Вы находитесь на странице: 1из 32

Encyclopedia of

Nanoscience and
Nanotechnology

www.aspbs.com/enn

Nanocrystalline TiO2 for Photocatalysis


Hubert Gnaser, Bernd Huber, Christiane Ziegler
Universität Kaiserslautern, Kaiserslautern, Germany

CONTENTS indicated by the present theme, has stimulated great hopes


in accomplishing thereby achievements with paramount ben-
1. Introduction efits for human beings and the global environment. To out-
2. Electronic and Charge-Transfer Processes line the present state of that quest is the major goal of this
in Photocatalysis article.
3. Preparation of Nanostructured Materials “Catalysis” is probably the most familiar of the three
and Thin Films terms mentioned. A catalyst is incorporated in essentially
everybody’s automobile, with the goal of reducing or even
4. Structural Properties of Nanocrystalline
eliminating the engine’s toxic gaseous components by con-
TiO2 Films verting them into less harmful (albeit not necessarily benign)
5. Electrical Properties of Nanocrystalline substances. As is the case in all catalytic reactions, the cat-
TiO2 Films alyst itself is not part of the reaction, but is expected to
6. Photocatalytic Properties enhance its rate, that is, the velocity of the transformation
of Nanocrystalline TiO2 from the original components (the “educts” in the chemist’s
terminology) into the final ones (the “products”). Hence, a
7. Photocatalytic Applications catalyst is an entity that accelerates a chemical reaction with-
of Nanocrystalline TiO2 out being consumed itself in the process. Without catalysts,
Glossary various chemical reactions of great importance would pro-
References ceed too slowly [11]. The economic significance of catalysis
is enormous. In the U.S. alone, the annual value of products
manufactured with the use of catalysts is roughly in the vicin-
1. INTRODUCTION ity of one trillion dollars [12]. Indeed, more than 80% of the
The development of novel materials and the assessment of industrial chemical processes in use nowadays rely on one or
their potential application constitutes a major fraction of more catalytic reactions [13]. A number of those, including
today’s scientific reasearch efforts. In fact, there exist var- oil refining, petrochemical processing, and the manufactur-
ious major governmental research and development pro- ing of commodity chemicals (olefins, methanol, ethylene gly-
grams related to nanostructured materials. Furthermore, it col, etc.), are already well established. But many others, as
is estimated that nanotechnology has grown into a multi- will be seen in this contribution, represent challenges requir-
billion dollar industry and may become the most domi- ing the development of entirely new approaches. But apart
nant single technology of the twenty-first century. To allow from their industrial importance, catalytic phenomena effect
for this fact, this encyclopedia [1] encompasses a series virtually all aspects of our lives. They are crucial in many
of contributions devoted to a very prominent field of cur- processes occurring in living things, where enzymes are the
rent materials research activities, namely, nanoscience and catalysts. They are important in the processing of foods and
nanotechnology. The importance of these developments is the production of medicines. The reader may have noticed
reflected also in a number of recent books and articles that we have as yet refrained from specifying the meaning
reviewing this rapidly evolving field [2–10]. of photocatalysis; which will be one of the major topics of
This article focuses on a specific class of such novel this article. This term refers to a catalytic process that is
nano-scaled materials, nanocrystalline TiO2 , and its photo- triggered by illuminating the system by visible light or ultra-
catalytic properties. The title of this article encompasses violet irradiation. Ideally, that light flux would be the sun’s
three main terms (“(photo)catalysis,” “nanocrystalline,” and radiance.
“TiO2 ”) which, individually, stand for very important areas Next we shall consider the meaning of “nanocrystalline.”
of scientific research and of, perhaps even more important, First, it is noted that in today’s science world rather
technological applications. Their synergistic combination, as inflationary used, the prefix “nano” refers to a fraction of

ISBN: 1-58883-062-4/$35.00 Encyclopedia of Nanoscience and Nanotechnology


Copyright © 2004 by American Scientific Publishers Edited by H. S. Nalwa
All rights of reproduction in any form reserved. Volume 6: Pages (505–535)
506 Nanocrystalline TiO2 for Photocatalysis

one part in one billion (109  and, hence, its correct usage been elucidated through several decades of intense scientific
would require it being connected to some kind of unit (e.g., research (for a review see, e.g., [19]); some of them will be
of length, time, energy, mass, etc.). In the present context referred to in the course of the present overview. The fea-
(and in that of “nanotechnology”), “nano” most often relates ture probably most important in the present context is the
to the dimension, that is, the size of an object. Therefore, fact that TiO2 is a semiconductor with a bandgap of ∼ 3.2 eV.
nanocrystalline in the ensuing discussions will designate par- On the other hand, TiO2 , in its nanocrystalline form, consti-
ticles (of crystalline structure and, primarily, with the chem- tutes an enormously important commercial product. In fact,
ical composition of titanium dioxide) whose typical sizes are the world production of titanium dioxide white pigments
in the range of a few to several nanometers (nm), that is, of amounts to some 4.5 million tons per annum and the global
the order of the one billionth part of one meter. Obviously, consumption may be considered a distinct economic indi-
these are extremely tiny objects and can be “seen” and stud- cator. White pigments of TiO2 have average particle sizes
ied only with the help of sophisticated analytical instruments of around 200–300 nm, optimized for the scatter of white
like an electron microscope. light, resulting, thereby, in a hiding power. Reducing the
At first glance, it may appear that such tiny particles are crystallite size (to ≤ 100 nm), the reflectance of visible light
a rather modern contrivance, but this is probably a prema- (vis) decreases and the material becomes more transparent;
ture conclusion. In fact, it is quite firmly established that it is widely employed, for example, in paints, plastics, paper,
nm-sized particles (mostly very refractory ones like corun- or pharmaceuticals. Nanocrystalline TiO2 exhibits, in addi-
dum, diamond, or silicon carbide) are ubiquitous in the uni- tion, a pronounced absorption of ultraviolet (UV) radiation.
verse [14] and that they were already present at the time Because of this high UV absorption and the concurrent high
and the location of the formation of the solar system. This transparency for visible light, TiO2 particles with a size of
“stardust” originated from stellar outflows and supernova <100 nm have found widespread use in such diverse areas
ejecta, which may have occurred eons before the gas and as sun cosmetics, packaging materials, or wood protection
dust condensed into what is now the sun, the earth, and the coatings. Hence, although perhaps not generally realized,
planets. In fact, this dust has intrigued astronomers since TiO2 is ubiquitous in our everyday life. Apart from this well-
the days of William Herschel who noted, in the 1780’s, the established range of usage, an increasing number of catalytic
existence of small regions in the sky where there appeared applications of nanocrystalline TiO2 have emerged in recent
to be a complete absence of stars [15]. These regions are years.
most easily seen against the rich star-fields of the Milky Way. At this point, it appears appropriate to return to a more
Evidence of the presolar origin of these nanocrystalline par- general view of the present topic. Nanostructured materials
ticles comes primarily from their isotopic abundance pat- have generally the potential [3] for incorporating and taking
tern [16], which deviates typically to such an extent from advantage of a number of size-related effects in condensed
any other known matter that a terrestrial or solar origin matter ranging from electronic effects (so-called “quantum
is virtually impossible. (Most of these particles that have size effects”) caused by spatial confinement of delocalized
been investigated were extracted from primitive meteorites valence electrons and altered cooperative (“many-body”)
in which they were incorporated during the formation stage atom phenomena, like lattice vibrations or melting, to the
of the solar system; these did not experience any later mod- suppression of such lattice-defect mechanisms as dislocation
ification and, hence, preserved the presolar dust particles generation and migration in confined grain sizes. The pos-
unaltered [17].) sibilities to assemble size-selected atom clusters into new
Only now, some billions of years later, mankind has ini- materials with unique or improved properties will likely cre-
tiated the manufacture and application of such nanocrys- ate a revolution in our ability to engineer materials with
talline materials. Nanostructured materials with crystal sizes controlled optical, electronic, magnetic, mechanical, and
in the range of 5–50 nm of a variety of materials, includ- chemical properties for many pending future technological
ing metals and ceramics, have been artificially synthesized by applications.
many different techniques in the past couple of years [2, 3, Among those nanocrystalline materials, semiconductors
5–7]. Such new ultrafine-grained materials have properties appear to play a pivotal role in such distinct fields as [20]:
that are often significantly different and greatly enhanced
(i) heterogeneous photocatalysis;
as compared to coarser-grained or bulk substances. These
(ii) photoelectrochemistry, including electrochemical
favorable changes in properties result generally from their
photovoltaic cells;
small grain sizes, the large percentage of atoms in grain
(iii) photochemistry in zeolites, intercalated materials,
boundaries and at surfaces, the large surface-to-bulk ratio,
thin films, and membranes (like self-assembled struc-
and the interaction between individual crystallites. Since
tures);
these features can be tailored to a considerable extent, dur-
(iv) supramolecular photochemistry.
ing synthesis and processing, such nanophase materials are
thought to have great technological potential even beyond This diversity is thought to be largely due to the fact that
their current applications. heterogeneously dispersed semiconductor surfaces provide
Let us finally turn to a brief discussion of the third term, both a fixed environment to influence the chemical reactivity
“TiO2 ” ( i.e., titanium dioxide). TiO2 has three different crys- of a wide range of adsorbates and, in addition, a means to
tal structures [18]: rutile, anatase, and brookite; only the for- initiate light-induced redox reactivity in these weakly associ-
mer two of them are commonly used in photocatalysis. Like ated molecules [21]. Upon photoexcitation of semiconductor
for many other metal oxides (also for titanium oxide) have nanoparticles, either in solutions or fixed to a suitable sub-
the respective structural, optical, and electronic properties strate, simultaneous oxidation and reduction reactions may
Nanocrystalline TiO2 for Photocatalysis 507

occur; molecular oxygen is often assumed to serve as the oxi- following scheme
dizing agent. Such semiconductor-mediated redox reactions
are commonly grouped under the rubric of heterogeneous
photocatalysis [21].
TiO2 + 2h → 2e− + 2h+ electron-hole pair
In a heterogeneous photocatalytic system, photo-induced
formation in TiO2
molecular reactions take place at the surface of the cat-
alyst. Depending on where the initial excitation occurs, H2 O + 2h+ → 1/2O2 + 2H+ reaction at the TiO2
photocatalysis can be generally divided into two classes of electrode
processes [22]: In the case that the initial photoexcitation 2H+ + 2e− → H2 reaction at the
occurs in the adsorbate molecule, which then interacts with Pt electrode
the ground state catalyst substrate, the process is referred
to as a catalyzed photoreaction. On the other hand, when H2 O + 2h → 1/2O2 + H2 overall reaction
the initial excitation takes place in the catalyst substrate and
the excited catalyst transfers an electron or energy into a
ground state molecule, this process is referred to as a sensi- It appears to be generally accepted that this discovery
tized photoreaction [22]. Apparently, a considerable degree boosted extensive research efforts in the era of heteroge-
of synergism may be crucial when, for example, a transition
neous photocatalysis [21, 25–27]. These studies, often car-
metal oxide photocatalysts is combined with supramolecular
ried out in an interdisciplinary fashion with the participation
spectral sensitizing ligand complexes used to harvest light
of physicists, chemists, and chemical engineers, focused on
and vectorially transfer photo-generated electrons and holes
issues that are of great relevance both economically as well
along selected energetic pathways.
as ecologically like energy renewal and storage [28–30], the
In 1972, Fujishima and Honda reported [23] the photocat-
alytic splitting (i.e., the simultaneous oxidizing and reducing) decomposition of organic compounds in polluted air and
of water upon illumination of a TiO2 single-crystal electrode; wastewaters [31–33], chemical energy generation [34, 35],
in analogy to natural photosynthesis, they demonstrated the and photovoltaic devices [36, 37]. Most of these either
photoelectrolysis of water making use of a photoexcited already proven or envisaged applications are intimately
semiconductor in what was essentially a photochemical bat- linked to the extraordinary properties of nanocrystalline
tery. In that system, an n-type TiO2 semiconductor elec- TiO2 . In fact, nanocrystalline metal-oxide semiconductors
trode, which was connected through an electrical load to a such as TiO2 have been applied successfully in modern
platinum counter electrode, was exposed to near-UV light technologies including solar energy conversion, gas sensors,
(cf. Fig. 1). When the surface of the TiO2 electrode was catalysis, and photocatalysis [38–42].
illuminated with light of wavelength shorter than ∼415 nm, Following the first examination in 1977, using TiO2 to
photocurrent was observed to flow [23, 24]. Furthermore, decompose cyanide in water [43, 44], a great deal of effort
oxygen evolution (i.e., an oxidation reaction) on the TiO2 has been devoted in recent years to developing heteroge-
and hydrogen evolution (a reduction) on the Pt electrode neous photocatalysts with high photocatalytic activities for
was observed. The photoexcitation of TiO2 injects electrons their applications in solving environmental cleanup and pol-
from its valence band into its conduction band; the electrons lution remediation problems [31, 32, 45, 46]. Photocatalytic
flow through the external circuit to the Pt cathode where reactions on TiO2 surfaces are very important in such envi-
water molecules are reduced to hydrogen gas and the holes ronmental processes, as the oxidation of organic materials
remain in the TiO2 anode where water molecules are oxi- and the reduction of heavy metal ions in industrial waste
dized to oxygen. These observations indicate that water can waters. Apart from the utilization for water and air purifi-
be decomposed by means of UV-VIS light according to the cation, TiO2 photocatalysis has been shown useful for the
destruction of microorganisms such as bacteria [47] and
viruses [48], for the inactivation of cancer cells [46, 49], the
e- load e- clean-up of oil spills [50, 51], and other applications [45]; a
more detailed account will be given later in this article.
e- e- As mentioned, TiO2 is a semiconductor with a bandgap
of 3.2 eV; when excited by light of energy equal to or
CB - hν exceeding that value, electrons are promoted from the
e- Pt valence band to the conduction band leaving positive holes
H2O H2O
VB + in the valence band. These electrons and holes are capa-
TiO2 ble of, respectively, reducing and oxidizing compounds at
H2 O2 the TiO2 surface. If the electrons and holes do not recom-
bine (and produce heat), they can follow various reaction
Figure 1. Schematic arrangement for the photosplitting of water in an pathways; it is commonly accepted that the hole is quickly
electrochemical cell (in the actual setup, both electrodes are immersed
converted to the hydroxyl radical (• OH) upon oxidation of
in an aqueous solution and the chambers are separated by an ionically
conducting porous material). When the TiO2 photoanode is irradiated surface-adsorbed water and that the hydroxyl radical is the
with light, O2 evolves from it, whereas H2 evolves from the Pt counter- major reactant responsible for the oxidation of organic com-
electrode, while electrons will flow through an external load. pounds. Typical reductive and oxidative reactions could be
508 Nanocrystalline TiO2 for Photocatalysis

the following [25, 52–54] approach clearly has room for further improvements, it may
ultimately face limitations in terms of cost efficiency (manu-
TiO2 + h → TiO2 e− + h+  electron-hole facturing costs per unit of energy produced). Novel materials
pair formation and fabrication schemes are therefore explored. A promis-
e− + Mn+ → Mn−1+ reduction reaction ing approach consists of electrochemical photovoltaic cells
utilizing nanoporous semiconducting electrodes fabricated
+ • +
h + H2 Oads → OH + H oxidation of by lightly sintering nanosized TiO2 particulates, followed by
adsorbed water spectral sensitization with an appropriate dye.

OH + Rads → • Rads + H2 O oxidation of Metal oxide particles with diameters of some 10 nm
organic species can be prepared as paste and spread out over a surface
of fluorine-doped SnO2 conducting glass to form a three-
where Mn+ is the oxidized compound and Rads the dimensional network of interconnected nanoparticles. These
adsorbed organic species. Because the production of the nanostructured metal oxide layers, and in particular those
hydroxyl radical is considered the decisive step, the determi- constructed from anatase titanium dioxide (TiO2 , have
nation and optimization of the corresponding quantum yield aroused great interest because of their unprecedented prop-
in illuminated TiO2 is an important task [55, 56]. erties as electrodes. They find application in dye-sensitized
According to those concepts, TiO2 nanoparticles are solar cells, which nowadays show light-to-electricity conver-
expected to show a unique surface chemistry due to their sion efficiencies of 10% [69, 73–77]. The nature of electron
larger surface area [57]. The origin of the distinct photo- migration in these electrodes has been debated in past years
catalytic activities exhibited by nanoparticles of TiO2 is cru- as the experimental results and their interpretation does not
cial in understanding the reaction mechanisms, for example, converge to a generally accepted model. Instead, the exact
if they are purely due to the increased surface area or if role of electron trapping and the concomitant screening of
they have to be addressed to the presence of distorted sites the electric field remain unclear. It is reported that soon
on the surface or to the whole lattice of the particles. In after electrons are injected into the conduction band of TiO2
order to commercialize these treatment techniques, it is of a large fraction of them get trapped in surface states. Migra-
great importance to improve the preparative methods of tion of these electrons must then proceed with a hopping-
titania, because the photocatalytic activity and phase tran- type process in which the electrons remain most of the time
sition behavior of TiO2 are significantly influenced by the in localized states [78–83]. On the other hand, it is also
preparative conditions [58–63]. As previously mentioned, reported that the injection of electrons into the conduction
these catalytic processes constitute a major issue of this work band shows the so-called “free-electron” absorption, which
and will be outlined in the following sections. extends over a wide spectral range from the visible to the
The photocatalytic activity of TiO2 is very useful not infrared [84–88]. This suggests that not all injected elec-
only in environmental purification by decomposition of trons become trapped but that a substantial fraction of them
organic substances, but also in the materials industry such as remain in the conduction band.
mirrors and glasses due to its self-cleaning [64] and antifog- This article is organized in the following way: Section 2
ging effects. The latter has been attributed to the photo- outlines the electronic and charge-transfer processes as rel-
induced hydrophilic nature of the surface [65, 66]. A further
evant for photocatalysis, with a special emphasis towards
enhancement of the photocatalytic activity could be effected
nanocrystalline TiO2 . Since in photocatalysis and related
by means of composite TiO2 materials; examples would be
applications the respective nanostructured materials are
metal doping, mixing with insulating substances like SiO2 or
employed either in colloidal solutions or attached to a suit-
Al2 O3 [67], and monolayer coverage by SiO2 [68].
able support (e.g., as electrodes or thin films), both of these
Another prominent future application of nanocrystalline
semiconductors is thought [69–71] to lie in photovoltaics, preparation techniques are discussed (Section 3). Further-
that is, the conversion of sunlight into electrical power. more, various approaches for surface and thin-film modifi-
The limited reserves of fossil fuels and the increasing con- cation are described, as well as novel deposition methods
cern of global warming due to the greenhouse effect caused and structures. The structural and electronic properties of
by the combustion of those fuels has triggered, in the last nanocrystalline TiO2 films are examined in Sections 4 and
decades, widespread efforts into the development of photo- 5, respectively. The photocatalytic properties of nanocrys-
voltaic devices. With an energy supply from the sun to the talline TiO2 constitute the central theme of Section 6, high-
earth of 3 × 1024 Joule per year (about 10,000 times the lighting the dependence of the photocatalytic activity on
global annual energy consumption), this enterprise appears different parameters like film structure and phase, surface
all but unreasonable. In such solar cells, photon incident on morphology, electronic properties, and the effects induced
a semiconductor can create electron-hole pairs, basically a by various surface modifications. Representative examples
result of the photoelectric effect, discovered by Becquerel of photocatalytic applications utilizing nanocrystalline TiO2
already in 1839 [72]. At a junction between two different materials are presented in Section 7. Finally, an extensive
materials, this may establish an electrical potential differ- set of references is provided that should be useful for fur-
ence across this interface. Until now, the material of choice ther study: although the number is substantial, no attempt
for manufacturing such junctions has been (doped) silicon was made, however, to be comprehensive; any such attempt
(crystalline or amorphous), with compound semiconductors might be bound to fail due to the rapidity with which this
also being considered more recently. While this traditional field is evolving.
Nanocrystalline TiO2 for Photocatalysis 509

2. ELECTRONIC hν
AND CHARGE-TRANSFER
- CB hν
PROCESSES IN PHOTOCATALYSIS
2.1. Electronic Excitations + VB
a - +
and Charge-Carrier Trapping b d
-
+
A photocatalytic process is initiated by the absorption of + c - +
+
D+
+
photons by a molecule or the substrate to produce highly
reactive electronically excited states. The efficiency is con-
-
trolled by the system’s light absorption properties. Three A-
fundamental steps are of relevance: (1) the electronic excita- D
tion of a molecule upon photon absorption, (2) the band-gap
excitation of the semiconductor substrate, and (3) the inter-
facial electron transfer. Since a detailed account of these A
processes has been given in a lucid treatise by Linsebigler
Figure 2. Schematic illustration of the photoexcitation in a semicon-
et al. [22], we shall briefly summarize here only the more
ductor particle followed by deexcitation events. CB and VB designate
important points, referring thereby partly to that work. The the conduction and valence band, respectively. For further details see
probability of an electronic transition can be calculated from text.
quantum mechanical perturbation theory and is propor-
tional to the square of the amplitude of the radiation field oxidizing the donor (pathway d). Competing with charge
and the square of the transition dipole moment [89, 90]. transfer to adsorbed species is electron and hole recombi-
The latter may be computed via the molecular wave func- nation, occurring either in the volume (pathway b) or on
tion which, in turn, depends on the product of the electronic the surface of the semiconductor (pathway a; in both cases
spatial wave function, the electronic spin wave function, and heat will be released. Naturally, electron and hole recom-
the nuclear wave function. Further arguments [89] lead to bination is detrimental to the efficiency of a semiconductor
some general selection rules in terms of which electronic photocatalyst. Modifications to semiconductor surfaces, such
transitions are allowed and which might be forbidden. Typ- as metal addition, dopants, or combination with materials,
ically, the excitation of a weak transition will not effectively can be beneficial in decreasing the recombination and con-
induce a photochemical reaction, because few of the inci- currently increasing the quantum yield of the process.
dent photons will be absorbed (low cross-section); however, An efficient means to retard the recombination of pho-
an overall high reaction rate may still be possible in the case toexcited electron-hole pairs may proceed via the trapping
of a high quantum yield, that is, if the probability of prod- of charge carriers. The occurrence of surface and bulk irreg-
uct molecule formation per absorbed photon is high. The ularities resulting from the preparation process is associated
photochemical efficiency will also be determined by which with surface electron states; these may serve as charge car-
deexcitation channels are dominant. In particular, the perti- rier traps and can suppress the recombination of electrons
nent lifetimes of the involved processes are to be considered. and holes. The charge carriers trapped in such states are
Whereas the absorption of a photon occurs very rapidly on localized to a specific site on the surface or in the bulk; their
the order of 10−15 s, deexcitation events are much slower, population is dependent on the energy difference between
favoring the decay channel which minimizes the lifetime of the trap and the bottom of the conduction band. Experi-
the excited state. mental observations of such trapped states in TiO2 will be
The initial process for heterogeneous photocatalysis of reported later.
organic and inorganic compounds by semiconductors is the On the basis of laser flash photolysis measurements
generation of electron-hole pairs in the semiconductor par- [45, 92], the characteristic times for the individual steps
ticles. Once excitation across the bandgap has occurred, occurring during heterogeneous photocatalysis on TiO2 have
the lifetime is sufficient (in the nanosecond regime [91]) been derived. Whereas the primary process of charge-carrier
for the created electron-hole pair to undergo charge trans- generation is extremely fast (∼fs), charge-carrier recombina-
fer to adsorbed species on the semiconductor surface from tion may occur on time scales of 10–100 ns. Charge-carrier
solution or gas phase. Figure 2 exemplifies these processes. trapping can be very fast (100 ps) for the (reversible) trap-
The enlarged section shows the excitation of an electron ping of a conduction-band electron in a shallow trap, but
from the valence band to the conduction band initiated moderately fast (∼10 ns) for a deep trap or for the surface
by light absorption with energy equal or greater than the trapping of a valence-band hole, resulting in a surface-bound
bandgap of the semiconductor. The figure also illustrates hydroxyl radical. Finally, interfacial charge transfer can be
several deexcitation pathways for the electrons and holes. slow (∼100 ns) for the oxidation of an electron donor or
The electron transfer to adsorbed species or to the solvent very slow (∼ms) for the reduction of an electron acceptor.
results from migration of electrons or holes to the sur- In general, the valence-band holes are powerful oxidants
face. At the surface, the semiconductor can donate elec- (+10 to +35 V versus NHE depending on the type of semi-
trons to reduce an electron acceptor (often oxygen in an conductor and pH), while the conduction-band electrons are
aerated solution), corresponding to pathway c in Figure 2; good reductants (+0.5 to −1.5 V vs NHE) [93]; most organic
conversely, a hole can migrate to the surface where an elec- photodegradation reactions utilize the oxidizing power of
tron from a donor species can combine with the surface hole these photo-generated holes.
510 Nanocrystalline TiO2 for Photocatalysis

2.2. Band-Edge Position semiconductor electrolyte


and Band Bending + - + - - + - + - +
- + -
+ + + +
- + - - + + + - - +
The ability of a semiconductor to undergo photo-induced
electron transfer to adsorbed species on its surface is gov- E
erned by the band energy position of the semiconductor and CB CB E
the redox potentials of the adsorbate. The relevant potential Ec Ec
of the acceptor must be below (more positive than) the con- EF Eref EF
duction band potential of the semiconductor. By contrast, Eref
the potential of the donor needs to be above (more nega- EV EV
tive than) the valence-band position of the semiconductor in VB VB
order to donate an electron to the vacant hole. The band-
edge positions of several semiconductors are depicted in (a) (b)
Figure 3; the internal energy scale is given both with respect
to the vacuum level (left scale) and for comparison to nor- + - - + + - - + + + - -
+ - + + + -
mal hydrogen electrode (NHE) in a solution of an aqueous - - + + - - + + + - -
electrolyte at pH = 1.
When a semiconductor is brought into contact with E
another phase (e.g., liquid, gas, or metal), the transfer E
of mobile charges across this junction occurs until elec-
Eref
tronic equilibrium is reached. This redistribution of charges CB CB
Ec Eref Ec
involves the formation of a space-charge layer, that is, the
charge distribution on each side of the interface differs EF EF
from the bulk material (cf. Fig. 4). Depending on whether
EV EV
the electrons accumulate or deplete at the semiconductor
side, an accumulation or depletion layer is formed, causing VB VB
concurrently a shift in the electrostatic potential and a down- (c) (d)
ward (or upward) bending of the bands in the semiconduc-
tor. The depletion of electrons may reach such an extent that Figure 4. Space-charge layer of an n-type semiconductor in contact
their concentration at the surface decreases below the intrin- with another phase (e.g., an electrolyte or gas): (a) flat band situation,
sic level. As a consequence, the Fermi level is closer to the (b) accumulation layer, (c) depletion layer, and (d) inversion layer.
valence than to the conduction band; this situation is called
an inversion layer, as the semiconductor is p-type at the
when an n-type semiconductor like TiO2 is in contact with
surface and n-type in the bulk. These features are well docu-
mented [94] and will not be further discussed here. Of some an electrolyte as in a photoelectrochemical cell [71]; such
interest in the present context is the situation encountered devices are thought to have great potential both in pho-
tovoltaics for producing electric current and as fuel cells
for the generation of hydrogen via photo-cleavage of water.
E [eV] ENHE [eV] Because of these potential applications, the characteristics
0.0 vacuum of the semiconductor-electrolyte interface have been inves-
tigated in great details [93, 95, 96].
In particular, the potential distribution within a spherical
-2.5 -2.0
SiC semiconductor particle could be derived [97] using a lin-
-3.0 GaAsP GaP (n,p) -1.5
(n,p) (n,p) earized Poisson–Boltzmann equation. As discussed in [69],
-3.5 GaAs CdS -1.0 two limiting cases are of special importance for light-induced
(n)
-4.0
(n,p) CdSe ZnO TiO2 -0.5 Eu2+/3+
electron transfer in semiconductor dispersions. For large
(n) (n) (n)
WO3 SnO2
H2/H+
particles, the total potential drop within the particle is
-4.5 (n) (n) 0.0
-5.0 0.5 [Fe(CN)6]3-/4-  2
kT w
-5.5 1.0
Fe2+/Fe3+
 = (1)
∆E= Ru(bipy)2+/3+ 2e LD
1.4 2.25 2.25 Ce4+/3+
-6.0 eV eV eV 1.7 1.5
eV
3.0
eV
where w is the width of the space charge layer and LD =
-6.5 2.0
2.5
eV
0 kT /e2 Nd 05 is the Debye screening length [94], which
-7.0 2.5
depends on the dielctric constant, , of the material and on
-7.5
3.2
3.0 the number density of ionized dopants, Nd . This potential
3.2
-8.0
eV
3.2 eV 3.5 variation is identical with that of a planar Schottky deple-
eV tion layer [98]. For very small semiconductor particles (with
-8.5 4.0
3.8 radius R) the potential drop within the particle becomes
eV
 2
kT R
Figure 3. Bandgap energies of various semiconductors in an aqueous  = (2)
electrolyte at pH = 1. 6e LD
Nanocrystalline TiO2 for Photocatalysis 511

From the latter equation, it is found that the electrical field the empty acceptor orbital and, simultaneously, an electron
in nano-sized semiconductors will usually be small and that is donated from the filled donor level to recombine with the
high dopant levels are required to produce a significant original hole. This is a very general process for wide bandgap
potential difference between the center and the surface. For oxide semiconductors like TiO2 and others. For example,
example [69], in order to obtain a 50 meV potential drop in a the oxidation of many organic substances in colloidal sus-
nanocrystalline TiO2 particle with R = 6 nm, a concentration pensions has been investigated [102]. The energetics of the
of 5 × 1019 cm−3 of ionized donor impurities is necessary. semiconductor valence band and the oxidation potential of
Undoped TiO2 nanocrystallites have a much lower carrier the redox couple influence this photocatalytic oxidation. For
concentration and the band bending within the particles is example, the enhancement in the efficiency of halide oxida-
therefore negligibly small. tion at TiO2 follows the sequence Cl− < Br− < I− , corre-
lating with the decrease in the oxidation potential. Again,
2.3. Photo-Induced Charge-Transfer some recent examples will be presented in Section 6.
The kinetic analysis [103] of electron transfer in colloidal
Processes on the Catalyst Surface semiconductor systems is often complex. Apart from the
The principle of electron and hole transfer at the photoex- energetics of the conduction band of the semiconductor
cited semiconductor particle has already been alluded to and the redox potential of the acceptor, factors such as
previously in Fig. 2. Both free and trapped charge carri- the surface charges of the colloids, adsorption of the sub-
ers participate in these interfacial redox reactions. Due to strates, participation of surface states, and competition with
the quantization effects, by decreasing the particle’s size, it charge recombination influence the rate of charge transfer at
is possible to shift the conduction band to more negative the semiconductor interface [102]. This fact is evident from
potentials and the valence band to more positive values. It the widely differing rates of experimentally observed charge
was concluded [99] that the shift of the bandgap is propor- transfer rates, with time scales ranging from picoseconds to
tional to 1/R2 , R being the particle size. Therefore, redox milliseconds for different experimental conditions and vari-
processes that cannot occur in bulk materials can be facili- ous semiconductor systems.
tated in quantized semiconductor particles. Figure 5 shows
schematically such possible transfer reactions for an adsor- 2.4. Quantum-Size Effects
bate at the surface of a semiconductor. When there are
accessible energy levels in the substrate, an electron may Size quantization effects in metals or semiconductors
be transferred from the donor (D) into a substrate level have attracted considerable attention in the past decade
and then into the acceptor orbital as shown in Figure 5(a). [104–107]. Semiconductor nanoparticles may experience a
This scheme operates in the photosensitization of semicon- transition in terms of electronic properties from those typ-
ductor particles by dye molecules. An electron is injected ical for a solid to that of a molecule, in which a com-
from the excited state dye molecule into the semiconduc- plete electron delocalization has not yet occurred. These
tor, which then reduces another adsorbate particle. Early quantum-size effects arise when the Bohr radius of the first
experimental confirmation [100, 101] of these processes used exciton (an interacting electron-hole pair) and the semicon-
the reduction of methyl viologen in colloidal semiconductor ductor becomes comparable with or larger than that of the
systems and the water splitting process. Later, such reduc- particle; the Bohr radius [94]
tive processes have been investigated for many different sys- rB = 40  2 /e2 m∗  (3)
tems (see, e.g., [102]); some illustrative examples will be
presented in Section 6. depends on the dielectric constant  and the effective mass
For an initial excitation on the semiconductor substrate m∗ of the charge carriers (electrons and holes). The lat-
(Fig. 5(b)), a positively charged hole is created at the band- ter is frequently radically different for holes and electrons
edge of the valence band. An electron is transferred into and, in some cases, m∗ is more than an order of magni-
tude smaller than the free-electron mass me . Hence, such
quantum-size effects play a role for crystallites of approx-

CB CB
imately 1–10 nm in diameter. Under these conditions, the
CB
energy levels available for the electrons and holes in the con-
D D* D+
- duction and valence bands become discrete and the effec-
A A A- tive bandgap of the semiconductor is increased. To a first
VB VB VB approximation, the energy spacing between quantized lev-
els is inversely proportional to the effective mass and the
(a)
square of the particle diameter. A schematic energy diagram
CB - CB - e- resulting from such confinement effects is shown in Figure 6.
CB
Several attempts have been carried out to compute the
A A A-
hν electronic energy levels in such quantum dots [99, 108–110].
D D D+ According to these concepts, the energy of the lowest excited
VB + VB + e -
state of a semiconductor particle with radius R is given
VB

(b) approximately by
 
2 2 1 1 18e2
Figure 5. Sensitized photoreaction with (a) an initial excitation of the ER = Eg + + − (4)
adsorbate, or (b) an initial excitation of the solid. 2R2 m∗e m∗h R
512 Nanocrystalline TiO2 for Photocatalysis

bulk semiconductor cluster In colloidal solution, semiconductor particles reduce the


light intensity of the incident beam both by scattering and
LUMO absorption. In the absence of quantum-size effects the extinc-
CB
tion spectrum is described by the Mie theory [113, 114].
shallow This theoretical approach can be applied to an assembly of
Eg trap deep trap spherical particles if the interparticle distance is larger than
deep trap
surface
state
the wavelength of the incident light (i.e., the particles scat-
ter independently); if, furthermore, the particle size is much
VB smaller than the wavelength, the energy-dependent absorp-
HOMO tion cross-section   for the irradiation of a solution con-
taining the particles can be derived [115–117]:
Figure 6. Schematic correlation diagram relating bulk-semiconductor
electronic states to quantum crystalline states. Adapted from [110], s3/2 2
M. G. Bawendi et al., Annu. Rev. Phys. Chem. 41, 477 (1990). © 1990,   = 9Vp (8)
c 1 + 2s 2 + 22
Annual Reviews.
where   = 1 + i2 is the complex, frequency-dependent
Here Eg is the bandgap of the bulk semiconductor, the sec- dielectric constant of the semiconductor particle, s is the
ond term is the quantum energy of localization, increasing dielectric constant of the embedding medium (the solvent),
as R−2 for both electron and hole, and the third term is the Vp is the volume of a particle, is the frequency of the
Coulomb attraction [99]; whereas the Coulomb term shifts incident light, and c is the velocity of light. For the case
ER to smaller energy as R, the quantum confinement of a dilute system of particles with the number density n,
contribution increases ER as R2 . As a result, the appar-   can be related to the reciprocal absorption length
ent bandgap will always increase for small enough R. But
 
  = n  . It follows that the imaginary part
while the shift can be appreciable (∼1 eV) for small band of the dielectric constant is a direct measure of the light
gap materials like InSb, it might be considerably smaller absorption by the particles; it increases steeply near the
(∼0.2 eV) for semiconductors with a large bandgap like absorption edge, that is, for close to the threshold fre-
TiO2 or ZnO [99]. Apart from a large effect on the opti- quency. As noted in [69], Mie’s theory has been widely
cal properties (e.g., a change in color of the material due employed to interpret the extinction spectra of colloidal sys-
to the blue shift of the absorption), size quantization may tems [118]. For example, the brilliant ruby or yellow colors
also lead to major changes in the photocatalytic properties. caused, respectively, by colloidal gold or silver particles are
While these effects have been studied in great detail for sev- distinctly explained by this theory.
eral compound semiconductor materials like CdS, ZnO, or
PbS, related data for TiO2 appear to be still rather limited.
3. PREPARATION OF
2.5. Optical Properties NANOSTRUCTURED MATERIALS
Semiconductors absorb light below a threshold wavelength
AND THIN FILMS
g which is related to the bandgap energy Eg by [93] Several distinct techniques have been utilized in recent years
to synthesize nanocrystalline TiO2 thin films by chemical,
g nm = 1240/Eg eV (5) electrochemical, and organized assembly methods [69, 119–
Within the solid, the extinction of light follows an exponen- 121]. A simple approach involves casting of the thin film
tial law directly from colloidal suspensions [122]. This method of
preparation is relatively simple and inexpensive compared
I = I0 exp−
z (6) with other existing methods such as chemical vapor deposi-
tion or molecular beam epitaxy. Preparation of nanoclusters
where z is the penetration depth and
is the reciprocal in polymer films and Langmuir–Blodgett films has also been
absorption length. For TiO2 , for example,
= 26 × 104 cm−1 considered. The sol–gel technique has been found to be use-
at a wavelength of 320 nm; this implies that such light is ful in developing nanostructured semiconductor membranes
extinguished to 90% after passing a distance of 390 nm in with either a two-dimensional or three-dimensional con-
the semiconductor. Near threshold, the value of
increases figuration. Organic-template-mediated synthesis has been
with increasing photon energy. Frequently, a proportionality employed to develop nanoporous materials. The nanostruc-
of the type tured materials are highly porous and can easily be surface

h = Ch − Eg n (7) modified with sensitizers, redox couples, and/or other nano-
structured films.
provides a satisfactory description of the absorption where
h is the photon energy. C is a constant scaling with the 3.1. Preparation from
effective masses of the charge carriers, but is independent
Colloidal Suspensions
of the photon frequency. The exponent n has a value of
0.5 for a direct semiconductor and 2 for an indirect one Nanostructured semiconductor films of TiO2 have been pre-
[94]. Experimental data [111, 112] obtained on both anatase pared frequently from colloidal suspensions [123–133]. By
and rutile TiO2 thin films indicated, however, that the actual controlling the preparative conditions of the precursor col-
situation might be more complex. loids, it is possible to tailor the properties of these films.
Nanocrystalline TiO2 for Photocatalysis 513

Typically, these thin TiO2 films exhibit interesting pho- particle growth rate [142–144]. In a recent study [145],
tochromic, electrochromic, photocatalytic, and photoelec- anatase nanoparticles were prepared in a Ti-peroxy gel with-
trochemical properties that are inherited from the native out the collapse of the gel during the particle formation
colloids. The synthetic procedure involves the preparation process. The gel was made from titanium tetraisopropoxide
of ultrasmall particles (diameter 2–10 nm) in aqueous or (Ti(O iPr)4  and H2 O2 .
ethanolic solutions by controlled hydrolysis. The colloidal The crystallization rate of titania gels was found to be
suspension is coated onto a conducting glass plate (an opti- much higher in water than in methanol or n-hexane [146];
cally transparent electrode (OTE)) and dried; finally, the also, the crystallite size of anatase prepared in water was
film is annealed at 200–400  C in air for some 1–2 h. larger than those in organic solvents. Processing parameters
The conducting surface facilitates direct electrical contact of very often control the crystallite size and phase. Nonag-
the nanostructured thin film. This simple approach of coat- glomerated, ultrafine anatase particles have been gener-
ing preformed colloids onto a surface and annealing gener- ated by hydrothermally treating sol–gel-derived hydrous
ally produces an oxide film, which is robust and exhibits an oxides [147]. The degree of crystallinity and purity of the
excellent stability in both acidic and alkaline media, a fea- synthesized materials may affect their structural evolution
ture very important in several potential applications. Gener- during any heat treatment.
ally, some optimization is required for thicker films and for TiO2 thin films with different surface structures were pre-
higher colloidal concentrations in order to avoid cracking pared from alkoxide solutions containing polyethylene glycol
of the films upon drying. Further details of the methodol- (PEG) via the sol–gel method [148]. The larger the amount
ogy of preparation can be found in [134–136]. Transmis- of PEG added to the precursor solution, the larger the size
sion electron micrographs of nanostructured films prepared and number of pores produced in the resultant films. When
from colloidal suspensions show a three-dimensional net- PEG was added to the gel, the films decomposed completely
work of oxide nanocrystals of particle diameter as small during heat treatment. The adsorbed hydroxyl content of
as a few nanometers. No significant aggregation or sinter- such porous thin films is found to increase with increasing
ing effects are observed during the annealing process. X-ray amount of PEG. However, the transmittance of the films
diffraction analysis also confirms the crystallinity of these decreases due to the scattering of light by pores of larger
nanostructured films. Composite films, which in some cases size and a higher number in the films. Photocatalytic degra-
may exhibit improved properties as compared to single- dation experiments show that methyl orange is efficiently
component films, can be manufactured by mixing two or decolorized in the presence of the TiO2 thin films by expos-
more components prior to casting of the film. ing its aqueous solution to ultraviolet light. However, in films
Titania sol and gel prepared through the hydrolysis of deposited on soda-lime glass [149], diffusion of sodium and
tetrabutyl titanate in acid water solution are sensitive to calcium ions from the glass into the nascent TiO2 films was
ultraviolet (UV) irradiation and turn into blue color [137]. found to be detrimental to the photocatalytic activity of the
Electron spin-resonance measurement showed that the pho- resulting films. Sodium and calcium diffusion into nascent
tochromism was ascribed to the presence of titanium (III) TiO2 films was effectively retarded by a 03 m SiO2 inter-
ions in the Ti-O-Ti network. The titanium (III) ions could facial layer formed on the soda-lime glass [149, 150].
be oxidized by the oxygen in the atmosphere, and then the TiO2 thin film photocatalysts coated onto glass plates were
sol and gel faded slowly. The absorption peaks in the optical prepared [151] by thermal decomposition of tetraisopropyl
absorption spectra of the titania gel were attributed to the orthotitanate with a dip-coating process using alpha-
transition of 3-dimensional electrons of a trivalent titanium terpineol as a highly viscous solvent. Two types of ligands—
in certain environments. polyethylene glycol 600 and (ethoxyethoxy)ethanol—were
The morphology of TiO2 particles affects their catalytic added to the dip-coating solution as the stabilizer of tita-
activity and electrical properties. In recent years, many nium alkoxide and thin films were obtained after calcination
methods for preparing TiO2 nanoparticles and thin films at 450  C for 1 h. The film thickness obtained with a single
have been studied [138]. TiO2 nanocrystals prepared by the dipping was proportional to the viscosity of the dip-coating
sol–gel method often have fully hydroxylated surfaces and solutions. The thin films obtained were transparent with a
these hydroxyl groups have a strong influence on the cat- thickness of 1 m. The crystal form of the films was anatase
alytic and photocatalytic properties such as electron-transfer alone. The thin films were formed with aggregated nano-
rates and reducing properties [139]. In order to develop sized TiO2 single crystals (7–15 nm), and the size of the TiO2
photocatalysts with high activities, it is very important to crystals became smaller for the polymer-added systems.
prepare porous anatase nanoparticles with a high specific Transparent anatase TiO2 -based multilayered photocat-
surface area. Furthermore, the preparation method should alytic films synthesized via a sol–gel process on porous
be simple and should be a low temperature process. Mix- alumina and glass substrates showed a sponge-like micro-
tures of rutile and anatase precipitates could be obtained structure and a mean crystallite dimension of ca. 8 nm [152].
by the hydrothermal treatment of an alcohol solution of Ti Doping such films with iron(III) impeded the photocatalytic
alkoxide at 573 K [140], while anatase nanoparticles were activity.
prepared by heat treatment of a H2 O–EtOH solution of The effects of calcination on the microstructures of nano-
TiOSO4 at 373 K [141]. Thus, the anatase and rutile particles sized titanium dioxide powders prepared by vapor hydroly-
were usually formed in a solution by conventional soft chem- sis was investigated in detail [153–155]: Among the factors
ical synthesis methods. The preparation of monodispersed examined [153], large surface area and good dispersion
oxide particles by the “sol–gel method” enables the manu- of the powders in the reaction mixture are favorable to
facture of oxide particles through a gel state with a regulated photoactivity. Conversely, prolonged calcination at high
514 Nanocrystalline TiO2 for Photocatalysis

temperatures is detrimental to photoactivity. Powders pro- nucleated must be removed from the region of high super-
duced at higher temperatures are predominantly anatase saturation. Since the aggregates are already entrained in the
and are generally more photoactive. condensing gas, this is readily accomplished by setting up
The formation, structure, and photophysical properties conditions for moving this gas.
of functional mixed films of 5,10,15,20-tetra-4-(2-decanoic Typically, there are three fundamental rates that essen-
acid)phenyl porphyrin (TDPP) with TiO2 nanoparticles tially control the formation of the clusters in the gas-
formed from the two-dimensional sol–gel process of tetra- condensation process [160]. These are
butoxyltitanium (TBT) at the air/water interface was reported
(i) the rate of supply of atoms to the region of supersat-
[156]. The composite multilayer films were assembled by
uration where condensation occurs,
transferring the mixed monolayer onto quartz plates. The
(ii) the rate of energy removal from the hot atoms via
sensitization of TDPP upon TiO2 nanoparticles was con-
the condensing medium, the gas,
firmed by the spectral changes of UV-visible absorption and
(iii) the rate of removal of the cluster upon nucleation
fluorescence of TDPP in the composite films. Furthermore
from the supersaturated region.
the photosensitization greatly affected the photocatalytic
activity of TiO2 particles with respect to the degradation of The clusters that are collected via thermophoresis on the
methylene blue. surface of a cold finger usually form very open, fractal struc-
Crystalline titania thin films were obtained [157] on glass tures. The clusters are held on the collector surface rather
and various kinds of organic substrates at 40–70  C by deposi- weakly, via Van-der-Waals forces, and are easily removed by
tion from aqueous solutions of titanium tetrafluoride. Trans- means of a scraper. The material removed is consolidated
parent films consisting of small anatase particles (∼20 nm) in a compaction unit at typical pressures of 1–2 GPa; the
exhibited excellent adhesion to relatively hydrophilic sur- scraping and consolidation are carried out under ultra-high-
faces. Uniform coatings were successfully prepared on sub- vacuum conditions in-situ after the removal of the gas from
strates with complex shapes such as cotton and felt fiber. the chamber, in order to maximize the cleanliness of the
Growth rate and particle size were controlled by both the particle surfaces and the interfaces that are formed.
deposition conditions and the addition of an organic surfac-
tant. Organic dyes were incorporated into the anatase films 3.4. Film Deposition by Sputtering
using organic-dye dissolving solutions and a surfactant.
and Vacuum-Based Techniques
Crystalline titanium dioxide films are often deposited by
3.2. Chemical Precipitation
various techniques employing vacuum conditions, using,
Rutile-phase nano-sized TiO2 powders having a high spe- for example, RF magnetron [161–166], DC sputtering
cific surface area of 180 m2 /g were prepared by homoge- [167–171], chemical vapor deposition [172, 173], plasma
neous precipitation at ambient or very low temperatures spraying [174, 175], or related techniques [176, 177]. Rapid
(<100  C) [158]. Ultrafine SnO2 TiO2 -coupled particles electroplating of photocatalytically highly active TiO2 -Zn
could be synthesized [159] by homogeneous precipitation; nanocomposite films on steel was achieved [178] and the
they were employed for photocatalytic degradation of azo gas-phase oxidation of CH3 CHO was employed as an indi-
dye active red X-3B in aerated solution. The results show cator of the photocatalytic activity. Not surprisingly, the film
that a very rapid and complete decolorization of the azo and surface morphologies and the crystallinity are strongly
dye can be achieved, and the photoactivity of the coupled dependent on the total and the oxygen partial pressure, the
particles is higher than that of pure ultrafine TiO2 , and the deposition rate, and the phase composition; the resulting
optimum loading of SnO2 on TiO2 is 18.4%. The enhanced photocatalytic properties can be modified over a wide range
degradation rate of X-3B using coupled photocatalysts is by those parameters. Such films may show good unifor-
attributed to increased charge separation in these systems. mity of thickness over large areas, high optical transmittance
(∼80%) in the visible region, and considerable mechanical
durability [169].
3.3. Gas Condensation and Consolidation
Transmission electron microscopy was used to study [179]
Another method to synthesize nanostructured materials is the structure, morphology, and orientation of thin TiO2 films
by way of gas condensation followed by the in-situ consol- prepared by reactive magnetron sputtering on glass slides
idation under high-vacuum conditions [2]. This approach at different substrate temperatures (100 to 400  C). The
can produce ultrafine-grained materials which may exhibit microstructure and photocatalytic reactivity of TiO2 films
size-related effects. The basic aspects of the generation of have been shown to be functions of deposition tempera-
nanometer-sized materials via gas condensation are concep- ture. In the temperature range examined, all film samples
tually rather simple [2]: A precursor material, either an ele- have a porous nanostructure and the dimension of parti-
ment or a compound, is evaporated in a gas maintained at a cles grew with increasing deposition temperature. Films are
low pressure, usually well below one atmosphere. The evap- amorphous at temperatures of 100  C and only the anatase
orated atoms or molecules lose energy via collisions with phase forms at 200  C and above. Films deposited between
the gas atoms and undergo a homogeneous condensation 200 to 300  C show a preferred orientation, while films at
to form atomic or molecular clusters in the highly supersat- 400  C change into complete random orientation. Deposi-
urated vicinity of the precursor source. In order to main- tion at 250  C yields high efficiency in photocatalytic degra-
tain small cluster sizes, by minimizing further atom/molecule dation owing to the high degree of preferred orientation and
accretion and cluster-cluster coalescence, the clusters once nanocrystalline/nanoporous anatase phase.
Nanocrystalline TiO2 for Photocatalysis 515

Another frequently employed and convenient way for the the films experienced a rising phase, followed by a dramatic
preparation of TiO2 thin films is pulsed laser deposition drop with an increasing number of Au particles. Ultrafine
[180], although this technique does not produce nanocrys- Pt particles [191] were embedded into rutile TiO2 particles
talline structures. by decomposing a colloidal organic-Pt complex, resulting in
very narrow size distribution with a mean diameter of 3 nm.
These nm-sized Pt particles were found to grow epitaxially
3.5. Surface and Film Modifications
on the TiO2 crystallites with a well-defined crystallographic
The surface of as-deposited films has frequently been mod- relationship.
ified in the quest to enhance the catalytic activity. For Doping a nano-structured TiO2 electrode sensitized with
example, Fe [181] or Sn [182] ions have been implanted tetrasulfonated gallium phthalocyanine with tetrasulfonated
into transparent and colorless TiO2 thin films fabricated on zinc porphyrin (ZnTsPP) greatly enhances the photoelec-
microscope glass slides by DC magnetron reactive sputter- tric conversion at long wavelengths, with 20- and 60-
ing using Ar and O2 as working gases. The efficiency of fold improvement of the quantum efficiency at 680 and
this procedure could not as yet be proven unambiguously. 700 nm [195].
Apart from grain size, the presence of reactive species on Semiconductor/metal composite nanoparticles have been
the surface [183] may influence the photocatalytic activity. synthesized [196] by chemically reducing HAuCl4 on the sur-
Implanted metal ions (V or Cr) were observed [184] located face of preformed TiO2 nanoparticles. These gold-capped
at the lattice positions of Ti4+ in TiO2 and were stabilized TiO2 particles (diameter 10–40 nm) were stable in acidic
after calcination of the samples in an O2 ambient at around (pH 2–4) aqueous solutions. The role of the gold layer
775 K. Spectroscopic studies showed that the presence of in promoting the photocatalytic charge transfer has been
these substitutional metal ion species are, in fact, responsible probed using thiocyanate oxidation at the semiconductor
for a large shift in the absorption spectra of these catalysts interface. More than 40% enhancement in the oxidation effi-
toward visible light regions. Porous anatase TiO2 films were ciency was found with TiO2 /Au nanoparticles capped with
densified by Zn+ ion implantation up to the ion penetra- low concentration of the noble metal.
tion depth [185]. After the subsequent annealing at 800  C, Magnetic photocatalysts were synthesized by coating tita-
the phase transformation from anatase to rutile accompa- nium dioxide particles onto colloidal magnetite and nano-
nied with grain growth up to the film thickness was observed. magnetite particles [197]. The photoactivity of the prepared
In addition, the phase transformation was not induced by coated particles was lower than that of single-phase TiO2
the annealing up to 800  C with or without preceding Ar+ and was found to decrease with an increase in the heat
ion implantation. Thus, the implanted impurity Zn assisted treatment. These observations were explained in terms of
the phase transformation. Annealing in O2 tends to reduce an unfavorable heterojunction between the titanium dioxide
the rate of phase transformation and create ZnTiO3 . Optical and the iron oxide core, leading to an increase in electron-
absorption above the photon energy of 2.9 eV was increased hole recombination.
remarkably by the Zn+ or Zn+ and O+ ion implantation and TiO2 -based powders, doped with a small amount of
subsequent annealing and is due to the phase transformation. SiO2 , were prepared by a sol–gel method and were sub-
The presence of active species such as Ti3+ and hydoxyl sequently heated to precipitate fine anatase crystals [198].
on the surface of ultrafine TiO2 particles, prepared by a col- The obtained powders have large specific surface areas
loidal chemical approach and subjected to different heating (∼200 m2 /g) and upon treating them chemically with aque-
treatments, was inferred [186] from optical absorption and ous NaOH, the photocatalytic property of the powders was
photoelectron spectra; these species may enhance the pho- extremely improved and the powders showed higher activity
tocatalytic activity of particles. Treating TiO2 powder by a than the undoped TiO2 powders.
hydrogen plasma resulted in a reduction of the oxide par- In composite TiO2 -SiO2 thin films prepared by a sol–gel
ticles and electrons trapped at the oxygen-defect sites were process, the refractive index and the photocatalytic activity
found [187]. Also laser ablation of TiO2 photocatalysts aim- decrease with increasing SiO2 content in these films [199].
ing at the enhancement of the activity was reported [188]. Alkoxide sol–gel processing has been investigated for the
Noble metal particles of Au, Pt, and Ir were deposited synthesis of stable SiO2 -TiO2 high-permeability catalytic
on nanostructured TiO2 films using an electrophoretic membranes to be used in alkene isomerization [200].
approach [189]. The improved photoelectrochemical per- Nanocrystalline TiO2 was prepared on mesoporous sil-
formance of the semiconductor-metal composite film was ica both by sol–gel processes [201] and by an impregna-
attributed to the shift in the quasi-Fermi level of the com- tion method with titanium complexes featuring different
posite to more negative potentials. Continuous irradiation ligands [202].
of the composite films over a long period causes the pho- Binary mixed oxide of Fe/Ti (1:1 composition) with homo-
tocurrent to decrease as the semiconductor-metal interface geneous distribution of iron into the TiO2 has been prepared
undergoes chemical changes. by sol–gel impregnation using metal alkoxide precursors and
Doping of nanostructured TiO2 both as particles and in firing at different temperatures (500, 700, and 900  C) [203].
thin films with a variety of metals has been reported, some The mixed oxide exhibits excellent absorption in the visible
common examples being Pt [190, 191], Pb [192], Au [193], spectral region (570–600 nm). The photocatalytic activity of
or others [194]. A shift of the UV-vis absorption towards the Fe/Ti oxide reduces to a large extent at a high sintering
longer wavelengths was observed upon Pb doping, which temperature of the sample due to the presence of a increas-
indicates a decrease in the bandgap of TiO2 . In TiO2 films ing amount of the inactive (Fe2 /TiO5  pseudobrookite phase.
embedding Au nanoparticles [193], the specific resistance of Nanostructured TiO2 /SnO2 binary oxides were prepared by
516 Nanocrystalline TiO2 for Photocatalysis

combustion of stearic acid precursors [204], with metal pre- with anatase and brookite phases has been developed by
cursors being dispersed in the stearic acid at the molecular hydrolysis of titanium tetraisopropoxide in pure water or
level. It was found that preparative methods affected the a 1 + 1 EtOH–H2 O solution under ultrasonic irradiation;
crystalline structure of the powders and the anatase phase the photocatalytic activity of TiO2 particles prepared by this
of TiO2 was stabilized by the addition of SnO2 . method exceeded that of Degussa P25 and was the first
Spray “painted” (spray deposited) titanium dioxide report that showed high photocatalytic activity of a photo-
coatings were sensitized [205] with chemically deposited catalyst containing an 80% anatase and 20% brookite phase.
cadmium selenide thin films; the structural, optical, and pho- A novel and convenient nonhydrolytic approach to the
toelectrochemical characterization of these composite films preparation of uniform, quantum confined TiO2 nanocrys-
indicate the importance of thermal treatments in improving tals, using an intramolecular adduct stabilized alkoxide
the photocurrent quantum yields. Up to 400  C, the effect precursor, was described recently [220]. In contrast to estab-
of air annealing is to shift the onset of absorption to longer lished aqueous sol–gel-techniques, the processing in hydro-
wavelengths and improve the photocurrent substantially. carbon solvents at high temperatures allows access to very
Organic compounds may play a crucial role in chemical small free-standing crystallites, and opens up new possibili-
processes for ceramic coatings [206]. Organic compounds ties for control over size distribution, surface chemistry, and
remained in a fixed position in the coating, which was pre- particle agglomeration.
pared from the chemically modified titanium tetraisopropox- It has been reported that the columnar morphologies
ide solution and heated at temperatures as high as 673 K. It in sputtered TiO2 films enhances the photocatalytic [221]
was not until the organic compounds decomposed to carbon and photovoltaic [222] efficiency. Following these studies,
dioxide and the gas phase was left from the coating that the enhanced surface-reaction efficiency has been demonstrated
nanostructure, consisting of nano-sized pores and anatase in the photocatalysis of sculptured thin films of TiO2 [223].
crystallites with preferred orientation, developed at 723 K. In obliquely deposited films with variously shaped columns
Cobalt(II) 4,4 ,4 ,4 -tetrasulfophthalocyanine, covalently such as zigzag, cylinder, and helix, the columnar thickness
linked to the surface of titanium dioxide particles, TiO2 - and spacing play an important role in the enhancement of
CoTSP, was shown [207] to be an effective photocatalyst for the effective surface area, while the columnar shape is less
the oxidation of sulfur (IV) to sulfur (VI) in aqueous sus- important. The optimum morphology for a surface reaction
pensions. Upon bandgap illumination of the semiconductor, has been obtained at the deposition angle of 70 , where the
conduction-band electrons and valence-band holes are sep- photocatalytic activity is 2.5 times larger than that at 0 . The
arated; the electrons are channeled to the bound CoTSP morphology of these obliquely deposited thin films appears
complex resulting in the reduction of dioxygen, while the well suited for application as solar cells, electro- and pho-
holes react with adsorbed S(IV) to produce S(VI) in the tochromic devices, and photocatalysts.
form of sulfate. The template method for synthesizing nanostructures
The photoactivity of the Pt/TiO2 system in the visible involves the synthesis of the desired material within the
region was improved [208] by the addition of the sensitizer pores of a nanoporous membrane or other solid. This
([Ru(dcbpy)2 (dpq)]2+  [where dcbpy = 4,4 -dicarboxy 2,2 - approach has been used in several experiments [224–229] for
bipyridine and dpq = 2,3-bis-(2 -pyridyl)-quinoxaline] lead- the preparation of TiO2 nanotubes and nanorods; typically,
ing to an efficient water reduction. porous aluminum oxide (PAO) nano-templates were used.
Photocatalytic properties for hydrogen production were Compact, continuous, and uniform anatase nanotubules
investigated [209] on layered titanium compounds interca- with diameters in the range 50–70 nm were produced
lating CdS in the interlayer, which were prepared by direct inside PAO nano-templates by pressure impregnating the
cation exchange reactions and sulfurization processes. The PAO pores with titanium isopropoxide and then oxidatively
photocatalytic activity of the compounds intercalating CdS decomposing the reagent at 500  C [230]. Cleaning the sur-
was superior to those of simple CdS and the physical mix- face of the template and repeating the process several times
ture of CdS and metal oxides. The improvement might be produced titania nanotubules with a wall thickness of ∼3 nm
attributed to the formation of microheterojunctions between per impregnation. The tube exteriors appeared to be faithful
the CdS nanoparticles and the layers of oxides. replicas of the pores in which they were formed.
Nano-TiO2 whiskers were prepared by various techniques
3.6. Novel Deposition Methods [213, 231]; using, for example, controlled hydrolysis of
titanium butoxide [231] it was found that the nano-TiO2
and Structures whiskers obtained were anatase and grew selectively in the
In recent years, there has been increased interest in studying [001] direction with a diameter of about 4 nm and a length
and manufacturing nanoscaled TiO2 materials as nanoparti- of about 40 nm. Acetic acid played an important role in the
cles [210], nanowires [211], nanorods [212], whiskers [213], oriented growth of nano-TiO2 whiskers.
and nanotubes [214–217]. Highly ordered TiO2 nanowire (TN) arrays were prepared
There are many synthetic routes for the creation of [211] in anodic alumina membranes by a sol–gel method.
nanocrystals of oxides and the controlled hydrolysis of metal The TNs are single crystalline anatase phase with uniform
alkoxides is the most generalized solution-phase synthetic diameters around 60 nm. At room temperature, photolu-
strategy [218]. Increased photocatalytic activity was reported minescence measurements of the TN arrays show a visible
recently for TiO2 prepared by ultrasonic irradiation and gly- broadband with three peaks, which are located at about 425,
cothermal methods. This novel method [219] for prepar- 465, and 525 nm that are attributed to self-trapped excitons,
ing highly photoactive nanometer-sized TiO2 photocatalysts F , and F + centers, respectively.
Nanocrystalline TiO2 for Photocatalysis 517

4. STRUCTURAL PROPERTIES On the other hand, La retarded both transformation and


OF NANOCRYSTALLINE TiO2 FILMS densification.
Transmission electron microscopy (TEM) is typically used
4.1. The Lattice Structure of Rutile to investigate the crystal size distribution, grain-boundary
and Anatase disorder, and defect structure in nanocrystalline TiO2 mate-
rials prepared by the various techniques outlined in the pre-
Three different crystal structures of TiO2 exist [18]: rutile, vious section. In a recent study of films with an average grain
anatase, and brookite; only the former two of them are com- size of 15 nm prepared by reactive sputtering [238], evidence
monly used in photocatalysis, with anatase typically exhibit- of both ordered and disordered grain-boundary regions was
ing the higher photocatalytic activity [232]. The structure found and planar defects were observed in grain interiors
of rutile and anatase can be described in terms of chains identified as (011) deformation twins. Also, crystallographic
of TiO6 octahedra. The two structures differ by the dis- shear defects can occur as a result of aggregation of oxygen
tortion of each octahedron and the actual pattern of the vacancies in understoichiometric titanium oxide. Figure 8
chains. Figure 7 depicts the unit cell structures of rutile and shows results of TEM investigations [239] of nanocrystalline
anatase crystals [233–235]. Each Ti4+ ion is surrounded by an TiO2 films prepared from colloidal suspensions; the TiO2
octahedron of six O2− ions. The octahedron in rutile shows crystallites were nominally pure anatase phase with a size of
a slightly orthorhombic distortion, whereas the respective 16 nm.
octahedra in anatase are significantly distorted, resulting
in a symmetry that is lower than orthorhombic. The Ti-Ti
distances in anatase are greater (0.379 and 0.304 nm as com- 4.2. Structure of Crystalline TiO2 Surfaces
pared to 0.357 and 0.296 nm in rutile) while the Ti-O dis-
tances are shorter than in rutile (0.1934 and 0.1980 nm in The geometric structure of crystalline surfaces of TiO2 has
anatase versus 0.1949 and 0.1980 nm in rutile). In the rutile been studied predominantly on (macroscopically) large sin-
structure, each octahedron is in contact with 10 neighboring gle crystals; in particular, the (110) surface of rutile, being
ones (two sharing edge oxygen pairs and eight sharing cor- thermodynamically the most stable one [240], has been
ner oxygen atoms), whereas in the anatase crystal each octa- investigated extensively by a broad variety of surface science
hedron is in contact with eight neighbors (four sharing an tools [241–243]. Among the different questions addressed
edge and four sharing a corner). These differences in lattice therein, a prominent one was related to the possible types
structure result in different mass densities ( = 4250 g/cm3 of oxygen defects at the surface; in fact, three distinct oxy-
for rutile and  = 3894 g/cm3 for anatase) and electronic gen vacancy sites were tentatively identified: lattice, single-
band structures for the two forms of TiO2 . bridging, and double-bridging vacancies [244]. With the
Synthetic titanium oxide crystallizes in two polymorphs: widespread use of scanning-tunneling microscopy (STM),
anatase and rutile. Anatase is metastable and transforms atomically resolved investigations of TiO2 surfaces became
exothermally and irreversibly to rutile. Some properties of feasible and a more detailed picture of the rutile (110) sur-
TiO2 may strongly depend on its polymorphic phase. The face structure emerged [245–248]. Consistently, these studies
anatase-rutile transformation is strongly influenced by the corroborate the longstanding notion about the prominent
synthesis method, atmosphere, grain growth, and impuri- importance of surface defects as the active sites for vari-
ties. Some additives, such as ZrO2 and Al2 O3 , retard the ous types of chemical reactions [249–251], for example, for
anatase-rutile transformation, whereas others, such as CoO
and ZnO, accelerate such a process [236]. The anatase-to-
rutile phase transformation of doped nanostructured titania
was studied [237] using differential thermal analysis (DTA)
and X-ray diffraction (XRD). The presence of Cu and Ni
was found to enhance transformation as well as sintering.

Titanium Oxygen

90˚ 78.12˚

81.21˚ 92.43˚

Figure 8. Cross-section transmission electron microscopy image of a


Figure 7. Crystal structures of rutile (left) and anatase (right) TiO2 . nanocrystalline TiO2 anatase film. The nominal crystallite size is 16 nm.
518 Nanocrystalline TiO2 for Photocatalysis

the dissociation of water [252, 253]. The number of stud- radio-frequency sputtering on the structural properties of
ies on the structural properties of anatase single crystals the layers has been studied [269]. This characterization
is considerably smaller, which appears to be largely due to allowed the correlation of the inhibition of the grain growth
the difficulties encountered in preparing such surfaces in a of titania to the presence of iron oxide and its segregation
defect-free state. Nevertheless, a rather distinct picture of at grain boundaries. This behavior could be ascribed to a
the anatase TiO2 surface structure [254–257] and its prop- superficial-tension phenomenon. As a possible application
erties in terms of adsorption/desorption reactivity [258–260] of these thin films, it was observed that they were able to
has been achieved. sense CO down to the level requested for environmental
The structure and composition of a nanocrystalline sur- monitoring.
face may have a particular importance in terms of chemi- A study [270] of the structure and morphology of a tita-
cal and physical properties because of their small size. For nium dioxide photocatalyst (Degussa P25) reveals multipha-
instance, nanocrystal growth and manipulation relies heav- sic material consisting of an amorphous state, together with
ily on surface chemistry [261]. The thermodynamic phase the crystalline phases anatase and rutile in the approximate
diagrams of nanocrystals are strongly modified from those proportions 80/20. Transmission electron microscopy pro-
of the bulk materials by the surface energies [262]. More- vides evidence that some individual particles are a mixture
over, the electronic structure of semiconductor nanocrystals of the amorphous state with either the anatase phase or
is influenced by the surface states that lie within the bandgap the rutile phase, and that some particles, which are mostly
but are thought to be affected by the surface reconstruc- anatase, are covered by a thin overlayer of rutile that man-
tion process [263]. Thus, a picture of the physical properties ifests its presence by the appearance of Moiré fringes. The
of nanocrystals is complete only when the structure of the photocatalytic activity of this form of titanium dioxide is
surface is determined. reported as being greater than the activities of either of the
To understand and improve the applications of titanium- pure crystalline phases, and an interpretation of this obser-
oxide nanoparticles, it is extremely important to perform a vation has been given in terms of the enhancement in the
detailed investigation of the surface and the interior struc- magnitude of the space-charge potential, which is created
tural properties of nanocrystalline materials, such as rutile by contact between the different phases present and by the
and anatase with diameter of a few nanometers. Detailed presence of localized electronic states from the amorphous
experiments using X-ray absorption near-edge structure phase.
spectroscopy (XANES) demonstrated [264] that the pres-
ence of both defects and surface states strongly influence
5. ELECTRICAL PROPERTIES
the X-ray absorption spectra, even though the first nearest-
neighbor geometrical arrangement around the central Ti OF NANOCRYSTALLINE TiO2 FILMS
atom in both rutile and anatase is quite similar: the dif- Most studies into the carrier transport in nanocrystalline
ferences in the XANES spectra arise from the outer-lying TiO2 were carried out with the films in contact with elec-
atomic shells, indicating that “medium” to “long-range” trolytes, mostly due to their use in highly efficient electro-
effects play an important role to the near-edge features. chemical solar cells, often called “Grätzel cells” [38, 71]. In
In another study of this kind [265], a shorter Ti-O dis- this application, the pore surface is covered with an ultrathin
tance for surface TiO2 , resulting from Ti-OH bonding was organic dye layer and contacted with an electrolytic solution
observed together with a minor disorder of the lattice that penetrates the pore structure. The experimental work
in smaller nanoparticles. Nevertheless, the Ti sites largely indicates [271–274] that in this configuration the electrolyte
remain octrahedral even in particles with diameters of 3 nm. screens any electric field within the porous structure and
Because the interfacial electron/hole transfer occurs via sur- establishes diffusion conditions for the carrier propagation.
face Ti or O atoms, the observed structural changes around On the other hand, investigations in which the nanopores
the surface Ti atoms in small TiO2 particles could be respon- are in contact with an insulating medium (a gas or vacuum)
sible [265] for the unique photocatalytic properties. may allow one to obtain quantitative insight into the elec-
A qualitative analysis of opaque nanostructured glass- tronic properties of the material and the basic feature of car-
supported TiO2 films was carried out [266] using scanning rier transport. In a series of measurements [275–277] on the
force microscopy (SFM), and surface parameters such as electron transport in nanoporous TiO2 films with gas-filled,
average grain diameter, roughness exponent, and fractal insulating pores employing Pt/TiO2 , Schottky barrier struc-
dimension [267] were determined. The TiO2 surfaces exhibit tures indicate a barrier height of 1.7 eV, compatible with an
distinct roughness due to the large aggregates formed by the electron affinity of 3.9 eV for the TiO2 films. Below ∼300 K,
interconnected TiO2 particles. Fractal dimension was found tunneling transport through the barrier occurs, resulting in
to range between 2.10 and 2.45, depending on the scanned barrier lowering effects. Carrier drift mobilities, recombi-
range and the preparation method. The surface morphol- nation lifetimes, and their dependence on injection level
ogy of nanocrystalline materials prepared by compacting in TiO2 are reported. It is found that the mobility-lifetime
nanometer-sized clusters was investigated by SFM [268]; product is independent of injection level, while drift mobility
these materials had a grain size of 5–15 nm and contained and recombination lifetime change strongly with injection.
about 1019 interfaces per cubic centimeter. Upon heat treat- A trap-filling model appears as the most plausible model
ment, grains were found to fuse together forming bamboo- compatible with the experimental findings [277]. Compari-
like structures and then lined up as tubular structures. son with recent experiments on nanoporous films in contact
The influence of the iron concentration in mixed- with electrolytes indicate that the transport and recombina-
oxide (TiO2 and Fe2 O3  thin films prepared by reactive tion mechanism is qualitatively similar for the two cases.
Nanocrystalline TiO2 for Photocatalysis 519

Various observations indicate that electron transport in Such a power dependence of on the oxygen partial pres-
the nanocrystalline TiO2 dominates the transient response sure was noted also in recent work [297] investigating electri-
of the system. Transient photocurrent measurements reveal cal and defect thermodynamic properties of nanocrystalline
a very slow (∼millisecond), multiphasic response to both titanium oxide. At high O2 pressures, p(O2  > 1 mbar,
continuous wave [278, 279] and pulsed [280–282] illumina- the conductivity is constant, whereas at values p(O2  <
tion. The characteristic rise or decay time of the response 10−14 mbar a steep increase of with decreasing pressure
is dependent upon the intensity of the light source [278, was found, following a power dependence ∝ p(O2 n with
281–283]. Comparison with the transient response without n = −1/2 [297]. The plateau of conductivity at high oxy-
electrolyte indicates that it is the TiO2 , and not the elec- gen pressures can be interpreted as being a domain of ionic
trolyte, which is responsible for the very long tail [284, 275]. conductivity, an unexpected behavior for titanium dioxide.
A slow and multiphasic time dependence has also been In a coarse-grained material, dominant hole conductivity is
observed in the rereduction of oxidized dye molecules in a observed in this partial pressure range. This difference may
redox inactive environment [285]; the same work indicates be due to the high density of grain boundaries in nanocrys-
that the rate of dye reduction is controlled by the concen- talline ceramics, which can be preferred paths for diffu-
tration of electrons introduced into the TiO2 by externally sion at reduced temperatures. Furthermore, an increase in
applied bias. The wide range of time scales is consistent with ionic conductivity is also expected due to enhanced defect
the assumption that electron transport within the TiO2 is the formation in the space charge regions adjacent to grain
rate limiting step. boundaries [298]. At low oxygen partial pressure, nanocrys-
The slow processes are attributed to the trapping of elec- talline TiO2 exhibits enhanced electronic conductivity as
trons by a high density of localized states in the TiO2 . Since compared to coarse-grained TiO2 . The power exponent n =
the TiO2 grains are normally crystalline [286], the localized −1/2 can be explained under the assumption that doubly
states are believed to be concentrated at the grain bound- charged titanium interstitials are formed. The intrinsic disor-
aries and on the very large surface. There is evidence for der of titanium dioxide is reputedly of the cationic Frenkel-
intraband-gap states in bias-dependent optical absorbance type [299–302], although alternative defect models based
spectra [287] and surface photovoltage spectra. It would, on Schottky disorder are also described in the literature
therefore, be extremely useful to correlate the density and [303, 304]. In the domain of ionic conductivity, the activa-
nature of those states with the electronic transport proper-
tion energy of conduction is ∼10 ± 01 eV [297], a value
ties of the material.
typical of migration enthalpies for ionic defects. By contrast,
Investigating electron migration in nanostructured
the activation energy in the reduction-controlled regime was
anatase TiO2 films with intensity-modulated photocurrent
found to be ∼39 ± 02 eV.
spectroscopy [288], it was found that, upon illumination, a
In titanium oxide thin films prepared by a d.c. sputter-
fraction of the electrons accumulated in the nanostructured
ing technique onto glass substrates with average grain sizes
film is stored in deep surface states, whereas another
of 100–200 nm, the surface structure and phase morphology
fraction resides in the conduction band and is free to move.
of the films was found [305] to depend on the deposition
These data indicate that the average concentration of the
excess conduction band electrons equals about one electron conditions. The current-voltage characteristics of these films
per nanoparticle, irrespective of the type of electrode, the are ohmic for values of applied voltage lower than 0.5 V.
film thickness, or the irradiation intensity. For higher values, the mechanism of electrical conduction
The photocurrent in thin film TiO2 electrodes prepared is determined by space-charge-limited currents [306]; then,
by sol–gel methods was studied in [289] as a function of a power-law dependence was observed with n ∼ 2.3–2.9. In
film thickness. Films with thickness smaller than the space much thicker Ti oxide films (1.9–8 m) deposited by sput-
charge layer were found to show a larger photocurrent than tering [307], both the surface roughness and the internal
films with thickness larger than the space charge layer. It surface area increased with film thickness; this resulted in an
was concluded that the increase in photocurrent is due to enhancement of the incident photon-to-current efficiency.
the effective electron-hole separation throughout the whole Electrical and optical spectroscopic studies of TiO2
film thickness and the reduction of bulk recombination. The anatase thin films deposited by sputtering showed [308, 309]
use of TiO2 thin-film electrodes for photocatalytic devices that the metastable phase anatase differs in electronic prop-
might therefore be useful to gain high device performance. erties from the well-known, stable phase rutile. (From the
In a series of papers, Dittrich et al. carried out extensive broadening of the X-ray diffraction peaks, the average grain
investigations of the electrical conductivity in nanoporous size of the films is estimated to be in the range of 30–40 nm
TiO2 films [290–296]. Studying the temperature- and oxy- [308].) Resistivity and Hall-effect measurements revealed an
gen partial pressure-dependent conductivity of rutile and insulator-metal transition in a donor band in anatase thin
anatase, they noted [292] that is thermally activated with films with high donor concentrations. Such a transition is
EA = 085 eV, independent of the absolute value of and not observed in rutile thin films with similar donor con-
depends on p(O2  by a power law for p(O2  < 1–10 mbar. centrations. This indicates a larger effective Bohr radius
The electrical properties of reduced nanoporous TiO2 are of donor electrons in anatase than in rutile, which in turn
determined by surface chemical reactions which lead to the suggests a smaller electron effective mass in anatase. The
formation of shallow donor and deep trap states. Further- smaller effective mass in anatase is consistent with the
more, this group examined in detail the photovoltage in high-mobility, bandlike conduction observed in anatase crys-
nanocrystalline TiO2 [293, 295] and the injection currents in tals. It is also responsible for the very shallow donor ener-
these porous specimens [294, 296]. gies in anatase. Luminescence of self-trapped excitons was
520 Nanocrystalline TiO2 for Photocatalysis

observed in anatase thin films, which implies a strong lat- 6.1. Dependence of Photocatalytic Activity
tice relaxation and a small exciton bandwidth in anatase. on Film Structure and Phase
Optical absorption and photoconductivity spectra show that
anatase thin films have a wider optical absorption gap than The photoactivities of ultrafine TiO2 nanoparticles in
rutile thin films. The extrapolated optical absorption gaps of anatase, rutile, or mixed phases were tested in the photocat-
anatase and rutile films were found to be 3.2 and 3.0 eV, alytic degradation of phenol [322]. For TiO2 nanoparticles,
mainly in the anatase phase and mixed-phases, the photo-
respectively, at room temperature.
catalytic activities increased significantly with the content of
The observation of space charge limited currents (SCLC)
the amorphous part decreasing. The completely crystallized
in nanoscaled pure and chromium-doped titania was
rutile nanoparticles exhibited size effects in this photocat-
reported [310] and both the free-charge carrier density and
alytic reaction and the photocatalytic activity of rutile-type
the trapped-charge carrier density were given.
TiO2 nanoparticles with a size of 7.2 nm was much higher
Photoconductivity was also studied in compound systems;
than that with 18.5 nm or 40.8 nm and was comparable to
for example, in a TiO2 -C60 bilayer system the conductivity that of anatase nanoparticles.
increases significantly in the fullerene upon irradiation at A modified sol–gel process was used [323] to prepare
wavelengths <300 nm [311]. nano-structured TiO2 catalysts of controlled particle size
Although being an efficient photocatalyst for the detox- (i.e., 6, 11, 16, and 20 nm). The effect of TiO2 parti-
ification of organically charged waste water, titanium diox- cle size on gas-phase photocatalytic oxidation of toluene
ide suffers from the drawback of poor absorption properties was examined under dry and humid conditions. Main reac-
because of a bandgap of 3.2 eV. Thus, wavelengths shorter tion products were CO2 and water, although small amounts
than 400 nm are needed for light-induced generation of of benzaldehyde were also detected. The smaller particle
electron-hole pairs. That is the reason why doping with tran- size (i.e., 6 nm) led to higher conversion and complete
sition metal ions is interesting for inducing a reduction of the mineralization of toluene into CO2 and H2 O. Electronic
bandgap. However, this doping changes other physical prop- and structural effects (i.e., size and ensemble effects) were
erties such as lifetime of electron-hole pairs and adsorption responsible for the excellent performance of a 6 nm TiO2
characteristics [312]. catalyst for toluene photodegradation. The dependency of
the photoactivity on the crystallite size of anatase titania for
the decomposition of trichloroethylene (TCE) was investi-
gated [324]. It was found that the photoactivity of all tita-
6. PHOTOCATALYTIC PROPERTIES nia samples was linearly increased as the crystallite size of
OF NANOCRYSTALLINE TiO2 the anatase phase increased, regardless of the preparation
method, as long as there was no significant rutile phase.
Most experimental investigations reported a higher photo-
The enhancement of the photocatalytic activity of TiO2
catalytic efficiency in the anatase TiO2 phase; as a possi-
was investigated as a function of added amount of other
ble reason it was suggested that the recombination of the
oxides to promote desired oxidation or reduction reac-
electron-hole pairs produced by UV irradiation occurs more tions [325]. Mixed oxides of Nb2 O5 or Li2 O with TiO2 were
rapidly on the surface of the rutile phase, and the amount of prepared by the sol–gel process. The target material of
reactants and hydroxides attached to this surface is smaller dichloroacetic acid (DCA) was chosen for oxidation reac-
than on the surface of the anatase phase. tions and K2 Cr2 O7 for reduction. While the Nb-oxide had
The study of the photocatalytic activity of nanocrystalline a deleterious effect on the decomposition rate of DCA, the
TiO2 materials is a longstanding research effort [313–315]; excess electrons due to the doping of Nb2 O5 into TiO2 pro-
in most lab-scale experiments it was evaluated by means moted the reduction process for Cr6+ . Li2 O (1 wt%) with
of the degradation observed for typical substances (e.g., TiO2 was found to be the most efficient photocatalyst for
aqueous methyl orange, methylene blue, etc.) upon expo- DCA oxidation, resulting in photocatalytic activity of 50%.
sure of the specimen to UV irradiation. In such a way, A highly sensitive biochemical oxygen demand (BOD)
the possible influence of light intensity, structural proper- sensor using a commercial TiO2 semiconductor and photo-
ties, surface morphology, phase and chemical composition, catalytic pretreatment was developed to evaluate low BOD
resulting from various deposition or preparation methods, levels in river waters [326]. The photocatalytic oxidation was
could conveniently be explored [316–320]. Furthermore, investigated as a function of irradiation times, TiO2 concen-
any correlation with the optical or electrical properties of trations, and pH. The optimal irradiation time was 4 min.
the nanocrystalline films could thereby be investigated. In The sensor response was increased with increasing pH and
addition, alternative approaches have come under scrutiny. the responses obtained by photocatalysis to river samples
A new simple method for characterizing photocatalytic were higher than those obtained without photocatalysis.
activity by measuring photo-generated transient charge sep-
aration at the surface of semiconductor photocatalysts was
6.2. Influence of Surface Morphology
proposed [321]. In this technique, the charge separation gen-
erated by a pulse dye laser is obtained as a function of the and Defects
incident laser energy; thereby, the photocatalytic activity and Photocatalytic reduction of CO2 to organic compounds was
the type of surface reaction (reduction or oxidation) in tita- carried out [327] in a semiconductor suspension system
nium dioxide films were rapidly determined. In the following under simulated solar power using a TiO2 catalyst. Exper-
sections some examples of such studies will be given. imental results show that the photocatalytic activity can be
Nanocrystalline TiO2 for Photocatalysis 521

improved by depositing Pd or Ru on the TiO2 surface. Films polycrystalline anatase titanium dioxide [336]. Self-cleaning
of TiO2 dispersed or coated with platinum were deposited and antifogging effects of TiO2 films prepared by magnetron
on glass and Pt-buffered polyamide substrates, respectively, sputtering were investigated [161, 163] in terms of the photo-
by magnetron sputtering [328]. The photocatalytic activity of catalytic behavior by measuring the decomposition of methy-
the films was evaluated through the decomposition of acetic lene blue and the reduction of the contact angle between
acid under UV irradiation. The Pt-dispersed TiO2 film with water and TiO2 under ultraviolet irradiation. The phase con-
approximately 1.5 wt% platinum shows a maximum activity version from the rutile to the anatase TiO2 film leads to an
due to the formation of anatase phase with a fine grain size. enhancement of the activity; the anatase films with the best
Platinum particles ∼2 nm in thickness coated on anatase photocatalytic behavior are prepared at higher total pres-
film greatly improves activity. The activity shows a steplike sures (>1.50 Pa) and characterized by a high decomposition
dependence of film thickness, where the critical thickness efficiency, a contact angle about 10 after irradiation, and a
varies between 150 and 200 nm depending on the deposition good stability in darkness.
temperatures. The correlation between defects and activity Titanium dioxide thin films prepared with various sur-
was verified by measuring either the temperature depen- face morphologies by metalorganic chemical vapor deposi-
dence of electric resistance or the shift of binding energy tion were found to exhibit reversible wettability control by
from X-ray photoelectron spectroscopy (XPS). light irradiation [337]. These TiO2 surfaces became highly
In crystalline TiO2 films deposited by reactive RF mag- hydrophilic by UV irradiation, and returned to the initial
netron sputtering on glass substrates without additional relatively hydrophobic state by visible-light (VIS) irradia-
external heating, the photocatalytic activity was evaluated by tion. The hydrophobic-hydrophilic conversion induced by
the measurement of the decomposition of methylene blue UV light was ascribed to the increase in dissociated water
under UV irradiation [162]. The results showed that crys- adsorption on the film surface. By contrast, the conver-
talline anatase, anatase/rutile, or rutile films can be suc- sion from hydrophilic to hydrophobic by VIS irradiation was
cessfully deposited on unheated substrates. Anatase TiO2 caused by the elimination of water adsorbed on the sur-
films with a more open surface, a higher surface roughness, face due to the heat generated. Changes of the water con-
and a larger surface area, formed at higher total pressures, tact angle between hydrophilic states and hydrophobic ones
exhibit the best photocatalytic activity. The photocatalytic strongly depended on the roughness of the film surface.
activity of polycrystalline anatase TiO2 films was found [329]
The self-cleaning property of thin transparent TiO2 coat-
to be affected by the crystalline orientation that depends
ings on glass under solar UV light was studied [338] for two
on the deposition temperature; it was greater on the (112)-
compounds: palmitic (hexadecanoic) acid and fluoranthene,
oriented than on the (001)-oriented film. The former film
which are both present in the atmospheric solid particles.
exhibited a columnar structure resulting in a larger sur-
The removal rates of layers of these compounds sprayed on
face area for photocatalytic reaction than the films with
the self-cleaning glass were found to be sufficient for the
the (001)-preferred orientation. Furthermore, the introduc-
expected application. The identified intermediates (about
tion of structural defects associated with oxygen vacancies
40 for each compound) show the gradual splitting of the
was found [169] to create some energy levels around the
mid-gap, indicating that they could work as recombination palmitic acid chain and the oxidative openings of the aro-
centers of photo-induced holes and electrons, causing the matic rings of fluoranthene. In the case of palmitic acid,
decrease in photocatalytic activity. Therefore, the decrease the products give some indications about the photocatalytic
in the structural defects associated with oxygen vacancies mechanism. About 20% of the organic carbon contained in
appears to be important for improving the photocatalytic the initial compounds was transformed into volatile carbonyl
activity of the films. products.
A marked difference of the photocatalytic activity An extreme photo-induced hydrophilicity was achieved
between the TiO2 films coated on quartz and glass substrates [339, 340] when TiO2 films were covered by SiO2 overlay-
was confirmed [330], which would be interpreted in terms of ers (with 10–20 nm in thickness). These multilayer films
the difference in the photocarrier’s diffusion length induced exhibited much more extreme hydrophilicity than a TiO2
by impurity Na+ ions. These results lead to a conclusion film without overlayer. The surface analyses revealed that
that the crystallinity and defects of TiO2 as well as the film the enhanced photo-induced hydrophilic surface of the mul-
thickness and surface area have a great influence on the tilayer films exhibited an improved photocatalytic activity
photocatalytic activity. An enhanced photocatalytic activity towards decomposition of organic substances on their sur-
of TiO2 could be achieved also by deposition of the films on faces. An extreme light-induced superhydrophilicity was also
sulfonated glass substrates [331] or by using special support reported [341] for mesoporous TiO2 thin films (crystallite
materials [332, 333]. size ∼15 nm, surface area ∼50 m2 /g, pore size ∼3.6 nm).
Photo-oxidative self-cleaning and antifogging effects of For such films, the water contact angle was found to be
transparent titanium dioxide films has attracted considerable reduced essentially to zero upon UV-irradation for a dura-
attention for the past decade [334, 335]. In order to under- tion of about 60 min. In addition, the photocatalytic activity
stand the photo-induced hydrophilic conversion on titanium of those films could be enhanced by treating the substrate
dioxide coatings in details, it is inevitably necessary to under- surfaces with an H2 SO4 solution [342].
stand the relationship between the photo reaction and the In order to investigate the cathodic photoprotection of
surface crystal structure; this can be done, for example, by the steel from corrosion, stainless steel was coated with
an evaluation of the photo-induced hydrophilic conversion TiO2 thin films, applied by a spray pyrolysis [343]. It
on the different crystal faces of rutile single crystals and also was concluded that these coatings exhibit both a cathodic
522 Nanocrystalline TiO2 for Photocatalysis

photo-protection effect against corrosion and the frequently hν


reported photocatalytic self-cleaning effect.
CB - hν

6.3. Influence of Electronic Properties


VB +

Detailed spectroscopic investigations of the processes occur- - +


ring upon bandgap irradiation in colloidal aqueous TiO2 -
suspensions in the absence of any hole scavengers showed +
Schottky
[344] that while electrons are trapped instantaneously, that Barrier
is, within the duration of the laser flash (20 ns), at least
metal semiconductor
two different types of traps have to be considered for the
remaining holes. Deeply trapped holes are rather long-lived
and unreactive, that is, they are not transferred to the Figure 9. Metal-modified semiconductor photocatalyst particle.
ions of model compounds for photocatalytic oxidation like
dichloroacetate or thiocyanate. Shallowly trapped holes, on activity. The Pt/TiO2 system is probably the most frequently
the other hand, are in a thermally activated equilibrium with studied metal-semiconductor pair (see, e.g., [350, 351]);
free holes which exhibit a very high oxidation potential. The Figure 10 exemplifies the enhancement of the photo-
overall yield of trapped holes can be considerably increased catalytic activity of nanocrystalline TiO2 by platinization
when small platinum islands are present on the TiO2 sur- [350]: three commercially available TiO2 -catalysts, namely,
face, which act as efficient electron scavengers competing Degussa P25, Sachtleben Hombikat UV100, and Millennium
with the undesired e− /h+ recombination. While molecular TiONA PC50, were platinized by a photochemical impreg-
oxygen, O2 , reacts in a relatively slow process with trapped nation method with two ratios of platinum deposits (0.5 and
electrons, the adsorption of the model compounds on the 1 wt%). The photocatalytic activities of these samples were
TiO2 surface prior to the bandgap excitation appears to be determined using three different model compounds, EDTA,
a prerequisite for an efficient hole scavenging. 4-chlorophenol (4-CP), and dichloroacetic acid (DCA). Pla-
tinization resulted in all cases in an enhancement of the
6.4. Enhanced Photocatalytic Activity activity; Figure 10 shows the degradation of DCA as a func-
Via Surface Modifications tion of illumination time (light intensity 23 W/m2 at a wave-
length of 320–400 nm) for pure TiO2 and the two platinized
A driving force for research in heterogeneous photocataly- specimens. After 2 h of illumination, the initial concentra-
sis using TiO2 (and semiconductors in general) is the cre- tion of 120 ppm total organic carbon (TOC) was reduced
ation and application of systems capable of using natural to 2.3 ppm at pH 3 employing the best photocatalyst, in
sunlight to degrade a variety of organic and inorganic con- this case, Hombikat UV100 with 0.5 wt% Pt. For this sys-
taminants in wastewater or polluted air. As mentioned, the tem, an initial photonic efficiency (i.e., number of degraded
photocatalytic activity depends strongly, among other fac- molecules per number of incident photons) of ∼43% was
tors, on the wavelength range response. Since the bandgap obtained [350]. Apart from platinum, an enhanced pho-
of TiO2 is ∼3.2 eV, it is active only in the ultraviolet region tocatalytic activity has been also noted for other metals
which amounts to <10% of the overall solar intensity. Prin- and semiconductors. Their influence on the photocatalytic
cipally, there are several remedies to circumvent (at least activities has been studied in detail, for example, utilizing
partially) this limitation: (i) Deposition of metals on the
semiconductor; (ii) using multicomponent semiconductors;
(iii) surface modification with sensitizing dyes. The merits of
1.0 without Pt
these options will be outlined briefly in the following. 0.5 wt% Pt
1 wt% Pt
0.8
6.4.1. Deposition of Metals on the Surface
TOC/TOC0

The selectivity and efficiency of a photochemical reaction


0.6
can be improved by modifying the surface with a noble
metal. The deposition of metal particles on oxide surfaces
has been the subject of several recent reviews [345–348] and 0.4
therefore, there is, no need to duplicate it here. In terms
of photocatalytic activity, a drastic enhancement has, for the 0.2
first time, been observed for the photocatalytic conversion
of water into hydrogen and oxygen upon a fractional cover-
age of the TiO2 surface with platinum [349]. After excitation 0.0
0 20 40 60 80 100 120
the electron migrates to the metal where it becomes trapped
illumination time (min)
and electron-hole recombination is suppressed. The hole is
then free to diffuse to the semiconductor surface where oxi- Figure 10. Degradation of dichloroacetic acid (expressed as the relative
dation of organic species can occur. These processes are change of TOC) with platinized TiO2 in comparison to pure TiO2 as a
schematically depicted in Figure 9. The presence of the function of illumination at pH 3. Data from [350], D. Hufschmidt et al.,
metal can be beneficial also because of its own catalytic J. Photochem. Photobiol. A: Chem. 148, 223 (2002).
Nanocrystalline TiO2 for Photocatalysis 523

the following systems: Au-modified TiO2 [352, 353], silver- thoroughly in combination with TiO2 : TiO2 CdS [359–361],
modified titanium particles [354], transition-metal doped TiO2 CdSe [362], TiO2 coupled SnO2 [363], TiO2 capped
TiO2 photocatalysts [355–357], or rare-earth-doped TiO2 SiO2 [364], and some mixed-oxide systems like Fe2 O3 TiO2
nanoparticles [358]. [365] or ZrO2 TiO2 [366].
It should be noted in this context that various analytical
techniques like transmission electron microscopy or scan- 6.4.3. Surface Modification
ning force microscopy are often very useful in determining
with Sensitizing Dyes
the size of the particles and their distribution in the bulk
and at the surface of these nanocrystalline materials. Surface sensitization of a wide bandgap semiconductor pho-
tocatalyst like TiO2 via adsorbed dyes can increase the effi-
6.4.2. Composite Semiconductors ciency of the excitation step and, as a consequence, the
activity. This process can also expand the wavelength range
The advantage of composite semiconductors is usually of excitation. Some common dyes that are used as sensi-
twofold: first, to extend the photo-response by coupling tizers include erythrosin B, eosin, rhodamines, cresyl vio-
a large bandgap semiconductor with another featuring a let, thionine, porphyrins, [Ru(bpy)3 ]2+ and its analogues,
smaller one and, second, to retard the recombination of and many others (see, e.g., [102] for a more comprehen-
photo-generated charge carriers by injecting electrons into sive list). The individual charge-transfer and excitation steps
the lower lying conduction band of the large bandgap mate- involved in the dye sensitizer surface process are exemplified
rial. Two types of geometries have been employed: cap- in Figure 12. The primary excitation of an electron in the
ping the nanocrystallites of one semiconductor with those dye molecule occurs in either the singlet or triplet excited
of the second or bringing the nanocrystalline particles of state of the molecule; if the oxidative energy level of the
the two materials into intimate contact. The principle of excited state of the dye molecule, with respect to the con-
charge exchange and separation for both arrangements is duction band energy level of the semiconductor, is more
illustrated in Figure 11. Let us consider the case of cou-
negative, then the dye molecule can transfer the electron
pling CdS with TiO2 ; the energy of the excitation light is
to the conduction band of the semiconductor. The surface
too small to directly excite the TiO2 particle, but it is large
accepts electrons from the dye molecules which, in turn,
enough to excite an electron from the valence band across
can be transferred to reduce an organic acceptor molecule
the bandgap of CdS (Eg = 25 eV) to the conduction band.
adsorbed on the surface. The dye-sensitized semiconduc-
According to the energetics, the hole produced in the CdS
tor can also be used in oxidative degradation of the dye
valence band remains in the CdS particle, whereas the elec-
molecule itself after charge transfer; this appears to be an
tron transfers to the conduction band of the TiO2 particle;
important process in view of the large quantities of dye sub-
this increases the charge separation and efficiency of the
stances in wastewater from the textile industries and others.
photocatalytic process. The separated electrons and holes
In passing, it can be noted that the process of utilizing sub-
are then free to undergo electron transfer with adsorbates at
bandgap excitation with dyes that absorb strongly in the vis-
the surface. While the mechanisms of charge separation in a
ible for photosensitization is frequently employed in color
capped system are similar, in a capped semiconductor only
photography and other imaging science applications. This
one of the charge carriers is accessible at the surface for cat-
approach of light-energy conversion is also similar to plant
alytic reactions. Several semiconductors have been studied
photosynthesis, in which chlorophyll molecules act as light
harvesting antenna molecules.
CB
-
Generally, the high porosity and strong bonding char-
TiO2
CB - acter of nanostructured TiO2 (and other) semiconductor

SnO2 films facilitate the surface modification with organic dyes
hν'
or organometallic complexes. For example, photoconversion
VB + A efficiencies in the range 10–15% in diffused daylight have
+
VB been reported [367] for nanostructured TiO2 films modified
A+ with a ruthenium complex. The charge injection from a sin-
glet excited sensitizer into the conduction band of a large
(a)
bandgap semiconductor is thought to be an ultrafast process
CdS TiO2
B
hν CB - -
CdS CB - B- S*
CB CB CB
hν -
S

hν' A S+ A S+ A-
A VB +
VB VB VB

+
A+ VB
Figure 12. Sequence of excitation and charge-transfer steps using a dye-
(b) molecule sensitizer. In the first step, the sensitizer (S) is excited by an
incident photon of energy h and an electron is transferred into the
Figure 11. Photo-induced charge separation in composite semiconduc- conduction band of the semiconductor particle; the electron then can
tor particles: (a) capped and (b) coupled semiconductor nanocrystal- be transferred to reduce an organic acceptor molecule (A) adsorbed on
lites. Photo-generated charge carriers move in opposite directions. the surface.
524 Nanocrystalline TiO2 for Photocatalysis

occurring at a picosecond time scale; in the case of different a dye-sensitized nanocrystalline TiO2 solar cell, depicting
organic dyes, it has been shown [368, 369] that charge injec- the relevant energy levels and the pathway for the photo-
tion takes place within 20 ps. A similar fast electron transfer excited electrons. In this specific example, a ruthenium com-
has also been noted for [Ru(H2 O)2 ]2− on a TiO2 surface at plex [367] was adsorbed as a dye onto the TiO2 and an
very low coverage [370]. Progress in femtosecond laser spec- I− /I−
3 redox couple was used in the electrolyte. Contrary to
troscopy opened a venue for investigations on even much conventional semiconductor devices, in the dye-sensitized
shorter time scales. In fact, in recent studies charge car- cells the function of light absorption is separated from the
rier injection times in the range of 20–200 fs were reported charge-carrier transport. The Ru-complex has to absorb the
[371–375] for various dyes on nanocrystalline TiO2 particles. incident sunlight and to effect, via this energy, the electron-
Significant enhancement effects of electron acceptors transfer reaction (numbers  1 and 2 in Fig. 13). Apart from
(additives) such as hydrogen peroxide, ammonium per- supporting the dye, the TiO2 film acts as an electron accep-
sulfate, potassium bromate, and potassium peroxymono- tor and electronic conductor: the electrons injected into the
sulfate (oxone) on the TiO2 photocatalytic degradation TiO2 conduction band drift across the nanocrystalline film
of various organic pollutants were observed already in to the conducting glass support which functions as current
early investigations [376]. The results showed that these collector (3 in Fig. 13). At the counterelectrode, the elec-
additives markedly improved the degradation rate of 2,4- tron is transferred to the redox couple in the electrolyte (4 )
dichlorophenol. The enhanced photocatalytic oxidation of which, in turn, serves to regenerate the dye ( 5 ). The cell
sulfide ions on phthalocyanine modified titania was ascribed voltage observed under illumination, V , is determined by
[377] to the additional formation of superoxide radicals. the difference between the Fermi-level of TiO2 and the elec-
Sensitization of wide bandgap semiconductor electrodes trochemical potential of the electrolyte (cf. Fig. 13 [69]).
by dyes absorbing visible light are routinely used also in dye-
sensitized photoelectrochemical cells, in order to achieve
high photon-to-current conversion efficiencies [378, 379]. 7. PHOTOCATALYTIC APPLICATIONS
The preparation and dynamics of interfacial photosensi- OF NANOCRYSTALLINE TiO2
tized charge separation in metal oxides such as TiO2 films
has been reviewed [70, 71, 380]. Principally, the photo- While many examples of the photocatalytic activity of
physical reactions occurring in those dye-sensitized injec- nanocrystalline TiO2 have been presented already in the
tion solar cells, which are based on a dye adsorbed onto foregoing sections, the following discussions will focus on
a porous TiO2 layer, are very similar to those relevant to novel and important (and sometimes large-scale) applica-
photocatalysis. Because of their importance as an energy tions. A very prominent area appears to be environmental
source, Figure 13 presents a schematic drawing [69] of such catalysis [31, 32, 45] which, in recent years, has expanded
considerably beyond the traditional fields like NOx removal
from stationary and mobile sources, or the conversion of
TCO glass dye TCO glass with Pt volatile organic compounds (VOC). According to [381],
these potential new areas include:
E S+/ S*
3 e- 2 hν (i) catalytic technologies for liquid or solid waste
reduction or purification;
1 4 ∆V (ii) use of catalysts in energy-efficient catalytic tech-
e
- nologies and processes;
I- / I3- (iii) reduction of the environmental impact in the use or
e-
5 disposal of catalysts;
S+/ S
e- (iv) new ecocompatible refinery, chemical or nonchem-
e- ical catalytic processes;
(v) catalysis for greenhouse gas control;
(vi) use of catalysts for user-friendly technologies and
reduction of indoor pollution;
por-TiO2 electrolyte I- / I3- (vii) catalytic processes for sustainable chemistry;
(viii) reduction of the environmental impact of transport.
e- load e- Hence, (photo)catalysis in environmental applications can
be instrumental in promoting the quality of life and environ-
Figure 13. Schematic outline of a dye-sensitized photovoltaic cell, ment, in promoting a more efficient use of resources, and in
showing the electron energy levels in the different phases. The system promoting sustainable processes and products.
consists of a semiconducting nanocrystalline TiO2 film onto which a Because of the tremendous importance of those environ-
Ru-complex is adsorbed as a dye and a conductive counterelectrode, mentally related areas, the use of nanocrystalline TiO2 for
while the electrolyte contains an I− /I−3 redox couple. The cell volt-
such photocatalytic applications is illustrated in this section
age observed under illumination corresponds to the difference, V ,
between the quasi-Fermi level of TiO2 and the electrochemical potential
by means of selected examples. Before doing so, it is stressed
of the electrolyte. S, S∗ , and S+ designate, respectively, the sensitizer, that there exists at least one other rather important issue
the electronically excited sensitizer, and the oxidized sensitizer. See text in this context. In fact, hydrogen production from aqueous
for details. Adapted from [69], A. Hagfeldt and M. Grätzel, Chem Rev. solutions using semiconductor particles such as, CdS, TiO2 ,
95, 49 (1995). © 1995, American Chemical Society. WO3 as photocatalysts is envisaged to become a potential
Nanocrystalline TiO2 for Photocatalysis 525

major application of these materials, and new concepts and whereas it inhibited that of acetone. As for the effect of pho-
approaches are developed continuously. For example, a new ton flux, it was found that photocatalytic degradation occurs
photocatalytic reaction that splits water into H2 and O2 in two regimes with respect to photon flux: for illumina-
was designed [382] by a two-step photo-excitation system tion levels distinctly blow 1000–2000 W/cm2 , the photocat-
composed of a IO− −
3 /I shuttle redox mediator and two dif- alytic degradation rate increased linearly with photon flux,
ferent TiO2 photocatalysts, Pt-loaded TiO2 -anatase for H2 whereas for power densities above that value, the rate was
evolution and TiO2 -rutile for O2 evolution. Simultaneous found to scale with the square root of the flux. Figure 14
gas evolution of H2 (180 mol/h) and O2 (90 mol/h) was shows some of those data [388], depicting in panel (a), the
observed from a basic (pH = 11) NaI aqueous suspension degradation of methanol as a function of UV illumination
of two different TiO2 photocatalysts under UV irradiation time for five different initial concentrations. (Using TiO2
( > 300 nm, 400 W high-pressure Hg lamp). An exten- anatase nanocrystallites with 7 nm diameter in a solution,
sive review [383] assesses photocatalytic efficiencies with ref- in this work photocatalytic TiO2 films were deposited onto
erence to hydrogen production by means of light energy glass substrates by dip-coating.) The reaction kinetics were
in the presence and absence of loaded metals, electron- found to follow the L-H model, in which the reaction rate r
donors/acceptors, and hole scavengers. varies proportionally with the surface coverage  according
to
7.1. Reduction/Removal of Toxic Gases
kKc
The conversion of nitrogen oxides to less toxic compounds is r = k = (9)
important both because of their toxicity and the global atmo- 1 + Kc
spheric pollution. NOx can be converted to N2 and other
nitrogen compounds by reduction. TiO2 -loaded zeolites and where c is the concentration of the VOC and k and K are,
the vanadium silicate-1 were found [384] to decompose NO respectively, the reaction rate constant and the adsorption
under irradiation, in particular, TiO2 included in zeolite cav- equilibrium constant. Figure 14(b) exemplifies this finding,
ities results in complete decomposition into N2 and O2 . Tita- showing the initial reaction rates r0 derived from data like
nium oxide catalysts prepared within the Y-zeolite cavities those in Figure 14(a) as a function of the respective initial
via an ion-exchange method exhibit [385] high and unique methanol concentrations c0 . The solid line in Figure 14(b)
photocatalytic reactivities for the decomposition of NO into is a fit to the data according to Eq. (9).
N2 and O2 , as well as the reduction of CO2 with H2 O show-
ing a high selectivity for the formation of CH3 OH. It was
also found that the charge transfer excited state of the tita-
concentration (10–3 mol/m3)

20
nium oxide species, (T3+ -O− )∗ , plays a vital role in these (a)
methanol
unique photocatalytic reactions. In yet another approach, an 15
efficient catalytic reduction of NO at low temperature by
means of NH3 could be achieved using Mn-, Cr-, or Cu-
10
oxides on a nanocrystalline TiO2 support [386].
The NOx removal process was studied experimentally in
a pulsed corona discharge combined with the TiO2 photo- 5
catalytic reaction [387]. NO2 was found to adsorb easily on
the photocatalyst surface, whereas NO was hardly adsorbed. 0
0 2 4 6 8 10
Addition of water vapor enhanced the NO2 adsorption. It illumination time (min)
initial reaction rate r0 (10–3 mol/m3min)

was concluded that the main role of the plasma-chemical


reaction in this system is the oxidation of NO into NO2 .
A considerable part of NO2 is adsorbed on the photocatalyst 2.0
surface, and is transformed to HNO3 through photocatalytic
reaction with OH. 1.5
The photocatalytic degradation of VOCs in the gas phase
constitutes another very important example in this range of 1.0
applications. Utilizing variously prepared TiO2 photocata-
lysts (e.g., deposited on glass fiber cloth, as pellets or as 0.5
(b)
thin films), the photo-induced reactions of trichloroethylene,
acetone, methanol, and toluene were investigated [388–390]. 0.0
0 5 10 15 20 25
The photocatalytic degradation rate was observed [388] to
initial concentration c0 (10–3 mol/m3)
increase with increasing initial concentration of the VOCs,
but remained almost constant beyond a certain concen-
Figure 14. (a) Photocatalytic degradation of methanol with different
tration. It matched well with the Langmuir–Hinshelwood
initial concentrations as a function of UV illumination time (light inten-
(L-H) kinetic model [11]. For the influence of water vapor sity 2095 W/cm2 at a wavelength of 254 nm) at a H2 O concentration
in a gas-phase photocatalytic degradation rate, there was an of 0.38 mol/m3 and a reaction temperature of 45  C. (b) Initial reaction
optimum concentration of water vapor in the degradation of rates r0 as derived from the data in (a) versus the initial methanol con-
trichloroethylene and methanol. Furthermore, water vapor centrations; the solid line is a fit according Eq. (9). Data from [388],
enhanced the photocatalytic degradation rate of toluene, S. B. Kim and S. C. Hong, Appl. Catal. B: Environ. 35, 305 (2002).
526 Nanocrystalline TiO2 for Photocatalysis

7.2. Degradation of Organic Compounds In a recent study [424], the photocatalytic degradation of
five dyes in TiO2 aqueous suspensions under UV irradiation
The degradation of organic compounds is probably the
has been investigated; it was attempted to determine the
most widely used photocatalytic application of nanocrys-
individual steps of such a degradation process by varying the
talline TiO2 and other semiconductor materials. In an aque-
aromatic structures, using either anthraquinonic (Alizarin
ous environment, the holes created under UV irradiation
S (AS)), or azoic (Crocein Orange G (OG), Methyl Red
are scavenged by surface hydroxyl groups to generate • OH (MR), Congo Red (CR)) or heteropolyaromatic (Methy-
radicals that then promote the oxidation of organics. This lene Blue (MB)) dyes. Figure 15 exemplifies the photocat-
radical-mediated oxidation has been successfully employed alytic degradation of three of these dyes (CR, OG, and MR)
in the mineralization of several hazardous chemical contam- as a function of UV irradiation. The initial reaction rates
inants such as hydrocarbons, haloaromatics, phenols, halo- were found to fall in the range from 1.9 mol/l min (for
genated biphenyls, surfactants, and textile and other dyes CR) to 3.6 mol/l min (for MR). In addition to a prompt
[102]. removal of the colors, TiO2 /UV-based photocatalysis was
The possible photocatalytic decomposition of a broad simultaneously able to fully oxidize the dyes, with a com-
range of organic compounds has been investigated using plete mineralization of carbon into CO2 . Sulfur heteroatoms
nanocrystalline TiO2 particles. Detailed studies reported were converted into innocuous SO2− 4 ions. The mineraliza-
the oxidation of dissolved cyanide [391], the degradation tion of nitrogen was more complex. Nitrogen atoms in the
of various kinds of acids [392–398], and of several herbi- 3-oxidation state, such as in amino groups, remain at this
cides [399–402], for the photocatalytic oxidation of toluene, reduction degree and produced NH+ 4 cations, subsequently
benzene, cyclohexene, and benzhydrol [403–406] or for and very slowly converted into NO− 3 ions. For azo-dye (OG,
the 1,1 -dimethyl-4,4 -bipyridium dichloride decomposition MR, CR) degradation, the complete mass balance in nitro-
[407]. In another application, a titanium oxide photocata- gen indicated that the central N N azo group was
lyst of ultra-high activity has been employed for the selective converted into gaseous dinitrogen, which is the ideal issue
N-cyclization of an amino acid in aqueous suspensions [408]. for the elimination of nitrogen-containing pollutants. The
Anatase crystallites of average diameter of ∼15 nm were aromatic rings were submitted to successive attacks by pho-
platinized by impregnation from aqueous chloroplatinic acid togenerated • OH radicals leading to hydroxylated metabo-
solution followed by hydrogen reduction. The catalyst was lites before the ring opening and the final evolution of CO2
suspended in an aqueous L-lysine (Lys) solution and photo- induced by repeated reactions with carboxylic intermediates.
irradiated under argon at ambient temperature to obtain The photocatalytic degradation of acid derived azo dyes
L-pipecolinic acid. in aqueous TiO2 suspensions follows apparently first-order
The photocatalytic degradation and oxidation of phenol kinetics [425, 426]. The site near the azo bond (C N N-
and phenol-based compounds has been examined quite fre- bond) is the attacked area in the photocatalytic degrada-
quently [409–414]. The decomposition of aqueous phenol tion process, while the TiO2 photocatalytic destruction of
solutions to carbon dioxide have been studied using natural the C N( ) bond and N N bonds leads to fading
sunlight in geometries simulating shallow ponds [415]. The of the dyes. The pH effect on the TiO2 photocatalytic degra-
photocatalyst was titanium dioxide freely suspended in the dation of the acid-derived azo dyes varies with dye struc-
solution or immobilized on sand or silica gel. Photodegra- ture. Hydroxyl radicals play an essential role in the fission
dation rates were approximately three times faster with the of the C N N conjugated system in azo dyes in TiO2
free suspension than with the immobilized catalyst under photocatalytic degradation. Metalized azo dyes were studied
the same conditions, and were dependent on the time of
year and the time of day. The seasonal variation correlated
80
roughly with seasonal solar irradiance tabulations for the
UV component of the spectrum. For 10 ppm of phenol, the CR
maximum rate of solar degradation resulted in a decrease OG
concentration (µmol/l)

in concentration to 10 ppb in less than 80 min with total 60 MR


mineralization in 110 min.
An efficient degradation of aqueous phenol was achieved
[416] by a new rotating-drum reactor coated with a TiO2 40
photocatalyst, in which TiO2 powders loaded with Pt are
immobilized on the outer surface of a glass drum. The reac-
tor can receive solar light and oxygen from the atmosphere 20
effectively. It was shown experimentally that phenol can be
decomposed rapidly by this reactor under solar light: with
the used experimental conditions, phenol with an initial con-
0
centration of 22.0 mg/dm3 was decomposed within 60 min 0 50 100 150 200
and was completely mineralized through intermediate prod- illumination time (min)
ucts within 100 min.
The photocatalytic degradation of various types of dyes Figure 15. Photocatalytic degradation of three different dyes, Congo
appears to be another prominent and extensively explored Red (CR), Crocein Orange (OG), and Methyl Red (MR), given in
application of nanocrystalline TiO2 in environmental cataly- terms of the concentration versus the time of illumination. Data from
sis [417–423]. [424], H. Lachheb et al., Appl. Catal. B: Environ. 39, 75 (2002).
Nanocrystalline TiO2 for Photocatalysis 527

[427] under TiO2 photocatalytic and photosensitized condi- the photocatalytic degradation of industrial residual waters,
tions in aqueous buffering solutions. The size and strength of the use of peroxydisulfate (S2 O2− 8 ) as an additional oxidant
intramolecular conjugation determines apparently the light- (electron scavenger) was observed to have an outstanding
fastness of the dyes; the more powerful OH radicals in TiO2 effect, producing an important increase in the degradation
photocatalytic process are highly reactive towards the azo rate [437]. The impact of pH and the presence of inorganic
dyes. ions and organic acids commonly found in natural waters on
rates of TiO2 photocatalyzed trinitrotoluene (TNT) trans-
formation and mineralization was examined [438]. Raising
7.3. Wastewater and Soil Remediation
the pH slightly increased the rate of TNT transformation,
The major causes [428] of surface water and groundwa- primarily as a result of an increased rate of TNT photoly-
ter contamination are industrial discharges, excess use of sis, but significantly reduced rates of mineralization due to
pesticides, fertilizers (agrochemicals), and landfilling domes- increased electrostatic repulsion between the catalyst surface
tic wastes. Typically, the wastewater treatment is based and anionic TNT intermediates. The presence of inorganic
upon various mechanical, biological, physical, and chemical anions did not substantially hinder TNT transformation at
processes. After filtration and elimination of particles in sus- alkaline pH, but mineralization rates were diminished when
pension, the biological treatment is the ideal process (natu- the anion either adsorbed strongly to the photocatalyst or
ral decontamination). Unfortunately, organic pollutants are was an effective hydroxyl radical scavenger.
not always biodegradable; a promising approach then relies Immobilized TiO2 photocatalysts were used to sterilize
on the formation of highly reactive chemical species, which and reclaim the wastewater of bean sprout cultivation from
degraded the more recalcitrant molecules into biodegrad- a continuous hydrocirculation system [439]. The photocata-
able compounds. These are called the advanced oxidation lysts effectively killed bacteria and degraded organic pollu-
processes (AOPs). Although there exist differences in their tants in the wastewater. Stimulation of bean sprout growth
detailed reaction schemes, their common feature is the pro- and suppression of decaying pathogens were also induced by
duction of OH radicals (• OH); these radicals are extraordi- the TiO2 photocatalytic activity.
narily reactive species (oxidation potential 2.8 V). They are Photocatalytic decomposition of seawater-soluble crude
also characterized by a low selectivity of attack, which is a oil fractions using high surface area colloid nanoparticles
useful attribute for an oxidant used in wastewater treatment of TiO2 under UV irradiation was explored [440]; although
and for solving pollution problems. These photocatalytic no mineralization occurred due to photolysis, important
degradation of wastewater employing nanocrystalline TiO2 chemical changes were observed in the presence of TiO2 ,
has been examined in various set-ups [429] and pilot-plant with the degradation reaching 90% (measured as dissolved
scale solar photocatalytic experiments have been realized organic carbon, (DOC)) in waters containing 9–45 mg C/l of
[428]. seawater-soluble crude oil compounds after 7 days of artifi-
Several recent studies reported on the removal or reduc- cial light exposure. During light exposure, transient interme-
tion of metals or metal-containing contaminants in waste- diates that showed higher toxicity than the initial compounds
water, based on the principles outlined in the foregoing were observed, but were subsequently destroyed. Hetero-
paragraph. Those investigations examined, for example, the geneous photocatalysis using TiO2 was considered to be a
removal of cadmium and mercury from water using mod- promising process to minimize the impact of crude oil com-
ified TiO2 nanoparticles [430, 431], the radical, mediated pounds on contaminated waters.
photo-reduction of manganese ions in UV-irradiated titania TiO2 -photocatalytic degradation of a cellulose effluent
suspensions [432], the simultaneous photocatalytic Cr(VI) was evaluated [441] using multivariate experimental design.
reduction and dye oxidation in a TiO2 slurry reactor [433], The effluent was characterized by general parameters such
or the removal of iron(III) cyanocomplexes [434]. as adsorbable organic halogens (AOX), TOC, COD, color,
While the efficient use of a photocatalytic process in the total phenols, acute toxicity, and by the analysis of chlori-
presence of TiO2 to degrade many different types of pol- nated low molecular weight compounds using GC/MS. The
lutants in wastewater has been confirmed repeatedly, the optimal concentration of TiO2 was found to be around 1 g/l.
question of how to efficiently separate and reuse TiO2 from After 30 min of reaction more than 60% of the toxicity was
treated wastewater became a notable problem in the appli- removed and after 420 min of reaction, none of the initial
cation of a TiO2 photo-oxidation process. A recent study chlorinated low molecular weight compounds were detected,
[435] aimed to develop an advanced process for dyeing suggesting an extensive mineralization.
wastewater treatment, in which dyeing wastewater was ini- Photocatalysts, based on titanium dioxide, were used for
tially treated by an intermittently decanted extended aer- the purification of contaminated soil polluted by oil [442].
ation (IDEA) reactor to initially remove biodegradable Commercially produced slurry of titanium dioxide was mod-
matters and further treated in a TiO2 photocatalytic reac- ified with barium, potassium, and calcium. The experiments
tor for complete decolorization and high chemical oxygen were performed under natural conditions in summer months
demand (COD) removal. Suspended TiO2 powder used in (July and August) applying direct solar-light irradiation. The
the photo-oxidation was separated from slurry by a mem- most active photocatalyst for soil purification was titanium
brane filter and recycled to the photo reactor continuously. dioxide modified with calcium.
Photocatalytic destruction of chlorinated solvents in water Two different photocatalysts, namely, Hombikat UV100
with solar energy was investigated [436] using a near- (Sachtleben Chemie) and P25 (Degussa) have been used in
commercial scale, single-axis tracking parabolic trough sys- batch experiments [443] to compare their ability to degrade
tem with a glass pipe reactor mounted at its focus. In the toxic components of a biologically pretreated landfill
528 Nanocrystalline TiO2 for Photocatalysis

leachate. A strong adsorption of the pollutant molecules was microcystin-LR was destroyed within the first 5 min of
observed for both TiO2 -powders, with a maximum of almost photocatalysis, with 97% of the toxin removed in 20 min.
70% TOC reduction for Hombikat UV100. The addition of 0.1% H2 O2 to the photocatalytic system
was found [448] to further enhance the degradation rate:
7.4. Purification of Drinking Water 99.6% of microcystin-LR was destroyed within 5 min and
no toxin was left after 10 min of photocatalysis.
Pathogens in drinking water supplies can be removed by Photocatalytic inactivation of different bacteria and bacte-
sand filtration followed by chlorine or ozone disinfection. riophages in drinking water at different TiO2 concentrations
These processes reduce the possibility of any pathogens with or without concurrent exposure to O2 was studied in
entering the drinking water distribution network. How- [449] using UV irradiation (5.5 mW/cm2 at 365 nm). For
ever, there is doubt about the ability of these methods example, for this light intensity, the most effective inacti-
to remove chlorine-resistant microorganisms including pro- vation of Escherichia coli CN13 was obtained at 1 g/l sus-
tozoan oocysts. Titanium dioxide (TiO2 ) photocatalysis is pension of TiO2 , resulting in a reduction by five orders of
a possible alternative/complementary drinking water treat- magnitude in 5 min. Under the same experimental condi-
ment method and several studies [444, 445] reported a tions, MS2 bacteriophage was reduced by four orders of
strong and swift photocatalytic inactivation of bacteria and magnitude, also in 5 min. The addition of O2 into the exper-
bacteriophages in aqueous solutions. For example, the rate imental environment increased the inactivation of Deinococ-
of disinfection was explored using TiO2 electrodes prepared cus radiophilus by four orders of magnitude in 60 min.
by the electrophoretic immobilization of TiO2 powders with
different crystallinity. These electrodes were tested for their
photocatalytic bactericidal efficiency with E. coli K12 as a 7.5. Miscellaneous Photocatalytic
model test organism [446]. Similar studies were reported for Applications
natural water from a river [447].
Cyanobacterial toxins produced and released by It may have become apparent from the foregoing dis-
cyanobacteria in freshwater around the world pose a cussions and examples that the solution of environmental
considerable threat to human health if present in drinking problems constitutes one of the (if not the) major driving
water sources. Therefore, various treatments have been forces in research and development in photocatalysis using
applied to remove these toxins. The effectiveness of TiO2 nanometer-sized TiO2 (and other semiconductor) particles.
photocatalysis for the removal of microcystin-LR from Another one, of course, is the production of hydrogen from
water has been established [448]. Not only does the process water splitting. Apart from these main applications, there
rapidly remove the toxin but also the by-products appear to exist, on the other hand, many attempts to explore novel
be nontoxic. The photocatalytic process has also significantly areas for the photocatalytic use of nanocrystalline TiO2
reduced the protein phosphatase 1 (PP1) inhibition. Protein materials. To give a flavor of the diversity of these efforts,
phosphatase 1 inhibition is potentially one of the most some selected (and mostly recent) examples follow.
serious harmful effects to humans who may consume water Nano-sized titanium oxide (TiO2  thin films have been
contaminated by microcystins. Figure 16 shows some of explored for alcohol-sensing applications. TiO2 thin films
these data, namely, the reduction of the microcystin-LR with different doping concentrations were prepared on alu-
concentration and the PP1 inhibition as a function of the mina substrates [450] using the sol–gel process using the
illumination time. The results indicate that about 86% of spin-coating technique for ethanol and methanol alcohol.
Experimental results indicated that the sensor is able to mon-
itor alcohols selectively at ppm levels; the films are stoichio-
concentration microcystin-LR (µg/ml)

1000 100
protein phosphatase inhibitor (%)

metric with carbon as the dominant impurity on the surface.


The morphologies and crystalline structures of the films were
800 80 studied by scanning electron microscopy (SEM) and XRD.
X-ray diffraction patterns showed that the films are pure
anatase phase up to an annealing temperature of 600  C.
600 60
As the annealing temperature increased to 800  C, a small
amount of rutile phase formed along with the anatase phase.
400 40 Optical waveguides were prepared by depositing a sol–gel-
derived titania film onto a silica substrate [451]. The titania
200 20 film is mesoporous, with pore sizes ranging from 3 to 8 nm.
Deposition of the titania does not change the critical angle
of total internal reflection. Thus, the titania-coated wave-
0 0 guides propagate light in an attenuated total reflection mode,
0 20 40 60
despite the relatively high refractive index (n = 1.8 in air) of
illumination time (min)
the titania film relative to the silica substrate (n = 05).
Figure 16. Destruction and protein phosphatase (PP1) inhibition of
The light output of electric lighting gradually decreases
microcystine-LR via TiO2 photocatalysis as a function of the duration due to stain buildup on lamps and covers during operation.
of UV illumination (xenon lamp with 480 W at a wavelength of 330– Roadway, and especially tunnel lighting, experiences a large
450 nm). Data from [448], I. Liu et al., J. Photochem. Photobiol. A: amount of contamination due to dust, carbon particles found
Chem. 148, 349 (2002). in vehicle engine exhausts and other airborne contaminants,
Nanocrystalline TiO2 for Photocatalysis 529

which results in the rapid deterioration of the light output. enhance the efficiency of the excitation step and, hence, the
Photocatalytic reactions caused by TiO2 are known to decom- catalytic activity.
pose such stains. This reaction is caused by the absorption Electron-hole pair The absorption of a photon of sufficient
of UV light ( ≤ 400 nm, corresponding to the bandgap energy may excite in a semiconductor an electron from the
of ∼3 eV) irradiated from lamps or the sun, and followed valence band to the conduction band, thereby creating a hole
by oxidation. Extensive field tests revealed [452] that a fine in the valence band.
film coating of TiO2 on lamps and luminaires can effec- Nanocrystalline A material composed of individual crys-
tively decompose various organic compounds such as vehicle tallites which have a size in the range of nanometer (nm);
exhaust gases, oil, nicotine, etc. This leads to an improvement 1 nm = 10−9 m.
of the luminous performance of installed lighting systems and
Photocatalysis A catalytic reaction triggered or enhanced by
reduces the cost of maintenance by approximately one-half.
illuminating the system with visible or ultraviolet irradiation.
It has been recently found [453] that photocatalytic TiO2
This reaction involves normally the electronic excitation of the
coated with polycarbonate (PC) releases a huge amount of
catalyst via the absorption of photons and an interfacial charge
exothermic energy in the temperature range between 200
transfer to an adsorbed species. Typically, the photocatalyst is
and 400  C (ca. 1.85 kJ/g). The strong interaction between
not consumed in the reaction.
oxygen-deficient sites in TiO2 and carbonyl groups of PC
mediated by a good PC solvent is found to be a prerequi- Photocatalytic degradation The removal or reduction of
site for a release of the enormous amount of exothermic (usually unwanted) substances via a photocatalytic reaction.
energy. This finding suggests that PC-coated TiO2 powders Quantum yield The probability of product formation per
or related oxides work as a combustion-assisting agent in a adsorbed photon in a photocatalytic reaction.
relatively lower temperature range and can be utilized for Titanium oxide Titanium dioxide with the nominal compo-
incineration applications in order to suppress the formation sition TiO2 is a semiconductor with a band gap of ∼3.2 eV; it
of extremely toxic dioxins. exists in three different crystalline modifications, two of which
A somewhat unusual application reported [454] the pho- (anatase and rutile) are commonly employed in photocatalysis.
tocatalytic deposition of a gold particle onto the top of a
SiN cantilever tip, employing the photocatalytic effect of
titanium dioxide. When the titanium dioxide immersed in REFERENCES
a solution including gold ions is subject to optical expo- 1. H. S. Nalwa, Ed., “Encyclopedia of Nanoscience and Nanotechnol-
sure, the excited electrons in the conduction band reduce ogy.” American Scientific Publisher, Stevenson Ranch CA, 2003, in
gold ions into gold metal. Illumination by an evanescent press.
wave generated with a total reflection configuration limits 2. J. H. Fendler, Ed., “Nanoparticles and Nanostructured Films.”
the deposition region to the very tip. In the experiments, Wiley-VCH, Weinheim, 1998.
100–300 nm gold particles on SiN cantilever tips for atomic 3. R. W. Siegel, in “Physics of New Materials” (E. Fujita, Ed.), p. 66.
Springer, Berlin, 1998.
force microscopes were obtained. In a related vein, photo-
4. G. L. Timp, Ed., “Nanotechnology.” Springer, New York, 1999.
induced deposition of copper on nanocrystalline TiO2 films 5. V. A. Shchukin and D. Bimberg, Rev. Mod. Phys. 71, 1125 (1999).
was proposed [455]. 6. H. S. Nalwa, Ed., “Handbook of Nanostructured Materials and
Solar photocatalytic oxidation processes (PCO) for degra- Nanotechnology,” Vols. 1–5. Academic, San Diego, 2000.
dation of water and air pollutants have received increasing 7. C. Binns, Surf. Sci. Rep. 44, 1 (2001).
attention. In fact, some field-scale experiments have demon- 8. J. Y. Ying, Ed., “Nanostructured Materials,” Academic, San Diego,
strated the feasibility of using a semiconductor (TiO2  in solar 2001.
collectors and concentrators to completely mineralize organic 9. M. Köhler, “Nanotechnologie.” Wiley-VCH, Weinheim, 2001.
contaminants in water and air [456]. Although successful prein- 10. H. S. Nalwa, Ed., “Nanostructured Materials and Nanotechnology.”
Academic, San Diego, 2001.
dustrial solar tests have been carried out, there are still dis-
11. J. M. Thomas and W. J. Thomas, “Principles and Practice of Hete-
crepancies and doubt concerning process fundamentals such rogeneous Catalysis.” VCH, Weinheim, 1997.
as the roles of active components, appropriate modelling of 12. J. H. Sinfelt, Surf. Sci. 500, 923 (2002).
reaction kinetics, or quantification of photo-efficiency. Chal- 13. F. Zaera, Surf. Sci. 500, 947 (2002).
lenges to development are catalyst deactivation, slow kinet- 14. J. M. Greenberg, Surf. Sci. 500, 793 (2002).
ics, low photo-efficiency and unpredictable mechanisms. The 15. D. A. Williams and E. Herbst, Surf. Sci. 500, 823 (2002).
development of specific nonconcentrating collectors for detox- 16. E. Zinner, Ann. Rev. Earth Planet. Sci. 26, 147 (1998).
ification and the use of additives such as peroxydisulfate have 17. L. Grossman, Geochim. Cosmochim. Acta 36, 597 (1972).
made competitive use of solar PCO possible. 18. B. Mason and L. G. Berry, “Elements of Mineralogy.” Freeman, San
Francisco, 1968.
19. V. E. Henrich and P. A. Cox, “The Surface Science of Metal Oxides.”
Cambridge University Press, Cambridge, 1996.
GLOSSARY 20. B. Levy, J. Electroceram. 1, 239 (1997).
Charge transfer The transfer of a charge carrier (electron or 21. M. A. Fox and M. T. Dulay, Chem. Rev. 93, 341 (1993).
hole) from an excited semiconductor to an adsorbed species on 22. A. L. Linsebigler, G. Q. Lu, and J. T. Yates, Chem. Rev. 95, 735 (1995).
23. A. Fuijshima and K. Honda, Nature 238, 37 (1972).
its surface. This transfer may initiate a reaction (oxidation or
24. A. Fuijshima, K. Kobayakawa, and K. Honda, J. Electrochem. Soc.
reduction) in the adsorbed molecule. 122, 1487 (1975).
Dye-sensitized semiconductor Adsorbing a suitable dye on 25. A. Mills and S. LeHunte, J. Photochem. Photobiol. A: Chem. 108, 1
the surface of a wide band gap semiconductor (like TiO2 ) can (1997).
530 Nanocrystalline TiO2 for Photocatalysis

26. A. Fujishima and T. N. Rao, Pure Appl. Chem. 70, 2177 (1998). 65. R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima,
27. A. Fujishima, K. Hashimoto, and T. Watanabe, “TiO2 Photocatalysis: A. Kitamura, M. Shimohigoshi, and T. Watanabe, Nature 388, 431
Fundamentals and Applications.” BKC, Tokyo, 1999. (1997).
28. A. J. Bard, J. Phys. Chem. 86, 172 (1982). 66. R. Wang, N. Sakai, A. Fujishima, T. Watanabe, and K. Hashimoto, J.
29. M. Grätzel, Ed., “Energy Resources Through Photochemistry and Phys. Chem. B 103, 2188 (1999).
Catalysis.” Academic, New York, 1983. 67. C. Anderson and J. A. Bard, J. Phys. Chem. 99, 9882 (1995); ibid. 101,
30. E. Pelizzetti and M. Schiavello, Eds., “Photochemical Conversion 2611 (1997).
and Storage of Solar Energy.” Kluwer Academic Publishers, Dor- 68. H. Tada, Y. Kubo, M. Akazawa, and S. Ito, Langmuir 14, 2936 (1998).
drecht, 1991. 69. A. Hagfeldt and M. Grätzel, Chem. Rev. 95, 49 (1995).
31. M. Schiavello, Ed., “Photocatalysis and Environment.” Kluwer Aca- 70. T. Gerfin, M. Grätzel, and L. Walder, Prog. Inorgan. Chem. 44, 345
demic Publishers, Dordrecht, 1988. (1997).
32. D. F. Ollis and H. Al-Ekabi, Eds., “Photocatalytic Purification and 71. M. Grätzel, Nature 414, 338 (2001).
Treatment of Water and Air.” Elsevier, Amsterdam, 1993. 72. E. Becquerel, Compt. Rend. 9, 145 (1839).
33. J. Peral, X. Domènech, and D. F. Ollis, J. Chem. Technol. Biotechnol. 73. U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissörtel, J. Salbeck,
70, 117 (1997). H. Spreitzer, and M. Grätzel, Nature 395, 583 (1998).
34. J. H. Fendler, J. Phys. Chem. 89, 2730 (1985). 74. S. Södergren, A. Hagfeldt, J. Olsson, and S. E. Lindquist, J. Phys.
35. M. Pessarakli, Ed., “Handbook of Photosynthesis.” Marcel Dekker, Chem. 98, 5552 (1994).
New York, 1997. 75. G. K. Boschloo and A. Goossens, J. Phys. Chem. 100, 19489 (1996).
36. K. Kalayanasundaram, M. Gräzel, and E. Pelizzetti, Coord. Chem. 76. G. K. Boschloo, A. Goossens, and J. Schoonman, J. Electroanal.
Rev. 69, 57 (1986). Chem. 428, 25 (1997).
37. B. A. Perkinson and M. T. Spitlet, Electrochem. Acta 37, 943 (1992). 77. D. Cahen, G. Hodes, M. Grätzel, J. F. Guillemoles, and I. Riess, J.
38. B. O’Regan and M. Grätzel, Nature 353, 737 (1991). Phys. Chem. B 104, 2053 (2000).
39. H. Weller, Angew. Chem. Int. Ed. Engl. 32, 41 (1993). 78. L. Kavan, K. Kratochvilová, and M. Grätzel, J. Electroanal. Chem.
40. L. N. Lewis, Chem. Rev. 93, 2693 (1993). 394, 93 (1995).
41. A. Henglein, J. Phys. Chem. 97, 5457 (1993). 79. P. E. de Jong and D. Vanmaekelbergh, J. Phys. Chem. B 101, 2716
42. C. Bechinger, E. Wirth, and P. Leiderer, Appl. Phys. Lett. 68, 2843 (1997).
(1996). 80. D. Vanmaekelbergh and P. E. de Jongh, J. Phys. Chem. B 103, 747
43. S. N. Frank and A. J. Bard, J. Am. Chem. Soc. 99, 303 (1977). (1999).
44. S. N. Frank and A. J. Bard, J. Phys. Chem. 81, 1484 (1977). 81. G. Schlichthörl, S. Y. Huang, J. Sprague, and A. J. Frank, J. Phys.
45. M. R. Hoffmann, S. T. Martin, W. Choi, and D. W. Bahnemann, Chem. B 101, 8141 (1997).
Chem. Rev. 95, 69 (1995). 82. S. Y. Huang, G. Schlichthörl, A. J. Nozik, M. Grätzel, and A. J.
46. A. Fujishima, T. N. Rao, and D. A. Tryk, J. Photochem. Photobiol. C: Frank, J. Phys. Chem. B 101, 2576 (1997).
Photochem. Rev. 1, 1 (2000). 83. G. Franco, J. Gehring, L. M. Peter, E. A. Ponomarev, and I. Uhlen-
47. J. C. Ireland, P. Klostermann, E. W. Rice, and R. M. Clark, Appl. dorf, J. Phys. Chem. B 103, 692 (1999).
Environ. Microbiol. 59, 1668 (1993). 84. G. K. Boschloo and D. Fitzmaurice, J. Phys. Chem. B 103, 2228
48. J. C. Sjogren and R. A. Sierka, Appl. Environ. Microbiol. 60, 344 (1999).
(1994). 85. B. O’Regan, M. Grätzel, and D. Fitzmaurice, Chem. Phys. Lett. 183,
49. R. X. Cai, Y. Kubota, T. Shuin, H. Sakai, K. Hashimoto, and 89 (1991).
A. Fujishima, Cancer Res. 52, 2346 (1992). 86. A. Kay, R. Humphry-Baker, and M. Grätzel, J. Phys. Chem. 98, 952
50. N. B. Jackson, C. M. Wang, Z. Luo, J. Schwitzgebel, J. G. Ekerdt, J. (1994).
R. Brock, and A. Heller, J. Electrochem. Soc. 138, 3660 (1991). 87. R. J. Ellingson, J. B. Asbury, S. Ferrere, H. N. Ghosh, J. R. Sprague,
51. H. Gerischer and A. Heller, J. Electrochem. Soc. 139, 113 (1992). T. Lian, and A. J. Nozik, J. Phys. Chem. B 102, 6455 (1998).
52. E. C. Butler and A. P. Davis, J. Photochem. Photobiol. A: Chem. 70, 88. J. B. Asbury, R. J. Ellingson, H. N. Ghosh, S. Ferrere, A. J. Nozik,
273 (1993). and T. Lian, J. Phys. Chem. B 103, 3110 (1999).
53. Y. Luo and D. F. Ollis, J. Catal. 163, 1 (1996). 89. W. S. Struve, “Fundamentals of Molecular Spectroscopy.” Wiley,
54. L. R. Skubal, N. K. Meshkov, and M. C. Vogt, J. Photochem. Photo- New York, 1989.
biol. A: Chem. 148, 103 (2002). 90. P. W. Atkins, “Molecular Quantum Mechanics,” 2nd ed. Oxford Uni-
55. C. Wang, J. Rabani, D. W. Bahnemann, and J. K. Dohrmann, J. Pho- versity Press, Oxford, 1983.
tochem. Photobiol. A: Chem. 148, 169 (2002). 91. Y. Nosaka and M. A. Fox, J. Phys. Chem. 92, 1893 (1988).
56. R. Gao, J. Stark, D. W. Bahnemann, and J. Rabani, J. Photochem. 92. S. T. Martin, H. Herrmann, and M. R. Hoffmann, Trans. Faraday
Photobiol. A: Chem. 148, 387 (2002). Soc. 90, 3315 and 3323 (1994).
57. K. J. Klabunde, J. Stark, O. Koper, C. Mohs, D. G. Park, S. Decker, 93. M. Grätzel, “Heterogeneous Photochemical Electron Transfer.”
Y. Jiang, I. Lagadic, and D. Zhang, J. Phys. Chem. 100, 12142 (1996). CRC Press, Boca Raton, FL, 1989.
58. K. N. P. Kumer, K. Keizer, A. J. Burggraaf, T. Okubo, H. Nagamoto, 94. P. Y. Yu and M. Cardona, “Fundamentals of Semiconductors.”
and S. Morooka, Nature 358, 48 (1992). Springer, Berlin, 2001.
59. M. Gopal, W. J. Moberly Chan, and L. C. De Jonghe, J. Mater. Sci. 95. G. J. Kavernos and N. J. Turro, Chem. Rev. 86, 401 (1986).
32, 6001 (1997). 96. P. V. Kamat, Chem. Rev. 93, 267 (1993).
60. S. Ito, S. Inoue, H. Kawada, M. Hara, M. Iwasaki, and H. Tada, 97. W. J. Albery and P. N. Bartlett, J. Electrochem. Soc. 131, 315 (1984).
J. Colloid Interface Sci. 216, 59 (1999). 98. H. Lüth, “Solid Surfaces, Interfaces and Thin Films.” Springer,
61. K. Terabe, K. Kato, H. Miyazaki, S. Yamaguchi, A. Imai, and Berlin, 2001.
Y. Iguchi, J. Mater. Sci. 29, 1617 (1994). 99. L. E. Brus, J. Chem. Phys. 80, 4403 (1984).
62. K. S. Suslick, Science 247, 1439 (1990). 100. A. Henglein, J. Phys. Chem. 86, 2291 (1982).
63. H. Tada, K. Teranishi, Y. Inubushi, and S. Ito, Chem. Comm. 2345 101. D. Duonghong, J. Ramsden, and M. Grätzel, J. Am. Chem. Soc. 104,
(1998). 2977 (1982).
64. S. Kambe, K. Murakoshi, T. Kitamura, Y. Wada, S. Yanagida, 102. P. V. Kamat, in “Handbook of Nanostructured Materials and Nano-
H. Kominami, and Y. Kera, Solar Energy Mater. Solar Cells 61, 427 technology” (H. S. Nalwa, Ed.), Vol. 3, p. 291. Academic, San Diego,
(2000). 2000.
Nanocrystalline TiO2 for Photocatalysis 531

103. N. S. Lewis, Annu. Rev. Phys. Chem. 42, 543 (1991). 145. N. Uekawa, J. Kajiwara, K. Kakegawa, and Y. Sasaki, J. Colloid Interf.
104. A. Henglein, Chem. Rev. 89, 1861 (1989). Sci. 250, 285 (2002).
105. M. L. Steigenwald and L. E. Brus, Acc. Chem. Res. 23, 183 (1990). 146. S. Yin, Y. Inoue, S. Uchida, Y. Fujishiro, and T. Sato, J. Mater. Res.
106. L. Banyai and S. W. Koch, “Semiconductor Quantum Dots.” World 13, 844 (1998).
Scientific, Rivers Edge, NJ, 1993. 147. C.-C. Wang and J. Y. Ying, Chem. Mater. 11, 3113 (1999).
107. L. Jacak, P. Hawrylak, and A. Wójs, “Quantum Dots.” Springer, 148. J. Yu, X. Zhao, and Q. Zhao, Thin Solid Films 379, 7 (2000).
Berlin, 1998. 149. J. Yu and X. Zhao, Mater. Res. Bull. 35, 1293 (2000).
108. L. E. Brus, J. Chem. Phys. 79, 5566 (1983). 150. N. Negishi, K. Takeuchi, T. Ibusuki, and A. K. Datye, J. Mater. Sci.
109. H. M. Schmidt and H. Weller, Chem. Phys. Lett. 129, 615 (1986). Lett. 18, 515 (1999).
110. M. G. Bawendi, M. L. Steigenwald, and L. E. Brus, Annu. Rev. Phys. 151. N. Negishi and K. Takeuchi, Thin Solid Films 392, 249 (2001).
Chem. 41, 477 (1990). 152. P. Sawunyama, A. Yasumori, and K. Okada, Mater. Res. Bull. 33, 795
111. H. Tang, H. Berger, P. E. Schmid, and F. Lévy, Solid State Comm. 92, (1998).
267 (1994). 153. C. K. Chan, J. F. Porter, Y.-G. Li, W. Guo, and C.-M. Chan, J. Am.
112. H. Tang, F. Lévy, H. Berger, and P. E. Schmid, Phys. Rev. B 52, 7771 Ceram. Soc. 82, 566 (1999).
(1995). 154. A. V. Vorontsov, A. A. Altynnikov, E. N. Savinov, and E. N. Kurkin,
113. G. Mie, Ann. Phys. 25, 377 (1908). J. Photochem. Photobiol., A: Chem. 144, 193 (2001).
114. C. F. Bohren and D. R. Huffman, “Absorption and Scattering of 155. S. Yin, R. Li, Q. He, and T. Sato, Mater. Chem. Phys. 75, 76 (2002).
Light by Small Particles.” Wiley, New York, 1983. 156. X.-S. Feng, S.-Z. Kang, H.-G. Liu, and J. Mu, Thin Solid Films 352,
115. U. Kreibig and M. Vollmer, “Optical Properties of Metal Clusters.” 223 (1999).
Springer, Berlin, 1995. 157. K. Shimizu, H. Imai, H. Hirashima, and K. Tsukuma, Thin Solid
116. U. Kreibig and C. v. Fragstein, Z. Phys. 224, 307 (1969). Films 351, 220 (1999).
117. M. M. Alvarez, J. T. Khoury, T. G. Schaaff, M. N. Shafigullin, I. Vez- 158. S. J. Kim, S. D. Park, C. K. Rhee, W. W. Kim, and S. Park, Scr. Mater.
mar, and R. L. Whetten, J. Phys. Chem. B 101, 3706 (1997). 44, 1229 (2001).
118. G. C. Papavassiliou, Prog. Sol. State Chem. 12, 185 (1979). 159. L. Shi, C. Li, H. Gu, and D. Fang, Mater. Chem. Phys. 62, 62 (2000).
119. P. V. Kamat, Prog. Inorgan. Chem. 44, 273 (1997). 160. R. W. Siegel, in “Materials Science and Technology” (R. W. Cahn,
120. G. Hodes, Isr. J. Chem. 33, 95 (1993). Ed.), Vol. 15, p. 583. VCH, Weinheim, 1991.
121. G. J. Meyer and P. C. Searson, Interface 23–27 (1993). 161. K. Takagi, T. Makimoto, H. Hiraiwa, and T. Negishi, J. Vac. Sci.
122. P. V. Kamat, Chemtech, June, 22–28 (1995). Technol. A 19, 2931 (1999).
123. N. Uekawa, T. Suzuki, S. Ozeki, and K. Kaneko, Langmuir 8, 1 (1992). 162. P. Zeman and S. Takabayashi, Surf. Coat. Technol. 153, 93 (2002).
124. D. V. Paranjape, M. Sastry, and P. Ganguly, Appl. Phys. Lett. 63, 18 163. P. Zeman and S. Takabayashi, J. Vac. Sci. Technol. A 20, 388 (2002).
(1993). 164. W. Zhang, Y. Li, and F. Wang, J. Mater. Sci. Technol. (China) 18, 101
125. K. Vinodgopal, U. Stafford, K. A. Gray, and P. V. Kamat, J. Phys. (2002).
Chem. 98, 6797 (1994). 165. Y. Hou, D. Zhuang, D. Zhao, J. Zhang, and C. Wang, in “Advanced
126. A. Hagfeldt, S. E. Lindquist, and M. Grätzel, Sol. Energy Mater. Sol. Photonic Sensors: Technology and Applications.” Proc. SPIE, Vol.
Cells 32, 245 (1994). 4220, 2000, p. 34.
127. N. A. Kotov, F. C. Meldrum, and J. H. Fendler, J. Phys. Chem. 98, 166. O. Banakh, P. E. Schmid, R. Sanjines, and F. Levy, Surf. Coat. Technol.
8827 (1994). 151–152, 272 (2002).
128. X.-Z. Ding, Z.-Z. Qi, and Y.-Z. He, J. Mater. Sci. Lett. 14, 21 (1995). 167. T. Wang, H. Wang, P. Xu, X. Zhao, Y. Liu, and S. Chao, Thin Solid
129. L. Su and Z. Lu, Spectrochim. Acta 53A, 1719 (1997). Films 334, 103 (1998).
130. X. Z. Ding and X. H. Liu, Mater. Sci. Eng. A 224, 210 (1997). 168. C.-C. Ting, S.-Y. Chen, and D.-M. Liu, J. Appl. Phys. 88, 4628 (2000).
131. D. Robert and J. V. Weber, J. Mater. Sci. Lett. 18, 97 (1999). 169. S. Takeda, S. Suzuki, H. Odaka, and H. Hosono, Thin Solid Films
132. H. Kozuka, Y. Takahashi, G. Zhao, and T. Yoko, Thin Solid Films 392, 338 (2001).
358, 172 (2000). 170. H. Ohsaki, Y. Tachibana, A. Mitsui, T. Kamiyama, and Y. Hayashi,
133. L. Znaidi, R. Seraphimova, J. F. Bocquet, C. Colbeau-Justin, and C. Thin Solid Films 392, 169 (2001).
Pommier, Mater. Res. Bull. 36, 811 (2001). 171. S. G. Springer, P. E. Schmid, R. Sanjines, and E. Levy, Surf. Coat.
134. C. J. Brinker and G. W. Scherer, “Sol-Gel Science.” Academic, San Techn. 151–152, 51 (2002).
Diego, 1990. 172. J. A. Ayllon, A. Figueras, S. Garelik, L. Spirkova, J. Durand, and L.
135. L. C. Klein, Ed., “Sol-Gel Optics: Processing and Applications.” Cot, J. Mater. Sci. Lett. 18, 1319 (1999).
Kluwer, Boston, 1994. 173. V. G. Bessergenev, I. V. Khmelinskii, R. J. F. Pereira, V. V. Krisuk,
136. K. Vinodgopal, S. Hotchandani, and P. V. Kamat, J. Phys. Chem. 97, A. E. Turgambaeva, and I. K. Igumenov, Vacuum 64, 275 (2002).
9040 (1993). 174. Y. C. Zhu and C. X. Ding, J. Eur. Ceram. Soc. 20, 127 (2000).
137. M.-P. Zheng, M.-Y. Gu, Y.-P. Jin, and G.-L. Jin, J. Mater. Sci. Lett. 175. Y. C. Zhu and C. X. Ding, Nanostruct. Mater. 11, 319 (1999).
20, 485 (2001). 176. A. Goossens and J. Schoonman, Eur. J. Sol. State Inorgan. Chem. 32,
138. J. Livage, M. Henry, and C. Sanchez, Prog. Solid State. Chem. 18, 259 779 (1995).
(1988). 177. B. Major, R. Ebner, P. Zieba, and W. Wolzynski, Appl. Phys. A 69,
139. J. Moser, S. Punchihewa, P. P. Infelta, and M. Grätzel, Langmuir 7, 921 (1999).
3012 (1991). 178. T. Deguchi, K. Imai, H. Matsui, M. Iwasaki, H. Tada, and S. Ito, J.
140. H. Kominami, J. Kato, S. Murakami, Y. Kera, M. Inoue, T. Inui, and Mater. Sci. 36, 4723 (2001).
B. Ohtani, J. Mol. Catal. A 144, 165 (1999). 179. H. Wang, T. Wang, and P. Xu, J. Mater. Sci., Mater. Electron. 9, 327
141. S. Ito, S. Inoue, H. Kawada, M. Hara, M. Iwasaki, and H. Tada, (1998).
J. Colloid Interface Sci. 216, 59 (1999). 180. M. Terashima, N. Inoue, S. Kashiwabara, and R. Fujimoto, Appl.
142. T. Sugimoto, H. Itoh, and T. Mochida, J. Colloid Interf. Sci. 205, 42 Surf. Sci. 169–170, 535 (2001).
(1998). 181. S. K. Zheng, T. M. Wang, C. Wang, and G. Xiang, Nucl. Instrum.
143. T. Sugimoto, X. Zhou, and A. Muramatsu, J. Colloid Interf. Sci. 252, Methods B 187, 479 (2002).
339 (2002). 182. S. K. Zheng, T. M. Wang, W. C. Hao, and R. Shen, Vacuum 65, 155
144. T. Sugimoto and X. Zhou, J. Colloid Interf. Sci. 252, 347 (2002). (2002).
532 Nanocrystalline TiO2 for Photocatalysis

183. X. Liu, C. Liang, H. Wang, X. Yang, L. Lu, and X. Wang, Mater. Sci. 220. H. Parala, A. Devi, R. Bhakta, and R. A. Fischer, J. Mater. Chem. 12,
Eng. A 326, 235 (2002). 1625 (2002).
184. H. Yamashita, Y. Ichihashi, M. Takeuchi, S. Kishiguchi, and 221. B. R. Weinberger and R. B. Garber, Appl. Phys. Lett. 66, 2409 (1995).
M. Anpo, J. Synchrotron Radiat. 6, 451 (1999). 222. M. M. Gómez, J. Lu, E. Olsson, A. Hagfeldt, and C. G. Granqvist,
185. K. Oyoshi, N. Sumi, I. Umezu, R. Souda, A. Yamazaki, H. Haneda, Sol. Energy Mater. Sol. Cells 64, 385 (2000).
and T. Mitsuhashi, Nucl. Instrum. Methods B 168, 221 (2000). 223. M. Suzuki, T. Ito, and Y. Taga, Appl. Phys. Lett. 78, 3968 (2001).
186. Z. Xu, J. Shang, C. Liu, C. Kang, H. Guo, and Y. Du, Mater. Sci. Eng. 224. P. Hoyer, Adv. Mater. 8, 857 (1996).
B 63, 211 (1999). 225. P. Hoyer, Langmuir 12, 1411 (1996).
187. T. Ihara, M. Miyoshi, M. Ando, S. Sugihara, and Y. Iriyama, J. Mater. 226. B. B. Lakshimi, C. J. Patrissi, and C. R. Martin, Chem. Mater. 9, 2544
Sci. 36, 4201 (2001). (1997).
188. K. Furusawa, K. Takahashi, S.-H. Cho, H. Kumagai, K. Midorikawa, 227. Y. J. Song and Z. L. Wang, Adv. Mater. 11, 469 (1999).
and M. Obara, J. Appl. Phys. 87, 1604 (2000). 228. M. Adachi, Y. Murata, M. Harada, and S. Yoshikawa, Chem. Lett.
189. V. Subramanian, E. Wolf, and P. V. Kamat, J. Phys. Chem. B 105, 942 (2000).
11439 (2001). 229. M. Zhang, Y. Bando, and K. Wada J. Mater. Sci. Lett. 20, 167 (2001).
190. T. Ohno, K. Fujihara, S. Saito, and M. Matsumura, Sol. Energy Mater. 230. A. Michailowski, D. AlMawlawi, G. Cheng, and M. Moskovits,
Sol. Cells 45, 169 (1997). Chem. Phys. Lett. 349, 1 (2001).
191. M. Tsujimoto, S. Moriguchi, S. Isoda, T. Kobayashi, and T. Komatsu, 231. Y. C. Zhu and C. X. Ding, Nanostruct. Mater. 11, 427 (1999).
J. Electron Microsc. 48, 361 (1999). 232. J. Augustynski, J. Electrochim. Acta 38, 43 (1993).
192. J. Yu, J. C. Yu, B. Cheng, and X. Zhao, J. Sol-Gel Sci. Technol. 24, 233. J. K. Burdett, Inorgan. Chem. 24, 2244 (1985).
39 (2002). 234. J. K. Burdett, T. Hughbands, J. M. Gordon, J. W. Richardson, and J.
193. H. Wang, P. Xu, and T. Wang, J. Vac. Sci. Technol. B 19, 645 (2001). V. Smith, J. Am. Chem. Soc. 109, 3639 (1987).
194. Y. Wang, H. Cheng, Y. Hao, J. Ma, W. Li, and S. Cai, Thin Solid 235. A. Fahmi, C. Minot, B. Silvi, and M. Causa, Phys. Rev. B 47, 11717
Films 349, 120 (1999). (1993).
195. H. Deng, Z. Lu, Y. Shen, H. Mao, and H. Xu, Chem. Phys. 231, 95 236. X.-F. Yu, N.-Z. Wu, Y.-C. Xie, and Y.-Q. Tang, J. Mater. Sci. Lett. 20,
(1998). 319 (2001).
196. A. Dawson and P. V. Kamat, J. Phys. Chem. B 105, 960 (2001). 237. J. Nair, P. Nair, F. Mizukami, Y. Oosawa, and T. Okubo, Mater. Res.
197. D. Beydoun, R. Amal, G. K. C. Low, and S. McEvoy, J. Phys. Chem. Bull. 34, 1275 (1999).
B 104, 4387 (2000). 238. D. G. Rickerby, Philos. Mag. B 76, 573 (1997).
198. T. Kasuga, M. Hiramatsu, M. Hirano, A. Hoson, and K. Oyamada, 239. B. Huber, H. Gnaser, and C. Ziegler, Anal. Bioanal. Chem. (2003).
J. Mater. Res. 12, 607 (1997). 240. P. A. Cox, F. W. H. Dean, and A. A. Williams, Vacuum 33, 839 (1983).
199. S.-H. Hahn, D.-J. Kim, S.-H. Oh, E.-J. Kim, and S.-W. Kim, in “Proc. 241. C. R. Henry, Surf. Sci. Rep. 31, 231 (1998).
5th Korea-Russia International Symp. on Science and Technology,” 242. H.-J. Freund, Faraday Discuss. 114, 1 (1999).
p. 337. IEEE, Piscataway, 2001. 243. U. Diebold, in “The Chemical Physics of Solid Surfaces, Oxide Sur-
200. F. P. Getton, V. A. Self, J. M. Ferguson, J. G. Leadley, P. A. Sermon, faces,” Vol. 9. (D. P. Woodruff, Ed.), Elsevier, Amsterdam, 2001.
and M. Montes, Ceram. Trans. 81, 355 (1998). 244. G. Lu, A. Linsebigler, and J. T. Yates, J. Phys. Chem. 98, 11733 (1994).
201. R. van Grieken, J. Aguado, M. J. López-Muñoz, and J. Marugán, J. 245. S. Fischer, A. W. Munz, K.-D. Schierbaum, and W. Göpel, Surf. Sci.
Photocem. Photobiol. A: Chem. 148, 315 (2002). 337, 17 (1995).
202. T. Tanaka, K. Teramura, T. Yamamoto, S. Takenaka, S. Yoshida, and 246. D. A. Bonnell, Prog. Surf. Sci. 57, 187 (1998).
T. Funabiki, J. Photocem. Photobiol. A: Chem. 148, 277 (2002). 247. U. Diebold, J. Lehman, T. Mahmoud, M. Kuhn, G. Leonardelli,
203. B. Pal, M. Sharon, and G. Nogami, Mater. Chem. Phys. 59, 254 (1999). W. Hebenstreit, M. Schmid, and P. Varga, Surf. Sci. 411, 137 (1998).
204. J. Yang, D. Li, X. Wang, X. Yang, and L. Lu, J. Solid State Chem. 165, 248. M. Li, W. Hebenstreit, U. Diebold, M. A. Henderson, and D. R.
193 (2002). Jennison, Faraday Discuss. 114, 245 (1999).
205. M. E. Rincon, A. Jimenez, A. Orihuela, and G. Martinez, Sol. Energy 249. R. A. Bennett, P. Stone, and M. Bowker, Faraday Discuss. 114, 267
Mater. Sol. Cells 70, 163 (2001). (1999).
206. K. Kato and K.-I. Niihara, Thin Solid Films 298, 76 (1997). 250. Y. Iwasawa, H. Onishi, K. Fukui, S. Suzuki, and T. Sasaki, Faraday
207. A. P. Hong, D. W. Bahnemann, and M. R. Hoffmann, J. Phys. Chem. Discuss. 114, 259 (1999).
91, 6245 (1987). 251. T. Fujino, M. Katayama, K. Inudzuka, T. Okuno, and K. Oura, Appl.
208. K. B. Dhanalakshmi, S. Latha, S. Anandan, and P. Maruthamuthu, Phys. Lett. 79, 2716 (2001).
Intern. J. Hydrogen Energy 26, 669 (2001). 252. I. M. Brookes, C. A. Muryn, and G. Thornton, Phys. Rev. Lett. 87,
209. S. Wenfeng and A. Yoshida, Sol. Energy Mater. Sol. Cells 69, 189 266103-1 (2001).
(2001). 253. R. Schaub, P. Thostrup, N. Lopez, E. Lægsgaard, I. Stensgarrd, J. K.
210. H. N. Ghosh and S. Adhikari, Langmuir 17, 4129 (2001). Nørskov, and F. Besenbacher, Phys. Rev. Lett. 87, 266104-1 (2001).
211. Y. Lei, L. D. Zhang, G. W. Meng, G. H. Li, X. Y. Zhang, C. H. Liang, 254. G. S. Herman, M. R. Sievers, and Y. Gao, Phys. Rev. Lett. 84, 3354
W. Chen, and S. X. Wang, Appl. Phys. Lett. 78, 1125 (2001). (2000).
212. T. Akita, K. Tanaka, K. Okuma, T. Koyanagi, and M. Haruta, J. Electr. 255. R. Hengerer, B. Bolliger, M. Erbudak, and M. Grätzel, Surf. Sci. 460,
Microsc. 50, 473 (2001). 162 (2000).
213. G. L. Li, G. H. Wang, and J. M. Hong, J. Mater. Res. 14, 3346 (1999). 256. Y. Liang, S. Gan, and S. A. Chambers, Phys. Rev. B 63, 235402-1
214. T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, and K. Niihara, Lang- (2001).
muir 14, 3160 (1998). 257. M. Lazzeri and A. Selloni, Phys. Rev. Lett. 87, 266105-1 (2001).
215. H. Imai, Y. Takei, K. Shimizu, M. Matsuda, and H. Hirashima, 258. A. Vittadini, A. Selloni, F. P. Rotzinger, and M. Grätzel, Phys. Rev.
J. Mater. Chem. 9, 2971 (1999). Lett. 81, 2954 (1998).
216. T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, and K. Niihara, Adv. 259. R. E. Tanner, Y. Liang, and E. I. Altman, Surf. Sci. 506, 251 (2002).
Mater. 11, 1307 (1999). 260. A. El-Azab, S. Gan, and Y. Liang, Surf. Sci. 506, 93 (2002).
217. G. H. Du, Q. Chen, R. C. Che, Z. Y. Yuan, and L. M. Peng, Appl. 261. X. Peng, J. Wickham, and A. P. Alivisatos, J. Am. Chem. Soc. 120,
Phys. Lett. 79, 3702 (2001). 5343 (1998).
218. A. Chemseddine and T. Moritz, Eur. J. Inorg. Chem. 235 (1999). 262. K. S. Hamad, R. Roth, J. Rockenberger, T. van Buuren, and A. P.
219. J. C. Yu, J. Yu, W. Ho, and L. Zhang, Chem. Comm. 1942 (2001). Alivisatos, Phys. Rev. Lett. 83, 3474 (1999).
Nanocrystalline TiO2 for Photocatalysis 533

263. N. A. Hill and K. B. Whaley, J. Chem. Phys. 100, 2831 (1994). 298. J. Maier, Prog. Solid State Chem. 23, 171 (1995).
264. Z. Y. Wu, J. Zhang, K. Ibrahim, D. C. Xian, G. Li, Y. Tao, T. D. Hu, 299. J. A. S. Ikeda and Y.-M. Chiang, J. Am. Ceram. Soc. 76, 2437 (1993).
S. Bellucci, A. Marcelli, Q. H. Zhang, L. Gao, and Z. Z. Chen, Appl. 300. J. F. Baumard and E. Tani, J. Chem. Phys. 67, 857 (1977).
Phys. Lett. 80, 2973 (2002). 301. D. M. Smyth, Prog. Solid State Chem. 15, 145 (1984).
265. L. X. Chen, T. Rajh, Z. Y. Wang, and M. Thurnauer, J. Phys. Chem. 302. K. Hoshino, N. L. Peterson, and C. L. Wiley, J. Phys. Chem. Solids
B 101, 10688 (1997). 46, 1397 (1985).
266. A. P. Xagas, E. Androulaki, A. Hiskia, and P. Falaras, Thin Solid 303. F. Millot, M.-G. Blanchin, R. Tetot, J.-F. Marucco, B. Poumellec, C.
Films 357, 173 (1999). Picard, and B. Touzelin, Prog. Solid State Chem. 17, 263 (1987).
267. A. Provata, P. Falaras, and A. Xagas, Chem. Phys. Lett. 297, 484 304. A. Bernasik, M. Rekas, M. Sloma, and W. Weppner, Solid State Ionics
(1998). 72, 12 (1994).
268. J. Y. Ying, L. F. Chi, H. Fuchs, and H. Gleiter, Nanostruct. Mater. 3, 305. D. Mardare, M. Tasca, M. Delibas, and G. I. Rusu, Appl. Surf. Sci.
273 (1993). 156, 200 (2000).
269. E. Comini, G. Sberveglieri, M. Ferroni, V. Guidi, C. Frigeri, and D. 306. D. Mardare, C. Baban, R. Gavrila, M. Modreanu, and G. I. Rusu,
Boscarino, J. Mater. Res. 16, 1559 (2001). Surf. Sci. 507–510, 468 (2002).
270. R. I. Bickley, T. Gonzalez-Carreno, J. S. Lees, L. Palmisano, and R. 307. M. Gómez, J. Rodríguez, S. Tingry, A. Hagfeldt, S.-E. Lindquist, and
J. D. Tilley, J. Sol. State Chem. 92, 178 (1991). C. G. Granqvist, Sol. Energy Mat. Sol. Cells 59, 277 (1999).
271. P. E. de Jongh and D. Vanmaekelbergh, Phys. Rev. Lett. 77, 3427 308. H. Tang, K. Prassad, R. Sanjinès, P. E. Schmid, and F. Lévy, J. Appl.
(1996). Phys. 75, 2042 (1994).
272. A. Zaban, A. Meier, and B. A. Gregg, J. Phys. Chem. B 101, 7985 309. R. Sanjinès, H. Tang, H. Berger, F. Gozzo, G. Margaritondo, and F.
(1997). Lévy, J. Appl. Phys. 75, 2945 (1994).
273. A. Solbrand, H. Lindström, H. Rensmo, A. Hagfeldt, and 310. H. Wittmer, S. Holten, H. Kliem, and H. D. Breuer, Phys. Stat. Sol.
S. Lindquist, J. Phys. Chem. B 101, 2514 (1997). (A) 181, 461 (2000).
274. J. Nelson, Phys. Rev. B 59, 15 374 (1999). 311. V. I. Makarov, S. A. Kochubei, and I. Khmelinskii, Chem. Phys. Lett.
275. R. Könenkamp, R. Henninger, and P. Hoyer, J. Phys. Chem. 97, 7328 355, 504 (2002).
(1993). 312. K. Wilke and H. D. Breuer, Z. Phys. Chem. 213, 135 (1999).
276. R. Könenkamp and R. Henninger, Appl. Phys. A 58, 87 (1994). 313. M. Anpo, Sol. Energy Mater. Sol. Cells 38, 221 (1995).
277. R. Könenkamp, Phys. Rev. B 61, 11057 (2000). 314. M. A. Fox, Sol. Energy Mater. Sol. Cells 38, 381 (1995).
278. F. Cao, G. Oskam, G. J. Meyer, and P. C. Searson, J. Phys. Chem. 100, 315. V. Augugliaro, V. Loddo, L. Palmisano, and M. Schiavello, Sol. Energy
17 021 (1996). Mater. Sol. Cells 38, 411 (1995).
279. P. M. Sommeling, H. C. Rieffe, J. M. Kroon, J. A. M. van Roosmalen, 316. D. Bahnemann, D. Bockelmann, and R. Goslich, Sol. Energy Mater.
A. Schonecker, and W. C. Sinke, in “Proceedings of the 14th EC Pho- 24, 564 (1991).
tovoltaic Solar Energy Conference” (H. Ossenbrink, P. Helm, and 317. D. M. Blake, J. Webb, C. Turchi, and K. Magrini, Sol. Energy Mater.
H. Ehmann, Eds.), pp. 1816–1819. H. S. Stephens and Associates, 24, 584 (1991).
Bedford, 1997. 318. J. Yu, X. Zhao, and Q. Zhao, J. Mater. Sci. Lett. 19, 1015 (2000).
280. A. Solbrand, H. Lindstrom, H. Rensmo, A. Hagfeldt, S.-E. Lindquist, 319. S. Ruan, F. Wu, T. Zhang, W. Gao, B. Xu, and M. Zhao, Mater. Chem.
and S. Sodergren, J. Phys. Chem. B 101, 2514 (1997). Phys. 69, 7 (2001).
281. K. Schwarzburg and F. Willig, Appl. Phys. Lett. 58, 2520 (1991). 320. K. Baba and R. Hatada, Surf. Coat. Technol. 136, 241 (2001).
282. B. Levy, W. Liu, and S. E. Gilbert, J. Phys. Chem. B 101, 1810 (1997). 321. T. Sumita, T. Yamaki, S. Yamamoto, and A. Miyashita, Jpn. J. Appl.
283. L. Dloczik, O. Ileperuma, I. Lauermann, L. M. Peter, E. A. Pono- Phys., Part 1 40, 4007 (2001).
marev, G. Redmond, N. J. Shaw, and I. Uhlendorf, J. Phys. Chem. B 322. L. Gao and Q. Zhang, Scr. Mater. 44, 1195 (2001).
101, 10 281 (1997). 323. A. J. Maira, K. L. Yeung, J. Soria, J. M. Coronado, C. Belver, C. Y.
284. A. Hagfeldt, S.-E. Lindquist, and M. Grätzel, Sol. Energy Mater. Sol. Lee, and V. Augugliaro, Appl. Catal. B: Environ. 29, 327 (2001).
Cells 32, 245 (1994). 324. K. Y. Jung, S. B. Park, and S.-K. Ihm, Appl. Catal. A: Gen. 224, 229
285. S. A. Haque, Y. Tachibana, D. R. Klug, and J. R. Durrant, J. Phys. (2002).
Chem. B 102, 1745 (1998). 325. M. S. Jeon, T. K. Lee, D. H. Kim, H. Joo, and H. T. Kim, Sol. Energy
286. C. J. Barbe, F. Arendse, P. Compte, M. Jirousek, F. Lenzmann, Mater. Sol. Cells 57, 217 (1999).
V. Shklover, and M. Grätzel, J. Am. Ceram. Soc. 80, 3157 (1997). 326. G.-J. Chee, Y. Nomura, K. Ikebukuro, and I. Karube, Sens. Actuators
287. F. Cao, G. Oskam, P. C. Searson, J. M. Stipkala, T. A. Heimer, B 80, 15 (2001).
F. Farzad, and G. J. Meyer, J. Phys. Chem. 99, 11 974 (1995). 327. Y. Xu, L. Jiang, X. Lu, J. Wang, and D. Zhou, High Technol. Letters
288. A. Goossens, B. van der Zanden, and J. Schooman, Chem. Phys. Lett. (China) 6, 63 (2000).
331, 1 (2000). 328. J. Sheng, L. Shivalingappa, J. Karasawa, and T. Fukami, J. Mater. Sci.
289. M. Takahashi, K. Tsukigi, T. Uchino, and T. Yoko, Thin Solid Films 34, 6201 (1999).
388, 231 (2001). 329. B. Kim, D. Byun, J. K. Lee, and D. Park, Jpn. J. Appl. Phys., Part 1
290. Th. Dittrich, E. A. Lebedev, and J. Weidmann, Phys. Stat. Sol. (A) 41, 222 (2002).
165, R5 (1998). 330. H. Tada and M. Tanaka, Langmuir 13, 360 (1997).
291. J. Weidmann, Th. Dittrich, E. Konstantinova, I. Lauermann, 331. C.-G. Wu, L.-F. Tzeng, Y.-T. Kuo, and C. H. Shu, Appl. Catal. A: Gen.
I. Uhlendorf, and F. Koch, Sol. Energy Mater. Sol. Cells 56, 153 (1999). 226, 199 (2002).
292. T. Dittrich, J. Weidmann, F. Koch, I. Uhlendorf, and I. Lauermann, 332. A. Rachel, B. Lavedrine, M. Subrahmanyam, and P. Boule, Catal.
Appl. Phys. Lett. 75, 3980 (1999). Comm. 3, 165 (2002).
293. V. Duzhko, V. Yu. Timoshenko, F. Koch, and Th. Dittrich, Phys. Rev. 333. R. L. Pozzo, J. L. Giombi, M. A. Baltanás, and A. E. Cassano, Appl.
B 64, 075204 (2001). Catal. B: Environ. 38, 61 (2002).
294. V. Kytin and Th. Dittrich, Phys. Stat. Sol. (A) 185, 461 (2001). 334. Y. Paz, Z. Luo, L. Rabenberg, and A. Heller, J. Mater. Res. 10, 2842
295. V. Kytin, V. Duzhko, V. Yu. Timoshenko, J. Rappich, and Th. Dittrich, (1995).
Phys. Stat. Sol. (A) 185, R1 (2001). 335. M. Nakamura, L. Sirghi, T. Aoki, and Y. Hatanaka, Surf. Sci. 507–
296. V. Kytin, Th. Dittrich, F. Koch, and E. Lebedev, Appl. Phys. Lett. 79, 510, 778 (2002).
108 (2001). 336. T. Watanabe, A. Nakajima, R. Wang, M. Minabe, S. Koizumi,
297. P. Knauth and H. L. Tuller, J. Appl. Phys. 85, 897 (1999). A. Fujishima, and K. Hashimoto, Thin Solid Films 351, 260 (1999).
534 Nanocrystalline TiO2 for Photocatalysis

337. M. Miyauchi, N. Kieda, S. Hishita, T. Mitsuhashi, A. Nakajima, 378. Q. Dai and J. Rabani, J. Photochem. Photobiol. A: Chem. 148, 17
T. Watanabe, and K. Hashimoto, Surf. Sci. 511, 401 (2002). (2002).
338. V. Romeas, P. Pichat, C. Guillard, T. Chopin, and C. Lehaut, J. Phys. 379. C. G. Garcia, C. J. Kleverlaan, C. A. Bignozzi, and N. Y. M. Iha, J.
IV, Proc. 9, PR3/247 (1999). Photochem. Photobiol. A: Chem. 147, 143 (2002).
339. K. Miyashita, S. Kuroda, T. Ubukata, T. Ozawa, and H. Kubota, 380. J. E. Moser, P. Bonhoete, and M. Grätzel, Coord. Chem. Rev. 171,
J. Mater. Sci. 36, 3877 (2001). 245 (1998).
340. K. Miyashita, S. Kuroda, T. Sumita, and T. Ubukata, J. Mater. Sci. 381. G. Centi, P. Ciambelli, S. Perathoner, and P. Russo, Catal. Today 75,
Lett. 20, 2137 (2001). 3 (2002).
341. J. C. Yu, J. Yu, W. Ho, and J. Zhao, J. Photochem. Photobiol. A: 382. R. Abe, K. Sayama, K. Domen, and H. Arakawa, Chem. Phys. Lett.
Chem. 148, 331 (2002). 344, 339 (2001).
342. J. C. Yu, J. Yu, and J. Zhao, Appl. Catal. B: Environ. 36, 31 (2002). 383. M. Ashokkumar, Int. J. Hydrog. Energy 23, 427 (1998).
343. Y. Ohko, S. Saitoh, T. Tatsuma, and A. Fujishima, J. Electrochem. 384. M. Anpo, H. Yamashita, Y. Ichihashi, Y. Fujii, and M. Honda, J. Phys.
Soc. 148, B24 (2001). Chem. B 101, 2632 (1997).
344. D. W. Bahnemann, M. Hilgendorff, and R. Memming, J. Phys. Chem. 385. M. Anpo, Nuovo Cimento 19D, 1641 (1997).
B 101, 4265 (1997). 386. P. G. Smirniotis, D. A. Peña, and B. S. Uphade, Angew. Chem. 113,
345. D. W. Goodman, Surf. Rev. Lett. 2, 9 (1995). 2537 (2001).
346. C. T. Campbell, Surf. Sci. Rep. 27, 1 (1997). 387. S. Daito, F. Tochikubo, and T. Watanabe, Jpn. J. Appl. Phys., Part 1
347. M. Bäumer and H.-J. Freund, Prog. Surf. Sci. 61, 127 (1999). 40, 2475 (2001).
348. H.-J. Freund, Surf. Sci. 500, 271 (2002). 388. S. B. Kim and S. C. Hong, Appl. Catal. B: Environ. 35, 305 (2002).
349. S. Sato and J. M. White, Chem. Phys. Lett. 72, 83 (1980). 389. P. B. Amama, K. Itoh, and M. Murabayashi, Appl. Catal. B: Environ.
350. D. Hufschmidt, D. Bahnemann, J. J. Testa, C. A. Emilio, and M. I. 37, 321 (2002).
Litter, J. Photochem. Photobiol. A: Chem. 148, 223 (2002). 390. A. Bouzaza and A. Laplanche, J. Photochem. Photobiol. A: Chem.
351. U. Siemon, D. Bahnemann, J. J. Testa, D. Rodríguez, M. I. Litter, 150, 207 (2002).
N. Bruno, J. Photochem. Photobiol. A: Chem. 148, 247 (2002). 391. B. D˛abrowski, A. Zaleska, M. Janczarek, J. Hupka, and J. D. Miller,
352. F. B. Li and X. Z. Li, Appl. Catal. A: Gen. 228, 15 (2002). J. Photochem. Photobiol. A: Chem. 151, 201 (2002).
353. M. Daté, Y. Ichihashi, T. Yamashita, A. Chiorino, F. Boccuzzi, and 392. J. Araña, O. González Díaz, M. Miranda Saracho, J. M. Doña
M. Haruta, Catal. Today 72, 89 (2002). Rodríguez, J. A. Herrera Melián, and J. Pérez Peña, Appl. Catal. B:
354. V. Vamathevan, R. Amal, D. Beydoun, G. Low, S. McEvoy, J. Pho- Environ. 36, 113 (2002).
tochem. Photobiol. A: Chem. 148, 233 (2002). 393. D. W. Bahnemann, S. N. Kholuiskaya, R. Dillert, A. I. Kulak, and A.
355. D. Dvoranová, V. Brezová, M. Mazúra, and M. A. Malati, Appl. I. Kokorin, Appl. Catal. B: Environ. 36, 161 (2002).
Catal. B: Environ. 37, 91 (2002). 394. A. Rachel, M. Sarakha, M. Subrahmanyam, and P. Boule, Appl. Catal.
356. H. Yamashita, M. Harada, J. Misaka, M. Takeuchi, K. Ikeue, and M. B: Environ. 37, 293 (2002).
Anpo, J. Photochem. Photobiol. A: Chem. 148, 257 (2002). 395. A. Rachel, M. Subrahmanyam, and P. Boule, Appl. Catal. B: Environ.
357. A. Di Paola, E. García-López, S. Ikeda, G. Marcì, B. Ohtani, and L. 37, 301 (2002).
Palmisano, Catal. Today 75, 87 (2002). 396. J. A. Byrne, A. Davidson, P. S. M. Dunlop, and B. R. Eggins, J. Pho-
358. A.-W. Xu, Y. Gao, and H.-Q. Liu, J. Catal. 207, 151 (2002). tochem. Photobiol. A: Chem. 148, 365 (2002).
359. R. Vogel, K. Pohl, and H. Weller, Chem. Phys. Lett. 174, 241 (1999). 397. M. Bekbolet, A. S. Suphandag, and C. S. Uyguner, J. Photochem.
360. S. Kohtani, A. Kudo, and T. Ssakata, Chem. Phys. Lett. 206, 166 Photobiol. A: Chem. 148, 121 (2002).
(1993). 398. Y. Cho and W. Choi, J. Photochem. Photobiol. A: Chem. 148, 129
361. R. Vogel, P. Hoyer, and H. Weller, J. Phys. Chem. 98, 3183 (1994). (2002).
362. D. Liu and P. V. Kamat, J. Phys. Chem. 97, 10769 (1993). 399. S. Parra, J. Olivero, and C. Pulgarin, Appl. Catal. B: Environ. 36, 75
363. K. Vinodgopal and P. V. Kamat, Environ. Sci. Technol. 29, 841 (1995). (2002).
364. I. Bedja and P. V. Kamat, J. Phys. Chem. 99, 9182 (1995). 400. M. Muneer and D. Bahnemann, Appl. Catal. B: Environ. 36, 95
365. R. Q. Long and R. T. Yang, J. Catal. 207, 158 (2002). (2002).
366. G. Colón, M. C. Hidalgo, and J. A. Navio, Appl. Catal. A: Gen. 231, 401. S. Parra, S. Malato, and C. Pulgarin, Appl. Catal. B: Environ. 36, 131
185 (2002). (2002).
367. M. K. Nazeeruddin, A. Kay, I. Rodicio, B. R. Humphry, E. Mueller, 402. E. Vulliet, C. Emmelin, J.-M. Chovelon, C. Guillard, and J.-M. Her-
P. Liska, N. Vlachopoulos, and M. Grätzel, J. Am. Chem. Soc. 115, rmann, Appl. Catal. B: Environ. 38, 127 (2002).
6382 (1993). 403. H. Einaga, S. Futamura, and T. Ibusuki, Appl. Catal. B: Environ. 38,
368. F. Willig, R. Eichberger, N. S. Sundaresan, and B. A. Parkinson, J. 215 (2002).
Am. Chem. Soc. 112, 2702 (1990). 404. M. C. Blount and J. L. Falconer, Appl. Catal. B: Environ. 39, 39 (2002).
369. P. V. Kamat, S. Das, K. G. Thomas, and M. V. George, Chem. Phys. 405. K.-I. Shimizu, T. Kaneko, T. Fujishima, T. Kodama, H. Yoshida, and
Lett. 178, 75 (1991). Y. Kitayama, Appl. Catal. A: Gen. 225, 185 (2002).
370. R. Eichberger and F. Willig, Chem. Phys. 141, 159 (1990). 406. S. Vijaikumar, N. Somasundaram, C. Srinivasan, Appl. Catal. A: Gen.
371. B. Burfeindt, T. Hannappel, W. Storck, and F. Willig, J. Phys. Chem. 225, 129 (2002).
100, 16463 (1996). 407. M. Kang, Appl. Catal. B: Environ. 37, 187 (2002).
372. N. J. Cherepy, G. P. Smestad, M. Grätzel, and J. Z. Zhang, J. Phys. 408. B. Ohtani, K. Iwai, H. Kominami, T. Matsuura, Y. Kera, and
Chem. B 101, 9342 (1997). S. Nishimoto, Chem. Phys. Lett. 242, 315 (1995).
373. T. Hannappel, B. Burfeindt, W. Storck, and F. Willig, J. Phys. Chem. 409. S. Horikoshi, N. Watanabe, H. Onishi, H. Hidaka, and N. Serpone,
B 101, 6799 (1997). Appl. Catal. B: Environ. 37, 117 (2002).
374. J. Randy, R. J. Ellingson, J. B. Asbury, S. Ferrere, H. N. Ghosh, J. R. 410. N. San, A. Hatipoǧlu, G. Koçtürk, and Z. Çýnar, J. Photochem. Pho-
Sprague, T. Lian, and A. J. Nozik, J. Phys. Chem. B 102, 6455 (1998). tobiol. A: Chem. 146, 189 (2002).
375. A. Furube, T. Asahi, H. Masuhara, H. Yamashita, and M. Anpo, J. 411. V. Iliev, J. Photochem. Photobiol. A: Chem. 151, 195 (2002).
Phys. Chem. B 103, 3120 (1999). 412. T. Nakashima, Y. Ohko, D. A. Tryk, A. Fujishima, J. Photochem.
376. H. Al-Ekabi, B. Butters, D. Delany, J. Ireland, N. Lewis, T. Powell, Photobiol. A: Chem. 151, 207 (2002).
and J. Story, Trace Met. Environ. 3, 321 (1993). 413. G. Mele, G. Ciccarella, G. Vasapollo, E. García-López, L. Palmisano,
377. V. Iliev and D. Tomova, Catal. Comm. 3, 287 (2002). and M. Schiavello, Appl. Catal. B: Environ. 38, 309 (2002).
Nanocrystalline TiO2 for Photocatalysis 535

414. V. Durgakumari, M. Subrahmanyam, K. V. Subba Rao, A. Ratna- 435. X. Z. Li and Y. G. Zhao, Water Sci. Technol. 39, 249 (1999).
mala, M. Noorjahan, and K. Tanaka, Appl. Catal. A: Gen. 248, 155 436. J. E. Pacheco, M. R. Prairie, and L. Yellowhorse, Trans. ASME, J.
(2002). Sol. Energy Engin. 115, 123 (1993).
415. R. W. Matthews and S. R. McEvoy, Sol. Energy 49, 507 (1992). 437. S. M. Rodriguez, C. Richter, J. B. Galvez, and M. Vincent, Sol. Energy
416. L. Zhang, T. Kanki, N. Sano, and A. Toyoda, Sol. Energy 70, 331 56, 401 (1996).
(2001). 438. D. C. Schmelling, K. A. Gray, and P. V. Kamat, Water Res. 31, 1439
417. J. Grzechulska and A. W. Morawski, Appl. Catal. B: Environ. 36, 45 (1997).
(2002). 439. J.-S. Hur and Y. Koh, Biotechnol. Lett. 24, 23 (2002).
418. J. Li, C. Chen, J. Zhao, H. Zhu, and J. Orthman, Appl. Catal. B: 440. R. L. Ziolli and W. F. Jardim, J. Photochem. Photobiol. A: Chem. 147,
Environ. 37, 331 (2002). 205 (2002).
419. Z. Sun, Y. Chen, Q. Ke, Y. Yang, and J. Yuan, J. Photochem. Photo- 441. M. Perez, F. Torrades, J. Peral, C. Lizama, C. Bravo, S. Casas, J. Freer,
biol. A: Chem. 149, 169 (2002). and H. D. Mansilla, Appl. Catal. B: Environ. 33, 89 (2001).
420. M. Sökmen and A. Özkan, J. Photochem. Photobiol. A: Chem. 147, 442. M. Hamerski, J. Grzechulska, and A. W. Morawski, Sol. Energy 66,
77 (2002). 395 (1999).
421. S. Sakthivel, M. V. Shankar, M. Palanichamy, B. Arabindoo, and V. 443. M. Bekbolet, M. Lindner, D. Weichgrebe, and D. W. Bahnemann,
Murugesan, J. Photochem. Photobiol. A: Chem. 148, 153 (2002). Sol. Energy 56, 455 (1996).
422. S. Al-Qaradawi and S. R. Salman, J. Photochem. Photobiol. A: Chem. 444. C. Wei, W. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L.
148, 161 (2002). Smith, K. Rajeshwar, Environ. Sci. Technol. 28, 934 (1994).
423. T. Sauer, G. Cesconeto Neto, H. J. José, R. F. P. M. Moreira, J. Pho- 445. M. Bekbölet, Water Sci. Technol. 11–12, 95 (1997).
tochem. Photobiol. A: Chem. 149, 147 (2002). 446. P. S. M. Dunlop, J. A. Byrne, N. Manga, and B. R. Eggins, J. Pho-
424. H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard, tochem. Photobiol. A: Chem. 148, 355 (2002).
and J.-M. Herrmann, Appl. Catal. B: Environ. 39, 75 (2002). 447. J. Wist, J. Sanabria, C. Dierolf, W. Torres, and C. Pulgarin, J. Pho-
425. H. Zhan and H. Tian, Dyes Pigments 37, 231 (1998). tochem. Photobiol. A: Chem. 147, 241 (2002).
426. H. Zhan, K. Chen, and H. Tian, Dyes Pigments 37, 241 (1998). 448. I. Liu, L. A. Lawton, B. Cornish, and P. K. J. Robertson, J. Photochem.
427. H. Zhan, H. Tian, K. Chen, and W. Zhu, Toxicol. Environ. Chem. 69, Photobiol. A: Chem. 148, 349 (2002).
531 (1999). 449. R. Armon, N. Laot, N. Narkis, and I. Neeman, J. Adv. Oxid. Technol.
428. S. Malato, J. Blanco, A. Vidal, and C. Richter, Appl. Catal. B: Environ. 3, 145 (1998).
37, 1 (2002). 450. M. Z. Atashbar, in “Proceedings 1st IEEE Conference on Nano-
429. T. An, G. Li, Y. Xiong, X. Zhu, H. Xing, and G. Liu, Mater. Phys. technology.” p. 544. IEEE-NANO 2001, IEEE, Piscataway, NJ,
Mech. 4, 101 (2001). 2001.
430. L. R. Skubal, N. K. Meshkov, T. Rajh, and M. Thurnauer, J. Pho- 451. L. W. Miller, M. I. Tejedor, B. P. Nelson, and M. A. Anderson, J. Phys.
tochem. Photobiol. A: Chem. 148, 393 (2002). Chem. B 103, 8490 (1999).
431. L. R. Skubal and N. K. Meshkov, J. Photochem. Photobiol. A: Chem. 452. H. Honda, A. Ishizahi, R. Soma, K. Hashimoto, and A. Fujishima, J.
148, 211 (2002). Illum. Eng. Soc. (USA) 27, 42 (1998).
432. Y. Ming, C. R. Chenthamarakshan, and K. Rajeshwar, J. Photochem. 453. J. Mizuguchi, J. Electrochem. Soc. 148, J55 (2001).
Photobiol. A: Chem. 147, 199 (2002). 454. T. Okamoto and I. Yamaguchi, J. Microsc. 202, 100 (2001).
433. S. G. Schrank, H. J. José, and R. F. P. M. Moreira, J. Photochem. 455. Z. V. Saponjic, T. Rajh, J. M. Nedeljkovic, and M. C. Thurnauer,
Photobiol. A: Chem. 147, 71 (2002). Mater. Sci. Forum 352, 91 (2000).
434. J. Aguado, R. van Grieken, M. J. López-Munoz, and J. Marugán, 456. M. Romero, J. Blanco, B. Sanchez, A. Vidal, S. Malato, A. I. Carbona,
Catal. Today 75, 95 (2002). and E. Garcia, Sol. Energy 66, 169 (1999).

Вам также может понравиться