154. [585] X. Zhang, M. Jin, Z. Liu, D.A. Tryk, S. Nishimoto, T. Murakami, A. Fujishima,
[534] I. Oller, W. Gernjak, M. Maldonado, P. Fernández-Ibáñez, J. Blanco, J. Sánchez- J. Phys. Chem. C 111 (2007) 14521–14529.
Pérez, S. Malato, Environ. Chem. Lett. 3 (2005) 118–121. [586] X. Zhang, H. Kono, Z. Liu, S. Nishimoto, D.A. Tryk, T. Murakami, H. Sakai,
[535] M.I. Maldonado, P.C. Passarinho, I. Oller, W. Gernjak, P. Fernández, J. Blanco, M. Abe, A. Fujishima, Chem. Commun. (2007) 4949–4951.
S. Malato, J. Photochem. Photobiol. A: Chem. 185 (2007) 354–363. [587] K. Kobayashi, Y. Tomita, Y. Maeda, Japan. J. Appl. Phys. 44 (2005) 5829–5831.
[536] C. Guillard, J. Disdier, C. Monnet, J. Dussaud, S. Malato, J. Blanco, [588] Y. Ohko, T. Tatsuma, T. Fujii, K. Naoi, C. Niwa, Y. Kubota, A. Fujishima, Nature
M.I. Maldonado, J.-M. Herrmann, Appl. Catal. B: Environ. 46 (2003) 319–332. Mater. 2 (2003) 29–31.
[537] Y. Ohko, I. Ando, C. Niwa, T. Tatsuma, T. Yamamura, T. Nakashima, Y. Kubota, [589] K. Kawahara, K. Suzuki, Y. Ohko, T. Tatsuma, Phys. Chem. Chem. Phys. 7 (2005)
A. Fujishima, Environ. Sci. Technol. 35 (2001) 2365–2368. 3851–3855.
[538] Y. Ohko, K.-i. Iuchi, C. Niwa, T. Tatsuma, T. Nakashima, T. Iguchi, Y. Kubota, [590] R. Gorges, S. Meyer, G. Kreisel, J. Photochem. Photobiol. A: Chem. 167 (2004)
A. Fujishima, Environ. Sci. Technol. 36 (2002) 4175–4181. 95–99.
[539] T. Nakashima, Y. Ohko, D.A. Tryk, A. Fujishima, J. Photochem. Photobiol. A- [591] K. Naoi, Y. Ohko, T. Tatsuma, J. Am. Chem. Soc. 126 (2004) 3664–3668.
Chem. 151 (2002) 207–212. [592] K. Naoi, Y. Ohko, T. Tatsuma, Chem. Commun. (2005) 1288–1290.
[540] T. Nakashima, Y. Ohko, Y. Kubota, A. Fujishima, J. Photochem. Photobiol. A: [593] S. Teekateerawej, J. Nishino, Y. Nosaka, J. Photochem. Photobiol. A: Chem. 179
Chem. 160 (2003) 115–120. (2006) 263–268.
[541] T. Saito, T. Iwase, J. Horie, T. Morioka, J. Photochem. Photobiol. B: Biol. 14 [594] G. Takei, T. Kitamori, H.-B. Kim, Catal. Commun. 6 (2005) 357–360.
(1992) 369–379. [595] E.T. Castellana, S. Kataoka, F. Albertorio, Cremer, Anal. Chem. 78 (2006)
[542] Z. Huang, P.-C. Maness, D.M. Blake, E.J. Wolfrum, S.L. Smolinski, W.A. Jacoby, 107–112.
J. Photochem. Photobiol. A: Chem. 130 (2000) 163–172. [596] G. Takei, M. Nonogi, A. Hibara, T. Kitamori, H.-B. Kim, Lab Chip 7 (2007)
[543] A.-G. Rincón, C. Pulgarin, Appl. Catal. B: Environ. 51 (2004) 283–302. 596–602.
[544] P. Fernández, J. Blanco, C. Sichel, S. Malato, Catal. Today 101 (2005) 345–352. [597] H. Nagai, T. Irie, J. Takahashi, S.-i. Wakida, Biosens. Bioelectr. 22 (2007)
[545] E.F. Duffy, et al., Solar Energy 77 (2004) 649–655. 1968–1973.
[546] M.I. Litter, Appl. Catal. B: Environ. 23 (1999) 89–114. [598] E. Baur, A. Rebmann, Helv. Chim. Acta 4 (1921) 256–262.
[547] M.R. Prairie, L.R. Evens, B.M. Stange, S.L. Martinez, Environ. Sci. Technol. 27 [599] J.M. White, J. Szanyi, M.A. Henderson, J. Phys. Chem. B 108 (2004) 3592–3602.
(1993) 1776–1782. [600] J.M. White, M.A. Henderson, J. Phys. Chem. B 109 (2005) 12417–12430.
[548] M.A. Ferguson, M.R. Hoffmann, J.G. Hering, Environ. Sci. Technol. 39 (2005) [601] M.A. Henderson, J.M. White, H. Uetsuka, H. Onishi, J. Catal. 238 (2006)
1880–1886. 153–164.
[549] A.-G. Rincón, C. Pulgarin, Solar Energy 77 (2004) 635–648. [602] H. Uetsuka, H. Onishi, M.A. Henderson, J.M. White, J. Phys. Chem. B 108 (2004)
[550] P. Pichat, J. Disdier, C. Hoang-Van, D. Mas, G. Goutailler, C. Gaysse, Catal. Today 10621–10624.
63 (2000) 363–369. [603] H. Uetsuka, M.A. Henderson, A. Sasahara, H. Onishi, J. Phys. Chem. B 108
[551] C.H. Ao, S.C. Lee, Appl. Catal. B: Environ. 44 (2003) 191–205. (2004) 13706–13710.
[552] C.H. Ao, S.C. Lee, J. Photochem. Photobiol. A: Chem. 161 (2004) 131–140. [604] M.A. Henderson, J. Phys. Chem. B 108 (2004) 18932–18941.
[553] J.-H. Kim, G. Seo, D.-L. Cho, B.-C. Choi, J.-B. Kim, H.-J. Park, M.-W. Kim, [605] M.A. Henderson, J. Phys. Chem. B 109 (2005) 12062–12070.
S.-J. Song, G.-J. Kim, S. Kato, Catal. Today 111 (2006) 271–274. [606] C.L. Perkins, M.A. Henderson, J. Phys. Chem. B 105 (2001) 3856–3863.
[554] S.A. Grinshpun, A. Adhikari, T. Honda, K.Y. Kim, M. Toivola, K.S. Ramchander- [607] M.A. Henderson, J. Phys. Chem. B 101 (1997) 221–229.
Rao, T. Reponen, Environ. Sci. Technol. 41 (2007) 606–612. [608] M.A. Henderson, Surf. Sci. 319 (1994) 315–328.
[555] T. Ibusuki, K. Takeuchi, J. Mol. Catal. 88 (1994) 93–102. [609] M. Murai, Y. Tamaki, A. Furube, K. Hara, R. Katoh, Catal. Today 120 (2007)
[556] H. Ichiura, T. Kitaoka, H. Tanaka, Chemosphere 51 (2003) 855–860. 214–219.
[557] Italcementi, http://www.italcementigroup.com/ENG/Research+and+ [610] H. Tada, M. Tanaka, Langmuir 13 (1997) 360–364.
Innovation/Innovative+Products/Innovative+Products.htm, 2008. [611] H. Wang, T. Wang, P. Xu, J. Mater. Sci.: Mater. Electr. 9 (1998) 327–330.
[558] Y. Ohko, Y. Utsumi, C. Niwa, T. Tatsuma, K. Kobayakawa, Y. Satoh, Y. Kubota, [612] A. Fernández, G. Lassaletta, V.M. Jiménez, A. Justo, A.R. González-Elipe,
A. Fujishima, J. Biomed. Mater. Res. 58 (2001) 97–101. J.M. Herrmann, H. Tahiri, Y. Ait-Ichou, Appl. Catal. B: Environ. 7 (1995) 49–63.
[559] Y. Sekiguchi, Y. Yao, Y. Ohko, K. Tanaka, T. Ishido, A. Fujishima, Y. Kubota, [613] S. Yanagida, A. Nakajima, Y. Kameshima, N. Yoshida, T. Watanabe, K. Okada,
Internat. J. Urology 14 (2007) 426–430. Mater. Res. Bull. 40 (2005) 1335–1344.
[560] A. Fujishima, D.A. Tryk, T. Watanabe, K. Hashimoto, Internat. Glass Rev. (1998) [614] Y. Gao, Y. Masuda, Z. Peng, T. Yonezawa, K. Koumoto, J. Mater. Chem. 13
114–116. (2003) 608–613.
[561] Y. Takata, S. Hidaka, J.M. Cao, T. Nakamura, H. Yamamoto, M. Masuda, T. Ito, [615] A. Goossens, E.-L. Maloney, J. Schoonman, Chem. Vapor Deposition 4 (1998)
Energy 30 (2005) 209–220. 109–114.
582 A. Fujishima et al. / Surface Science Reports 63 (2008) 515–582

[616] Z. Ding, X. Hu, P.L. Yue, G.Q. Lu, P.F. Greenfield, Catal. Today 68 (2001) [627] C. Lee, H. Choi, C. Lee, H. Kim, Surf. Coatings Technol. 173 (2003) 192–200.
173–182. [628] G.K. Mor, O.K. Varghese, M. Paulose, K. Shankar, C.A. Grimes, Solar Energy
[617] S.A. O’Neill, I.P. Parkin, R.J.H. Clark, A. Mills, N. Elliott, J. Mater. Chem. 13 Mater. Solar Cells 90 (2006) 2011–2075.
(2003) 56–60. [629] J.M. Macak, M. Zlamal, J. Krysa, P. Schmuki, Small 3 (2007) 300–304.
[618] P. Zeman, S. Takabayashi, J. Vac. Sci. Technol. A 20 (2002) 388–393. [630] K. Takami, T. Adachi, A. Nakajima, K. Hashimoto, T. Watanabe, Surface
[619] P.K. Song, M. Yamagishi, H. Odaka, Y. Shigesato, Japan. J. Appl. Phys. 42 (2003) Coatings Internat. B: Coatings Trans. 85 (2002) 155–158.
L1529–L1531. [631] K. Kobayakawa, C. Sato, Y. Sato, A. Fujishima, J. Photochem. Photobiol. A:
[620] M. Kitano, M. Takeuchi, M. Matsuoka, J.M. Thomas, M. Anpo, Chem. Lett. 34
Chem. 118 (1998) 65–69.
(2005) 616–617.
[632] M. Marezio, A. Gauzzi, F. Licci, E. Gilioli, Physica C: Superconductivity 338
[621] P.K. Song, Y. Irie, Y. Shigesato, Thin Solid Films 496 (2006) 121–125.
(2000) 1–8.
[622] M. Takeuchi, T. Yamasaki, K. Tsujimaru, M. Anpo, Chem. Lett. 35 (2006)
904–905. [633] T. Shibata, N. Sakai, K. Fukuda, Y. Ebina, T. Sasaki, Phys. Chem. Chem. Phys. 9
[623] M.-C. Yang, T.-S. Yang, M.-S. Wong, Thin Solid Films 469–470 (2004) 1–5. (2007) 2413–2420.
[624] T.-S. Yang, C.-B. Shiu, M.-S. Wong, Surf. Sci. 548 (2004) 75–82. [634] T. Sasaki, J. Ceram. Soc. Japan. 115 (2007) 9–16.
[625] T. Miyata, S. Tsukada, T. Minami, Thin Solid Films 496 (2006) 136–140. [635] C. Morterra, J. Chem. Soc., Faraday Trans. 1 (84) (1988) 1617–1637.
[626] N. Berger-Keller, G. Bertrand, C. Filiatre, C. Meunier, C. Coddet, Surf. Coatings [636] Y. Suda, T. Morimoto, Langmuir 3 (1987) 786–788.
Technol. 168 (2003) 281–290. [637] P. Jones, J.A. Hockey, Trans. Faraday Soc. 67 (1971) 2669–2678.

Вам также может понравиться