Você está na página 1de 32

Encyclopedia of

Nanoscience and
Nanotechnology

www.aspbs.com/enn

Nanocrystalline TiO2 for Photocatalysis


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

CONTENTS
1. Introduction
2. Electronic and Charge-Transfer Processes
in Photocatalysis
3. Preparation of Nanostructured Materials
and Thin Films
4. Structural Properties of Nanocrystalline
TiO2 Films
5. Electrical Properties of Nanocrystalline
TiO2 Films
6. Photocatalytic Properties
of Nanocrystalline TiO2
7. Photocatalytic Applications
of Nanocrystalline TiO2
Glossary
References

1. INTRODUCTION
The development of novel materials and the assessment of
their potential application constitutes a major fraction of
todays scientic reasearch efforts. In fact, there exist various major governmental research and development programs related to nanostructured materials. Furthermore, it
is estimated that nanotechnology has grown into a multibillion dollar industry and may become the most dominant single technology of the twenty-rst century. To allow
for this fact, this encyclopedia [1] encompasses a series
of contributions devoted to a very prominent eld of current materials research activities, namely, nanoscience and
nanotechnology. The importance of these developments is
reected also in a number of recent books and articles
reviewing this rapidly evolving eld [210].
This article focuses on a specic class of such novel
nano-scaled materials, nanocrystalline TiO2 , and its photocatalytic properties. The title of this article encompasses
three main terms ((photo)catalysis, nanocrystalline, and
TiO2 ) which, individually, stand for very important areas
of scientic research and of, perhaps even more important,
technological applications. Their synergistic combination, as

ISBN: 1-58883-062-4/$35.00
Copyright 2004 by American Scientic Publishers
All rights of reproduction in any form reserved.

indicated by the present theme, has stimulated great hopes


in accomplishing thereby achievements with paramount benets for human beings and the global environment. To outline the present state of that quest is the major goal of this
article.
Catalysis is probably the most familiar of the three
terms mentioned. A catalyst is incorporated in essentially
everybodys automobile, with the goal of reducing or even
eliminating the engines toxic gaseous components by converting them into less harmful (albeit not necessarily benign)
substances. As is the case in all catalytic reactions, the catalyst itself is not part of the reaction, but is expected to
enhance its rate, that is, the velocity of the transformation
from the original components (the educts in the chemists
terminology) into the nal ones (the products). Hence, a
catalyst is an entity that accelerates a chemical reaction without being consumed itself in the process. Without catalysts,
various chemical reactions of great importance would proceed too slowly [11]. The economic signicance of catalysis
is enormous. In the U.S. alone, the annual value of products
manufactured with the use of catalysts is roughly in the vicinity of one trillion dollars [12]. Indeed, more than 80% of the
industrial chemical processes in use nowadays rely on one or
more catalytic reactions [13]. A number of those, including
oil rening, petrochemical processing, and the manufacturing of commodity chemicals (olens, methanol, ethylene glycol, etc.), are already well established. But many others, as
will be seen in this contribution, represent challenges requiring the development of entirely new approaches. But apart
from their industrial importance, catalytic phenomena effect
virtually all aspects of our lives. They are crucial in many
processes occurring in living things, where enzymes are the
catalysts. They are important in the processing of foods and
the production of medicines. The reader may have noticed
that we have as yet refrained from specifying the meaning
of photocatalysis; which will be one of the major topics of
this article. This term refers to a catalytic process that is
triggered by illuminating the system by visible light or ultraviolet irradiation. Ideally, that light ux would be the suns
radiance.
Next we shall consider the meaning of nanocrystalline.
First, it is noted that in todays science world rather
inationary used, the prex nano refers to a fraction of

Encyclopedia of Nanoscience and Nanotechnology


Edited by H. S. Nalwa
Volume 6: Pages (505535)

506
one part in one billion (109  and, hence, its correct usage
would require it being connected to some kind of unit (e.g.,
of length, time, energy, mass, etc.). In the present context
(and in that of nanotechnology), nano most often relates
to the dimension, that is, the size of an object. Therefore,
nanocrystalline in the ensuing discussions will designate particles (of crystalline structure and, primarily, with the chemical composition of titanium dioxide) whose typical sizes are
in the range of a few to several nanometers (nm), that is, of
the order of the one billionth part of one meter. Obviously,
these are extremely tiny objects and can be seen and studied only with the help of sophisticated analytical instruments
like an electron microscope.
At rst glance, it may appear that such tiny particles are
a rather modern contrivance, but this is probably a premature conclusion. In fact, it is quite rmly established that
nm-sized particles (mostly very refractory ones like corundum, diamond, or silicon carbide) are ubiquitous in the universe [14] and that they were already present at the time
and the location of the formation of the solar system. This
stardust originated from stellar outows and supernova
ejecta, which may have occurred eons before the gas and
dust condensed into what is now the sun, the earth, and the
planets. In fact, this dust has intrigued astronomers since
the days of William Herschel who noted, in the 1780s, the
existence of small regions in the sky where there appeared
to be a complete absence of stars [15]. These regions are
most easily seen against the rich star-elds of the Milky Way.
Evidence of the presolar origin of these nanocrystalline particles comes primarily from their isotopic abundance pattern [16], which deviates typically to such an extent from
any other known matter that a terrestrial or solar origin
is virtually impossible. (Most of these particles that have
been investigated were extracted from primitive meteorites
in which they were incorporated during the formation stage
of the solar system; these did not experience any later modication and, hence, preserved the presolar dust particles
unaltered [17].)
Only now, some billions of years later, mankind has initiated the manufacture and application of such nanocrystalline materials. Nanostructured materials with crystal sizes
in the range of 550 nm of a variety of materials, including metals and ceramics, have been articially synthesized by
many different techniques in the past couple of years [2, 3,
57]. Such new ultrane-grained materials have properties
that are often signicantly different and greatly enhanced
as compared to coarser-grained or bulk substances. These
favorable changes in properties result generally from their
small grain sizes, the large percentage of atoms in grain
boundaries and at surfaces, the large surface-to-bulk ratio,
and the interaction between individual crystallites. Since
these features can be tailored to a considerable extent, during synthesis and processing, such nanophase materials are
thought to have great technological potential even beyond
their current applications.
Let us nally turn to a brief discussion of the third term,
TiO2 ( i.e., titanium dioxide). TiO2 has three different crystal structures [18]: rutile, anatase, and brookite; only the former two of them are commonly used in photocatalysis. Like
for many other metal oxides (also for titanium oxide) have
the respective structural, optical, and electronic properties

Nanocrystalline TiO2 for Photocatalysis

been elucidated through several decades of intense scientic


research (for a review see, e.g., [19]); some of them will be
referred to in the course of the present overview. The feature probably most important in the present context is the
fact that TiO2 is a semiconductor with a bandgap of 3.2 eV.
On the other hand, TiO2 , in its nanocrystalline form, constitutes an enormously important commercial product. In fact,
the world production of titanium dioxide white pigments
amounts to some 4.5 million tons per annum and the global
consumption may be considered a distinct economic indicator. White pigments of TiO2 have average particle sizes
of around 200300 nm, optimized for the scatter of white
light, resulting, thereby, in a hiding power. Reducing the
crystallite size (to 100 nm), the reectance of visible light
(vis) decreases and the material becomes more transparent;
it is widely employed, for example, in paints, plastics, paper,
or pharmaceuticals. Nanocrystalline TiO2 exhibits, in addition, a pronounced absorption of ultraviolet (UV) radiation.
Because of this high UV absorption and the concurrent high
transparency for visible light, TiO2 particles with a size of
<100 nm have found widespread use in such diverse areas
as sun cosmetics, packaging materials, or wood protection
coatings. Hence, although perhaps not generally realized,
TiO2 is ubiquitous in our everyday life. Apart from this wellestablished range of usage, an increasing number of catalytic
applications of nanocrystalline TiO2 have emerged in recent
years.
At this point, it appears appropriate to return to a more
general view of the present topic. Nanostructured materials
have generally the potential [3] for incorporating and taking
advantage of a number of size-related effects in condensed
matter ranging from electronic effects (so-called quantum
size effects) caused by spatial connement of delocalized
valence electrons and altered cooperative (many-body)
atom phenomena, like lattice vibrations or melting, to the
suppression of such lattice-defect mechanisms as dislocation
generation and migration in conned grain sizes. The possibilities to assemble size-selected atom clusters into new
materials with unique or improved properties will likely create a revolution in our ability to engineer materials with
controlled optical, electronic, magnetic, mechanical, and
chemical properties for many pending future technological
applications.
Among those nanocrystalline materials, semiconductors
appear to play a pivotal role in such distinct elds as [20]:
(i) heterogeneous photocatalysis;
(ii) photoelectrochemistry, including electrochemical
photovoltaic cells;
(iii) photochemistry in zeolites, intercalated materials,
thin lms, and membranes (like self-assembled structures);
(iv) supramolecular photochemistry.
This diversity is thought to be largely due to the fact that
heterogeneously dispersed semiconductor surfaces provide
both a xed environment to inuence the chemical reactivity
of a wide range of adsorbates and, in addition, a means to
initiate light-induced redox reactivity in these weakly associated molecules [21]. Upon photoexcitation of semiconductor
nanoparticles, either in solutions or xed to a suitable substrate, simultaneous oxidation and reduction reactions may

507

Nanocrystalline TiO2 for Photocatalysis

occur; molecular oxygen is often assumed to serve as the oxidizing agent. Such semiconductor-mediated redox reactions
are commonly grouped under the rubric of heterogeneous
photocatalysis [21].
In a heterogeneous photocatalytic system, photo-induced
molecular reactions take place at the surface of the catalyst. Depending on where the initial excitation occurs,
photocatalysis can be generally divided into two classes of
processes [22]: In the case that the initial photoexcitation
occurs in the adsorbate molecule, which then interacts with
the ground state catalyst substrate, the process is referred
to as a catalyzed photoreaction. On the other hand, when
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 sensitized photoreaction [22]. Apparently, a considerable degree
of synergism may be crucial when, for example, a transition
metal oxide photocatalysts is combined with supramolecular
spectral sensitizing ligand complexes used to harvest light
and vectorially transfer photo-generated electrons and holes
along selected energetic pathways.
In 1972, Fujishima and Honda reported [23] the photocatalytic splitting (i.e., the simultaneous oxidizing and reducing)
of water upon illumination of a TiO2 single-crystal electrode;
in analogy to natural photosynthesis, they demonstrated the
photoelectrolysis of water making use of a photoexcited
semiconductor in what was essentially a photochemical battery. In that system, an n-type TiO2 semiconductor electrode, which was connected through an electrical load to a
platinum counter electrode, was exposed to near-UV light
(cf. Fig. 1). When the surface of the TiO2 electrode was
illuminated with light of wavelength shorter than 415 nm,
photocurrent was observed to ow [23, 24]. Furthermore,
oxygen evolution (i.e., an oxidation reaction) on the TiO2
and hydrogen evolution (a reduction) on the Pt electrode
was observed. The photoexcitation of TiO2 injects electrons
from its valence band into its conduction band; the electrons
ow through the external circuit to the Pt cathode where
water molecules are reduced to hydrogen gas and the holes
remain in the TiO2 anode where water molecules are oxidized to oxygen. These observations indicate that water can
be decomposed by means of UV-VIS light according to the

e-

load

e-

e-

H2O

eCB

VB

H2O

e- Pt
TiO2

H2

O2

Figure 1. Schematic arrangement for the photosplitting of water in an


electrochemical cell (in the actual setup, both electrodes are immersed
in an aqueous solution and the chambers are separated by an ionically
conducting porous material). When the TiO2 photoanode is irradiated
with light, O2 evolves from it, whereas H2 evolves from the Pt counterelectrode, while electrons will ow through an external load.

following scheme

TiO2 + 2h 2e + 2h+

electron-hole pair
formation in TiO2

H2 O + 2h+ 1/2O2 + 2H+

reaction at the TiO2


electrode

2H+ + 2e H2

reaction at the
Pt electrode

H2 O + 2h 1/2O2 + H2

overall reaction

It appears to be generally accepted that this discovery


boosted extensive research efforts in the era of heterogeneous photocatalysis [21, 2527]. These studies, often carried out in an interdisciplinary fashion with the participation
of physicists, chemists, and chemical engineers, focused on
issues that are of great relevance both economically as well
as ecologically like energy renewal and storage [2830], the
decomposition of organic compounds in polluted air and
wastewaters [3133], chemical energy generation [34, 35],
and photovoltaic devices [36, 37]. Most of these either
already proven or envisaged applications are intimately
linked to the extraordinary properties of nanocrystalline
TiO2 . In fact, nanocrystalline metal-oxide semiconductors
such as TiO2 have been applied successfully in modern
technologies including solar energy conversion, gas sensors,
catalysis, and photocatalysis [3842].
Following the rst examination in 1977, using TiO2 to
decompose cyanide in water [43, 44], a great deal of effort
has been devoted in recent years to developing heterogeneous photocatalysts with high photocatalytic activities for
their applications in solving environmental cleanup and pollution remediation problems [31, 32, 45, 46]. Photocatalytic
reactions on TiO2 surfaces are very important in such environmental processes, as the oxidation of organic materials
and the reduction of heavy metal ions in industrial waste
waters. Apart from the utilization for water and air purication, 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
clean-up of oil spills [50, 51], and other applications [45]; a
more detailed account will be given later in this article.
As mentioned, TiO2 is a semiconductor with a bandgap
of 3.2 eV; when excited by light of energy equal to or
exceeding that value, electrons are promoted from the
valence band to the conduction band leaving positive holes
in the valence band. These electrons and holes are capable of, respectively, reducing and oxidizing compounds at
the TiO2 surface. If the electrons and holes do not recombine (and produce heat), they can follow various reaction
pathways; it is commonly accepted that the hole is quickly
converted to the hydroxyl radical ( OH) upon oxidation of
surface-adsorbed water and that the hydroxyl radical is the
major reactant responsible for the oxidation of organic compounds. Typical reductive and oxidative reactions could be

508

Nanocrystalline TiO2 for Photocatalysis

the following [25, 5254]


TiO2 + h TiO2 e + h+ 

electron-hole
pair formation

e + Mn+ Mn1+

reduction reaction

h + H2 Oads OH + H

OH + Rads Rads + H2 O

oxidation of
adsorbed water
oxidation of
organic species

where Mn+ is the oxidized compound and Rads the


adsorbed organic species. Because the production of the
hydroxyl radical is considered the decisive step, the determination and optimization of the corresponding quantum yield
in illuminated TiO2 is an important task [55, 56].
According to those concepts, TiO2 nanoparticles are
expected to show a unique surface chemistry due to their
larger surface area [57]. The origin of the distinct photocatalytic activities exhibited by nanoparticles of TiO2 is crucial in understanding the reaction mechanisms, for example,
if they are purely due to the increased surface area or if
they have to be addressed to the presence of distorted sites
on the surface or to the whole lattice of the particles. In
order to commercialize these treatment techniques, it is of
great importance to improve the preparative methods of
titania, because the photocatalytic activity and phase transition behavior of TiO2 are signicantly inuenced by the
preparative conditions [5863]. As previously mentioned,
these catalytic processes constitute a major issue of this work
and will be outlined in the following sections.
The photocatalytic activity of TiO2 is very useful not
only in environmental purication by decomposition of
organic substances, but also in the materials industry such as
mirrors and glasses due to its self-cleaning [64] and antifogging effects. The latter has been attributed to the photoinduced hydrophilic nature of the surface [65, 66]. A further
enhancement of the photocatalytic activity could be effected
by means of composite TiO2 materials; examples would be
metal doping, mixing with insulating substances like SiO2 or
Al2 O3 [67], and monolayer coverage by SiO2 [68].
Another prominent future application of nanocrystalline
semiconductors is thought [6971] to lie in photovoltaics,
that is, the conversion of sunlight into electrical power.
The limited reserves of fossil fuels and the increasing concern of global warming due to the greenhouse effect caused
by the combustion of those fuels has triggered, in the last
decades, widespread efforts into the development of photovoltaic devices. With an energy supply from the sun to the
earth of 3 1024 Joule per year (about 10,000 times the
global annual energy consumption), this enterprise appears
all but unreasonable. In such solar cells, photon incident on
a semiconductor can create electron-hole pairs, basically a
result of the photoelectric effect, discovered by Becquerel
already in 1839 [72]. At a junction between two different
materials, this may establish an electrical potential difference across this interface. Until now, the material of choice
for manufacturing such junctions has been (doped) silicon
(crystalline or amorphous), with compound semiconductors
also being considered more recently. While this traditional

approach clearly has room for further improvements, it may


ultimately face limitations in terms of cost efciency (manufacturing costs per unit of energy produced). Novel materials
and fabrication schemes are therefore explored. A promising approach consists of electrochemical photovoltaic cells
utilizing nanoporous semiconducting electrodes fabricated
by lightly sintering nanosized TiO2 particulates, followed by
spectral sensitization with an appropriate dye.
Metal oxide particles with diameters of some 10 nm
can be prepared as paste and spread out over a surface
of uorine-doped SnO2 conducting glass to form a threedimensional network of interconnected nanoparticles. These
nanostructured metal oxide layers, and in particular those
constructed from anatase titanium dioxide (TiO2 , have
aroused great interest because of their unprecedented properties as electrodes. They nd application in dye-sensitized
solar cells, which nowadays show light-to-electricity conversion efciencies of 10% [69, 7377]. The nature of electron
migration in these electrodes has been debated in past years
as the experimental results and their interpretation does not
converge to a generally accepted model. Instead, the exact
role of electron trapping and the concomitant screening of
the electric eld remain unclear. It is reported that soon
after electrons are injected into the conduction band of TiO2
a large fraction of them get trapped in surface states. Migration of these electrons must then proceed with a hoppingtype process in which the electrons remain most of the time
in localized states [7883]. On the other hand, it is also
reported that the injection of electrons into the conduction
band shows the so-called free-electron absorption, which
extends over a wide spectral range from the visible to the
infrared [8488]. This suggests that not all injected electrons become trapped but that a substantial fraction of them
remain in the conduction band.
This article is organized in the following way: Section 2
outlines the electronic and charge-transfer processes as relevant for photocatalysis, with a special emphasis towards
nanocrystalline TiO2 . Since in photocatalysis and related
applications the respective nanostructured materials are
employed either in colloidal solutions or attached to a suitable support (e.g., as electrodes or thin lms), both of these
preparation techniques are discussed (Section 3). Furthermore, various approaches for surface and thin-lm modication are described, as well as novel deposition methods
and structures. The structural and electronic properties of
nanocrystalline TiO2 lms are examined in Sections 4 and
5, respectively. The photocatalytic properties of nanocrystalline TiO2 constitute the central theme of Section 6, highlighting the dependence of the photocatalytic activity on
different parameters like lm structure and phase, surface
morphology, electronic properties, and the effects induced
by various surface modications. Representative examples
of photocatalytic applications utilizing nanocrystalline TiO2
materials are presented in Section 7. Finally, an extensive
set of references is provided that should be useful for further study: although the number is substantial, no attempt
was made, however, to be comprehensive; any such attempt
might be bound to fail due to the rapidity with which this
eld is evolving.

509

Nanocrystalline TiO2 for Photocatalysis

2. ELECTRONIC
AND CHARGE-TRANSFER
PROCESSES IN PHOTOCATALYSIS

2.1. Electronic Excitations


and Charge-Carrier Trapping
A photocatalytic process is initiated by the absorption of
photons by a molecule or the substrate to produce highly
reactive electronically excited states. The efciency is controlled by the systems light absorption properties. Three
fundamental steps are of relevance: (1) the electronic excitation of a molecule upon photon absorption, (2) the band-gap
excitation of the semiconductor substrate, and (3) the interfacial electron transfer. Since a detailed account of these
processes has been given in a lucid treatise by Linsebigler
et al. [22], we shall briey summarize here only the more
important points, referring thereby partly to that work. The
probability of an electronic transition can be calculated from
quantum mechanical perturbation theory and is proportional to the square of the amplitude of the radiation eld
and the square of the transition dipole moment [89, 90].
The latter may be computed via the molecular wave function which, in turn, depends on the product of the electronic
spatial wave function, the electronic spin wave function, and
the nuclear wave function. Further arguments [89] lead to
some general selection rules in terms of which electronic
transitions are allowed and which might be forbidden. Typically, the excitation of a weak transition will not effectively
induce a photochemical reaction, because few of the incident photons will be absorbed (low cross-section); however,
an overall high reaction rate may still be possible in the case
of a high quantum yield, that is, if the probability of product molecule formation per absorbed photon is high. The
photochemical efciency will also be determined by which
deexcitation channels are dominant. In particular, the pertinent lifetimes of the involved processes are to be considered.
Whereas the absorption of a photon occurs very rapidly on
the order of 1015 s, deexcitation events are much slower,
favoring the decay channel which minimizes the lifetime of
the excited state.
The initial process for heterogeneous photocatalysis of
organic and inorganic compounds by semiconductors is the
generation of electron-hole pairs in the semiconductor particles. Once excitation across the bandgap has occurred,
the lifetime is sufcient (in the nanosecond regime [91])
for the created electron-hole pair to undergo charge transfer to adsorbed species on the semiconductor surface from
solution or gas phase. Figure 2 exemplies these processes.
The enlarged section shows the excitation of an electron
from the valence band to the conduction band initiated
by light absorption with energy equal or greater than the
bandgap of the semiconductor. The gure also illustrates
several deexcitation pathways for the electrons and holes.
The electron transfer to adsorbed species or to the solvent
results from migration of electrons or holes to the surface. At the surface, the semiconductor can donate electrons to reduce an electron acceptor (often oxygen in an
aerated solution), corresponding to pathway c in Figure 2;
conversely, a hole can migrate to the surface where an electron from a donor species can combine with the surface hole

A-

CB

VB

D+

A
Figure 2. Schematic illustration of the photoexcitation in a semiconductor particle followed by deexcitation events. CB and VB designate
the conduction and valence band, respectively. For further details see
text.

oxidizing the donor (pathway d). Competing with charge


transfer to adsorbed species is electron and hole recombination, occurring either in the volume (pathway b) or on
the surface of the semiconductor (pathway a; in both cases
heat will be released. Naturally, electron and hole recombination is detrimental to the efciency of a semiconductor
photocatalyst. Modications to semiconductor surfaces, such
as metal addition, dopants, or combination with materials,
can be benecial in decreasing the recombination and concurrently increasing the quantum yield of the process.
An efcient means to retard the recombination of photoexcited electron-hole pairs may proceed via the trapping
of charge carriers. The occurrence of surface and bulk irregularities resulting from the preparation process is associated
with surface electron states; these may serve as charge carrier traps and can suppress the recombination of electrons
and holes. The charge carriers trapped in such states are
localized to a specic site on the surface or in the bulk; their
population is dependent on the energy difference between
the trap and the bottom of the conduction band. Experimental observations of such trapped states in TiO2 will be
reported later.
On the basis of laser ash photolysis measurements
[45, 92], the characteristic times for the individual steps
occurring during heterogeneous photocatalysis on TiO2 have
been derived. Whereas the primary process of charge-carrier
generation is extremely fast (fs), charge-carrier recombination may occur on time scales of 10100 ns. Charge-carrier
trapping can be very fast (100 ps) for the (reversible) trapping of a conduction-band electron in a shallow trap, but
moderately fast (10 ns) for a deep trap or for the surface
trapping of a valence-band hole, resulting in a surface-bound
hydroxyl radical. Finally, interfacial charge transfer can be
slow (100 ns) for the oxidation of an electron donor or
very slow (ms) for the reduction of an electron acceptor.
In general, the valence-band holes are powerful oxidants
(+10 to +35 V versus NHE depending on the type of semiconductor and pH), while the conduction-band electrons are
good reductants (+0.5 to 1.5 V vs NHE) [93]; most organic
photodegradation reactions utilize the oxidizing power of
these photo-generated holes.

510

Nanocrystalline TiO2 for Photocatalysis

2.2. Band-Edge Position


and Band Bending

semiconductor

The ability of a semiconductor to undergo photo-induced


electron transfer to adsorbed species on its surface is governed by the band energy position of the semiconductor and
the redox potentials of the adsorbate. The relevant potential
of the acceptor must be below (more positive than) the conduction band potential of the semiconductor. By contrast,
the potential of the donor needs to be above (more negative than) the valence-band position of the semiconductor in
order to donate an electron to the vacant hole. The bandedge positions of several semiconductors are depicted in
Figure 3; the internal energy scale is given both with respect
to the vacuum level (left scale) and for comparison to normal hydrogen electrode (NHE) in a solution of an aqueous
electrolyte at pH = 1.
When a semiconductor is brought into contact with
another phase (e.g., liquid, gas, or metal), the transfer
of mobile charges across this junction occurs until electronic equilibrium is reached. This redistribution of charges
involves the formation of a space-charge layer, that is, the
charge distribution on each side of the interface differs
from the bulk material (cf. Fig. 4). Depending on whether
the electrons accumulate or deplete at the semiconductor
side, an accumulation or depletion layer is formed, causing
concurrently a shift in the electrostatic potential and a downward (or upward) bending of the bands in the semiconductor. The depletion of electrons may reach such an extent that
their concentration at the surface decreases below the intrinsic level. As a consequence, the Fermi level is closer to the
valence than to the conduction band; this situation is called
an inversion layer, as the semiconductor is p-type at the
surface and n-type in the bulk. These features are well documented [94] and will not be further discussed here. Of some
interest in the present context is the situation encountered
ENHE [eV]

E [eV]
vacuum

0.0

-2.5
-3.0
-3.5
-4.0

SiC
(n,p)

GaAsP GaP
(n,p) (n,p)
GaAs
(n,p)

CdSe
(n)

CdS
(n)

WO3 SnO2
(n) (n)

TiO2
(n)

-5.0

-6.0
-6.5
-7.0
-7.5
-8.0

-1.5
-1.0

ZnO
(n)

-4.5

-5.5

-2.0

-0.5

Eu2+/3+

0.0

H2/H+

0.5

[Fe(CN)6]3-/4Fe2+/Fe3+

1.0

E=
1.4 2.25 2.25
eV
eV eV 1.7
eV

1.5
3.0
eV

2.5
eV

Ru(bipy)2+/3+

2.0

3.0
3.2
eV

3.2
eV

-8.5

- + - +
- + +
- +
+ + -

E
CB

CB

Ec
EF

Ec
EF

Eref

EV

Eref

EV
VB

VB

(a)

(b)

- +

+ -

+ + -

+
+

+
+

+ + -

E
E
CB

Eref

CB

Eref

Ec
EF

Ec
EF

EV

EV
VB

VB

(c)

(d)

Figure 4. Space-charge layer of an n-type semiconductor in contact


with another phase (e.g., an electrolyte or gas): (a) at band situation,
(b) accumulation layer, (c) depletion layer, and (d) inversion layer.

when an n-type semiconductor like TiO2 is in contact with


an electrolyte as in a photoelectrochemical cell [71]; such
devices are thought to have great potential both in photovoltaics for producing electric current and as fuel cells
for the generation of hydrogen via photo-cleavage of water.
Because of these potential applications, the characteristics
of the semiconductor-electrolyte interface have been investigated in great details [93, 95, 96].
In particular, the potential distribution within a spherical
semiconductor particle could be derived [97] using a linearized PoissonBoltzmann equation. As discussed in [69],
two limiting cases are of special importance for light-induced
electron transfer in semiconductor dispersions. For large
particles, the total potential drop within the particle is
 =

kT
2e

w
LD

2
(1)

Ce4+/3+

2.5
3.2
eV

electrolyte

+ - + + +
- + - +

3.5
4.0

3.8
eV

Figure 3. Bandgap energies of various semiconductors in an aqueous


electrolyte at pH = 1.

where w is the width of the space charge layer and LD =


0 kT /e2 Nd 05 is the Debye screening length [94], which
depends on the dielctric constant, , of the material and on
the number density of ionized dopants, Nd . This potential
variation is identical with that of a planar Schottky depletion layer [98]. For very small semiconductor particles (with
radius R) the potential drop within the particle becomes
 =

kT
6e

R
LD

2
(2)

511

Nanocrystalline TiO2 for Photocatalysis

From the latter equation, it is found that the electrical eld


in nano-sized semiconductors will usually be small and that
high dopant levels are required to produce a signicant
potential difference between the center and the surface. For
example [69], in order to obtain a 50 meV potential drop in a
nanocrystalline TiO2 particle with R = 6 nm, a concentration
of 5 1019 cm3 of ionized donor impurities is necessary.
Undoped TiO2 nanocrystallites have a much lower carrier
concentration and the band bending within the particles is
therefore negligibly small.

2.3. Photo-Induced Charge-Transfer


Processes on the Catalyst Surface
The principle of electron and hole transfer at the photoexcited semiconductor particle has already been alluded to
previously in Fig. 2. Both free and trapped charge carriers participate in these interfacial redox reactions. Due to
the quantization effects, by decreasing the particles size, it
is possible to shift the conduction band to more negative
potentials and the valence band to more positive values. It
was concluded [99] that the shift of the bandgap is proportional to 1/R2 , R being the particle size. Therefore, redox
processes that cannot occur in bulk materials can be facilitated in quantized semiconductor particles. Figure 5 shows
schematically such possible transfer reactions for an adsorbate at the surface of a semiconductor. When there are
accessible energy levels in the substrate, an electron may
be transferred from the donor (D) into a substrate level
and then into the acceptor orbital as shown in Figure 5(a).
This scheme operates in the photosensitization of semiconductor particles by dye molecules. An electron is injected
from the excited state dye molecule into the semiconductor, which then reduces another adsorbate particle. Early
experimental conrmation [100, 101] of these processes used
the reduction of methyl viologen in colloidal semiconductor
systems and the water splitting process. Later, such reductive processes have been investigated for many different systems (see, e.g., [102]); some illustrative examples will be
presented in Section 6.
For an initial excitation on the semiconductor substrate
(Fig. 5(b)), a positively charged hole is created at the bandedge of the valence band. An electron is transferred into
h
CB

CB

CB

D*

VB

VB

D+

AVB

(a)
CB
h
VB

CB

e-

CB

A-

D+

VB

VB

(b)
Figure 5. Sensitized photoreaction with (a) an initial excitation of the
adsorbate, or (b) an initial excitation of the solid.

the empty acceptor orbital and, simultaneously, an electron


is donated from the lled donor level to recombine with the
original hole. This is a very general process for wide bandgap
oxide semiconductors like TiO2 and others. For example,
the oxidation of many organic substances in colloidal suspensions has been investigated [102]. The energetics of the
semiconductor valence band and the oxidation potential of
the redox couple inuence this photocatalytic oxidation. For
example, the enhancement in the efciency of halide oxidation at TiO2 follows the sequence Cl < Br < I , correlating with the decrease in the oxidation potential. Again,
some recent examples will be presented in Section 6.
The kinetic analysis [103] of electron transfer in colloidal
semiconductor systems is often complex. Apart from the
energetics of the conduction band of the semiconductor
and the redox potential of the acceptor, factors such as
the surface charges of the colloids, adsorption of the substrates, participation of surface states, and competition with
charge recombination inuence the rate of charge transfer at
the semiconductor interface [102]. This fact is evident from
the widely differing rates of experimentally observed charge
transfer rates, with time scales ranging from picoseconds to
milliseconds for different experimental conditions and various semiconductor systems.

2.4. Quantum-Size Effects


Size quantization effects in metals or semiconductors
have attracted considerable attention in the past decade
[104107]. Semiconductor nanoparticles may experience a
transition in terms of electronic properties from those typical for a solid to that of a molecule, in which a complete electron delocalization has not yet occurred. These
quantum-size effects arise when the Bohr radius of the rst
exciton (an interacting electron-hole pair) and the semiconductor becomes comparable with or larger than that of the
particle; the Bohr radius [94]
rB = 40  2 /e2 m 

(3)

depends on the dielectric constant  and the effective mass


m of the charge carriers (electrons and holes). The latter is frequently radically different for holes and electrons
and, in some cases, m is more than an order of magnitude smaller than the free-electron mass me . Hence, such
quantum-size effects play a role for crystallites of approximately 110 nm in diameter. Under these conditions, the
energy levels available for the electrons and holes in the conduction and valence bands become discrete and the effective bandgap of the semiconductor is increased. To a rst
approximation, the energy spacing between quantized levels is inversely proportional to the effective mass and the
square of the particle diameter. A schematic energy diagram
resulting from such connement effects is shown in Figure 6.
Several attempts have been carried out to compute the
electronic energy levels in such quantum dots [99, 108110].
According to these concepts, the energy of the lowest excited
state of a semiconductor particle with radius R is given
approximately by


2 2 1
1
18e2
+
(4)

ER = Eg +
2R2 me
mh
R

512

Nanocrystalline TiO2 for Photocatalysis


bulk semiconductor

cluster
LUMO

CB
Eg

shallow
trap

deep trap

deep trap
surface
state

VB
HOMO
Figure 6. Schematic correlation diagram relating bulk-semiconductor
electronic states to quantum crystalline states. Adapted from [110],
M. G. Bawendi et al., Annu. Rev. Phys. Chem. 41, 477 (1990). 1990,
Annual Reviews.

Here Eg is the bandgap of the bulk semiconductor, the second term is the quantum energy of localization, increasing
as R2 for both electron and hole, and the third term is the
Coulomb attraction [99]; whereas the Coulomb term shifts
ER to smaller energy as R, the quantum connement
contribution increases ER as R2 . As a result, the apparent bandgap will always increase for small enough R. But
while the shift can be appreciable (1 eV) for small band
gap materials like InSb, it might be considerably smaller
(0.2 eV) for semiconductors with a large bandgap like
TiO2 or ZnO [99]. Apart from a large effect on the optical properties (e.g., a change in color of the material due
to the blue shift of the absorption), size quantization may
also lead to major changes in the photocatalytic properties.
While these effects have been studied in great detail for several compound semiconductor materials like CdS, ZnO, or
PbS, related data for TiO2 appear to be still rather limited.

2.5. Optical Properties


Semiconductors absorb light below a threshold wavelength
g which is related to the bandgap energy Eg by [93]
g nm = 1240/Eg eV

(5)

Within the solid, the extinction of light follows an exponential law


I = I0 exp
z

(6)

where z is the penetration depth and


is the reciprocal
absorption length. For TiO2 , for example,
= 26 104 cm1
at a wavelength of 320 nm; this implies that such light is
extinguished to 90% after passing a distance of 390 nm in
the semiconductor. Near threshold, the value of
increases
with increasing photon energy. Frequently, a proportionality
of the type

h = Ch Eg n

(7)

provides a satisfactory description of the absorption where


h is the photon energy. C is a constant scaling with the
effective masses of the charge carriers, but is independent
of the photon frequency. The exponent n has a value of
0.5 for a direct semiconductor and 2 for an indirect one
[94]. Experimental data [111, 112] obtained on both anatase
and rutile TiO2 thin lms indicated, however, that the actual
situation might be more complex.

In colloidal solution, semiconductor particles reduce the


light intensity of the incident beam both by scattering and
absorption. In the absence of quantum-size effects the extinction spectrum is described by the Mie theory [113, 114].
This theoretical approach can be applied to an assembly of
spherical particles if the interparticle distance is larger than
the wavelength of the incident light (i.e., the particles scatter independently); if, furthermore, the particle size is much
smaller than the wavelength, the energy-dependent absorption cross-section   for the irradiation of a solution containing the particles can be derived [115117]:
  = 9Vp


s3/2 2
c 1 + 2s 2 + 22

(8)

where   = 1 + i2 is the complex, frequency-dependent


dielectric constant of the semiconductor particle, s is the
dielectric constant of the embedding medium (the solvent),
Vp is the volume of a particle, is the frequency of the
incident light, and c is the velocity of light. For the case
of a dilute system of particles with the number density n,
  can be related to the reciprocal absorption length

 
  = n  . It follows that the imaginary part
of the dielectric constant is a direct measure of the light
absorption by the particles; it increases steeply near the
absorption edge, that is, for close to the threshold frequency. As noted in [69], Mies theory has been widely
employed to interpret the extinction spectra of colloidal systems [118]. For example, the brilliant ruby or yellow colors
caused, respectively, by colloidal gold or silver particles are
distinctly explained by this theory.

3. PREPARATION OF
NANOSTRUCTURED MATERIALS
AND THIN FILMS
Several distinct techniques have been utilized in recent years
to synthesize nanocrystalline TiO2 thin lms by chemical,
electrochemical, and organized assembly methods [69, 119
121]. A simple approach involves casting of the thin lm
directly from colloidal suspensions [122]. This method of
preparation is relatively simple and inexpensive compared
with other existing methods such as chemical vapor deposition or molecular beam epitaxy. Preparation of nanoclusters
in polymer lms and LangmuirBlodgett lms has also been
considered. The solgel technique has been found to be useful in developing nanostructured semiconductor membranes
with either a two-dimensional or three-dimensional conguration. Organic-template-mediated synthesis has been
employed to develop nanoporous materials. The nanostructured materials are highly porous and can easily be surface
modied with sensitizers, redox couples, and/or other nanostructured lms.

3.1. Preparation from


Colloidal Suspensions
Nanostructured semiconductor lms of TiO2 have been prepared frequently from colloidal suspensions [123133]. By
controlling the preparative conditions of the precursor colloids, it is possible to tailor the properties of these lms.

Nanocrystalline TiO2 for Photocatalysis

Typically, these thin TiO2 lms exhibit interesting photochromic, electrochromic, photocatalytic, and photoelectrochemical properties that are inherited from the native
colloids. The synthetic procedure involves the preparation
of ultrasmall particles (diameter 210 nm) in aqueous or
ethanolic solutions by controlled hydrolysis. The colloidal
suspension is coated onto a conducting glass plate (an optically transparent electrode (OTE)) and dried; nally, the
lm is annealed at 200400  C in air for some 12 h.
The conducting surface facilitates direct electrical contact of
the nanostructured thin lm. This simple approach of coating preformed colloids onto a surface and annealing generally produces an oxide lm, which is robust and exhibits an
excellent stability in both acidic and alkaline media, a feature very important in several potential applications. Generally, some optimization is required for thicker lms and for
higher colloidal concentrations in order to avoid cracking
of the lms upon drying. Further details of the methodology of preparation can be found in [134136]. Transmission electron micrographs of nanostructured lms prepared
from colloidal suspensions show a three-dimensional network of oxide nanocrystals of particle diameter as small
as a few nanometers. No signicant aggregation or sintering effects are observed during the annealing process. X-ray
diffraction analysis also conrms the crystallinity of these
nanostructured lms. Composite lms, which in some cases
may exhibit improved properties as compared to singlecomponent lms, can be manufactured by mixing two or
more components prior to casting of the lm.
Titania sol and gel prepared through the hydrolysis of
tetrabutyl titanate in acid water solution are sensitive to
ultraviolet (UV) irradiation and turn into blue color [137].
Electron spin-resonance measurement showed that the photochromism was ascribed to the presence of titanium (III)
ions in the Ti-O-Ti network. The titanium (III) ions could
be oxidized by the oxygen in the atmosphere, and then the
sol and gel faded slowly. The absorption peaks in the optical
absorption spectra of the titania gel were attributed to the
transition of 3-dimensional electrons of a trivalent titanium
in certain environments.
The morphology of TiO2 particles affects their catalytic
activity and electrical properties. In recent years, many
methods for preparing TiO2 nanoparticles and thin lms
have been studied [138]. TiO2 nanocrystals prepared by the
solgel method often have fully hydroxylated surfaces and
these hydroxyl groups have a strong inuence on the catalytic and photocatalytic properties such as electron-transfer
rates and reducing properties [139]. In order to develop
photocatalysts with high activities, it is very important to
prepare porous anatase nanoparticles with a high specic
surface area. Furthermore, the preparation method should
be simple and should be a low temperature process. Mixtures of rutile and anatase precipitates could be obtained
by the hydrothermal treatment of an alcohol solution of Ti
alkoxide at 573 K [140], while anatase nanoparticles were
prepared by heat treatment of a H2 OEtOH solution of
TiOSO4 at 373 K [141]. Thus, the anatase and rutile particles
were usually formed in a solution by conventional soft chemical synthesis methods. The preparation of monodispersed
oxide particles by the solgel method enables the manufacture of oxide particles through a gel state with a regulated

513
particle growth rate [142144]. In a recent study [145],
anatase nanoparticles were prepared in a Ti-peroxy gel without the collapse of the gel during the particle formation
process. The gel was made from titanium tetraisopropoxide
(Ti(O iPr)4  and H2 O2 .
The crystallization rate of titania gels was found to be
much higher in water than in methanol or n-hexane [146];
also, the crystallite size of anatase prepared in water was
larger than those in organic solvents. Processing parameters
very often control the crystallite size and phase. Nonagglomerated, ultrane anatase particles have been generated by hydrothermally treating solgel-derived hydrous
oxides [147]. The degree of crystallinity and purity of the
synthesized materials may affect their structural evolution
during any heat treatment.
TiO2 thin lms with different surface structures were prepared from alkoxide solutions containing polyethylene glycol
(PEG) via the solgel method [148]. The larger the amount
of PEG added to the precursor solution, the larger the size
and number of pores produced in the resultant lms. When
PEG was added to the gel, the lms decomposed completely
during heat treatment. The adsorbed hydroxyl content of
such porous thin lms is found to increase with increasing
amount of PEG. However, the transmittance of the lms
decreases due to the scattering of light by pores of larger
size and a higher number in the lms. Photocatalytic degradation experiments show that methyl orange is efciently
decolorized in the presence of the TiO2 thin lms by exposing its aqueous solution to ultraviolet light. However, in lms
deposited on soda-lime glass [149], diffusion of sodium and
calcium ions from the glass into the nascent TiO2 lms was
found to be detrimental to the photocatalytic activity of the
resulting lms. Sodium and calcium diffusion into nascent
TiO2 lms was effectively retarded by a 03 m SiO2 interfacial layer formed on the soda-lime glass [149, 150].
TiO2 thin lm photocatalysts coated onto glass plates were
prepared [151] by thermal decomposition of tetraisopropyl
orthotitanate with a dip-coating process using alphaterpineol as a highly viscous solvent. Two types of ligands
polyethylene glycol 600 and (ethoxyethoxy)ethanolwere
added to the dip-coating solution as the stabilizer of titanium alkoxide and thin lms were obtained after calcination
at 450  C for 1 h. The lm thickness obtained with a single
dipping was proportional to the viscosity of the dip-coating
solutions. The thin lms obtained were transparent with a
thickness of 1 m. The crystal form of the lms was anatase
alone. The thin lms were formed with aggregated nanosized TiO2 single crystals (715 nm), and the size of the TiO2
crystals became smaller for the polymer-added systems.
Transparent anatase TiO2 -based multilayered photocatalytic lms synthesized via a solgel process on porous
alumina and glass substrates showed a sponge-like microstructure and a mean crystallite dimension of ca. 8 nm [152].
Doping such lms with iron(III) impeded the photocatalytic
activity.
The effects of calcination on the microstructures of nanosized titanium dioxide powders prepared by vapor hydrolysis was investigated in detail [153155]: Among the factors
examined [153], large surface area and good dispersion
of the powders in the reaction mixture are favorable to
photoactivity. Conversely, prolonged calcination at high

514
temperatures is detrimental to photoactivity. Powders produced at higher temperatures are predominantly anatase
and are generally more photoactive.
The formation, structure, and photophysical properties
of functional mixed lms of 5,10,15,20-tetra-4-(2-decanoic
acid)phenyl porphyrin (TDPP) with TiO2 nanoparticles
formed from the two-dimensional solgel process of tetrabutoxyltitanium (TBT) at the air/water interface was reported
[156]. The composite multilayer lms were assembled by
transferring the mixed monolayer onto quartz plates. The
sensitization of TDPP upon TiO2 nanoparticles was conrmed by the spectral changes of UV-visible absorption and
uorescence of TDPP in the composite lms. Furthermore
the photosensitization greatly affected the photocatalytic
activity of TiO2 particles with respect to the degradation of
methylene blue.
Crystalline titania thin lms were obtained [157] on glass
and various kinds of organic substrates at 4070  C by deposition from aqueous solutions of titanium tetrauoride. Transparent lms consisting of small anatase particles (20 nm)
exhibited excellent adhesion to relatively hydrophilic surfaces. Uniform coatings were successfully prepared on substrates with complex shapes such as cotton and felt ber.
Growth rate and particle size were controlled by both the
deposition conditions and the addition of an organic surfactant. Organic dyes were incorporated into the anatase lms
using organic-dye dissolving solutions and a surfactant.

3.2. Chemical Precipitation


Rutile-phase nano-sized TiO2 powders having a high specic surface area of 180 m2 /g were prepared by homogeneous precipitation at ambient or very low temperatures
(<100  C) [158]. Ultrane SnO2 TiO2 -coupled particles
could be synthesized [159] by homogeneous precipitation;
they were employed for photocatalytic degradation of azo
dye active red X-3B in aerated solution. The results show
that a very rapid and complete decolorization of the azo
dye can be achieved, and the photoactivity of the coupled
particles is higher than that of pure ultrane TiO2 , and the
optimum loading of SnO2 on TiO2 is 18.4%. The enhanced
degradation rate of X-3B using coupled photocatalysts is
attributed to increased charge separation in these systems.

3.3. Gas Condensation and Consolidation


Another method to synthesize nanostructured materials is
by way of gas condensation followed by the in-situ consolidation under high-vacuum conditions [2]. This approach
can produce ultrane-grained materials which may exhibit
size-related effects. The basic aspects of the generation of
nanometer-sized materials via gas condensation are conceptually rather simple [2]: A precursor material, either an element or a compound, is evaporated in a gas maintained at a
low pressure, usually well below one atmosphere. The evaporated atoms or molecules lose energy via collisions with
the gas atoms and undergo a homogeneous condensation
to form atomic or molecular clusters in the highly supersaturated vicinity of the precursor source. In order to maintain small cluster sizes, by minimizing further atom/molecule
accretion and cluster-cluster coalescence, the clusters once

Nanocrystalline TiO2 for Photocatalysis

nucleated must be removed from the region of high supersaturation. Since the aggregates are already entrained in the
condensing gas, this is readily accomplished by setting up
conditions for moving this gas.
Typically, there are three fundamental rates that essentially control the formation of the clusters in the gascondensation process [160]. These are
(i) the rate of supply of atoms to the region of supersaturation where condensation occurs,
(ii) the rate of energy removal from the hot atoms via
the condensing medium, the gas,
(iii) the rate of removal of the cluster upon nucleation
from the supersaturated region.
The clusters that are collected via thermophoresis on the
surface of a cold nger usually form very open, fractal structures. The clusters are held on the collector surface rather
weakly, via Van-der-Waals forces, and are easily removed by
means of a scraper. The material removed is consolidated
in a compaction unit at typical pressures of 12 GPa; the
scraping and consolidation are carried out under ultra-highvacuum conditions in-situ after the removal of the gas from
the chamber, in order to maximize the cleanliness of the
particle surfaces and the interfaces that are formed.

3.4. Film Deposition by Sputtering


and Vacuum-Based Techniques
Crystalline titanium dioxide lms are often deposited by
various techniques employing vacuum conditions, using,
for example, RF magnetron [161166], DC sputtering
[167171], chemical vapor deposition [172, 173], plasma
spraying [174, 175], or related techniques [176, 177]. Rapid
electroplating of photocatalytically highly active TiO2 -Zn
nanocomposite lms on steel was achieved [178] and the
gas-phase oxidation of CH3 CHO was employed as an indicator of the photocatalytic activity. Not surprisingly, the lm
and surface morphologies and the crystallinity are strongly
dependent on the total and the oxygen partial pressure, the
deposition rate, and the phase composition; the resulting
photocatalytic properties can be modied over a wide range
by those parameters. Such lms may show good uniformity of thickness over large areas, high optical transmittance
(80%) in the visible region, and considerable mechanical
durability [169].
Transmission electron microscopy was used to study [179]
the structure, morphology, and orientation of thin TiO2 lms
prepared by reactive magnetron sputtering on glass slides
at different substrate temperatures (100 to 400  C). The
microstructure and photocatalytic reactivity of TiO2 lms
have been shown to be functions of deposition temperature. In the temperature range examined, all lm samples
have a porous nanostructure and the dimension of particles grew with increasing deposition temperature. Films are
amorphous at temperatures of 100  C and only the anatase
phase forms at 200  C and above. Films deposited between
200 to 300  C show a preferred orientation, while lms at
400  C change into complete random orientation. Deposition at 250  C yields high efciency in photocatalytic degradation owing to the high degree of preferred orientation and
nanocrystalline/nanoporous anatase phase.

Nanocrystalline TiO2 for Photocatalysis

Another frequently employed and convenient way for the


preparation of TiO2 thin lms is pulsed laser deposition
[180], although this technique does not produce nanocrystalline structures.

3.5. Surface and Film Modications


The surface of as-deposited lms has frequently been modied in the quest to enhance the catalytic activity. For
example, Fe [181] or Sn [182] ions have been implanted
into transparent and colorless TiO2 thin lms fabricated on
microscope glass slides by DC magnetron reactive sputtering using Ar and O2 as working gases. The efciency of
this procedure could not as yet be proven unambiguously.
Apart from grain size, the presence of reactive species on
the surface [183] may inuence the photocatalytic activity.
Implanted metal ions (V or Cr) were observed [184] located
at the lattice positions of Ti4+ in TiO2 and were stabilized
after calcination of the samples in an O2 ambient at around
775 K. Spectroscopic studies showed that the presence of
these substitutional metal ion species are, in fact, responsible
for a large shift in the absorption spectra of these catalysts
toward visible light regions. Porous anatase TiO2 lms were
densied by Zn+ ion implantation up to the ion penetration depth [185]. After the subsequent annealing at 800  C,
the phase transformation from anatase to rutile accompanied with grain growth up to the lm thickness was observed.
In addition, the phase transformation was not induced by
the annealing up to 800  C with or without preceding Ar+
ion implantation. Thus, the implanted impurity Zn assisted
the phase transformation. Annealing in O2 tends to reduce
the rate of phase transformation and create ZnTiO3 . Optical
absorption above the photon energy of 2.9 eV was increased
remarkably by the Zn+ or Zn+ and O+ ion implantation and
subsequent annealing and is due to the phase transformation.
The presence of active species such as Ti3+ and hydoxyl
on the surface of ultrane TiO2 particles, prepared by a colloidal chemical approach and subjected to different heating
treatments, was inferred [186] from optical absorption and
photoelectron spectra; these species may enhance the photocatalytic activity of particles. Treating TiO2 powder by a
hydrogen plasma resulted in a reduction of the oxide particles and electrons trapped at the oxygen-defect sites were
found [187]. Also laser ablation of TiO2 photocatalysts aiming at the enhancement of the activity was reported [188].
Noble metal particles of Au, Pt, and Ir were deposited
on nanostructured TiO2 lms using an electrophoretic
approach [189]. The improved photoelectrochemical performance of the semiconductor-metal composite lm was
attributed to the shift in the quasi-Fermi level of the composite to more negative potentials. Continuous irradiation
of the composite lms over a long period causes the photocurrent to decrease as the semiconductor-metal interface
undergoes chemical changes.
Doping of nanostructured TiO2 both as particles and in
thin lms with a variety of metals has been reported, some
common examples being Pt [190, 191], Pb [192], Au [193],
or others [194]. A shift of the UV-vis absorption towards
longer wavelengths was observed upon Pb doping, which
indicates a decrease in the bandgap of TiO2 . In TiO2 lms
embedding Au nanoparticles [193], the specic resistance of

515
the lms experienced a rising phase, followed by a dramatic
drop with an increasing number of Au particles. Ultrane
Pt particles [191] were embedded into rutile TiO2 particles
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
on the TiO2 crystallites with a well-dened crystallographic
relationship.
Doping a nano-structured TiO2 electrode sensitized with
tetrasulfonated gallium phthalocyanine with tetrasulfonated
zinc porphyrin (ZnTsPP) greatly enhances the photoelectric conversion at long wavelengths, with 20- and 60fold improvement of the quantum efciency at 680 and
700 nm [195].
Semiconductor/metal composite nanoparticles have been
synthesized [196] by chemically reducing HAuCl4 on the surface of preformed TiO2 nanoparticles. These gold-capped
TiO2 particles (diameter 1040 nm) were stable in acidic
(pH 24) aqueous solutions. The role of the gold layer
in promoting the photocatalytic charge transfer has been
probed using thiocyanate oxidation at the semiconductor
interface. More than 40% enhancement in the oxidation efciency was found with TiO2 /Au nanoparticles capped with
low concentration of the noble metal.
Magnetic photocatalysts were synthesized by coating titanium dioxide particles onto colloidal magnetite and nanomagnetite particles [197]. The photoactivity of the prepared
coated particles was lower than that of single-phase TiO2
and was found to decrease with an increase in the heat
treatment. These observations were explained in terms of
an unfavorable heterojunction between the titanium dioxide
and the iron oxide core, leading to an increase in electronhole recombination.
TiO2 -based powders, doped with a small amount of
SiO2 , were prepared by a solgel method and were subsequently heated to precipitate ne anatase crystals [198].
The obtained powders have large specic surface areas
(200 m2 /g) and upon treating them chemically with aqueous NaOH, the photocatalytic property of the powders was
extremely improved and the powders showed higher activity
than the undoped TiO2 powders.
In composite TiO2 -SiO2 thin lms prepared by a solgel
process, the refractive index and the photocatalytic activity
decrease with increasing SiO2 content in these lms [199].
Alkoxide solgel processing has been investigated for the
synthesis of stable SiO2 -TiO2 high-permeability catalytic
membranes to be used in alkene isomerization [200].
Nanocrystalline TiO2 was prepared on mesoporous silica both by solgel processes [201] and by an impregnation method with titanium complexes featuring different
ligands [202].
Binary mixed oxide of Fe/Ti (1:1 composition) with homogeneous distribution of iron into the TiO2 has been prepared
by solgel impregnation using metal alkoxide precursors and
ring at different temperatures (500, 700, and 900  C) [203].
The mixed oxide exhibits excellent absorption in the visible
spectral region (570600 nm). The photocatalytic activity of
the Fe/Ti oxide reduces to a large extent at a high sintering
temperature of the sample due to the presence of a increasing amount of the inactive (Fe2 /TiO5  pseudobrookite phase.
Nanostructured TiO2 /SnO2 binary oxides were prepared by

516
combustion of stearic acid precursors [204], with metal precursors being dispersed in the stearic acid at the molecular
level. It was found that preparative methods affected the
crystalline structure of the powders and the anatase phase
of TiO2 was stabilized by the addition of SnO2 .
Spray painted (spray deposited) titanium dioxide
coatings were sensitized [205] with chemically deposited
cadmium selenide thin lms; the structural, optical, and photoelectrochemical characterization of these composite lms
indicate the importance of thermal treatments in improving
the photocurrent quantum yields. Up to 400  C, the effect
of air annealing is to shift the onset of absorption to longer
wavelengths and improve the photocurrent substantially.
Organic compounds may play a crucial role in chemical
processes for ceramic coatings [206]. Organic compounds
remained in a xed position in the coating, which was prepared from the chemically modied titanium tetraisopropoxide solution and heated at temperatures as high as 673 K. It
was not until the organic compounds decomposed to carbon
dioxide and the gas phase was left from the coating that the
nanostructure, consisting of nano-sized pores and anatase
crystallites with preferred orientation, developed at 723 K.
Cobalt(II) 4,4 ,4 ,4 -tetrasulfophthalocyanine, covalently
linked to the surface of titanium dioxide particles, TiO2 CoTSP, was shown [207] to be an effective photocatalyst for
the oxidation of sulfur (IV) to sulfur (VI) in aqueous suspensions. Upon bandgap illumination of the semiconductor,
conduction-band electrons and valence-band holes are separated; the electrons are channeled to the bound CoTSP
complex resulting in the reduction of dioxygen, while the
holes react with adsorbed S(IV) to produce S(VI) in the
form of sulfate.
The photoactivity of the Pt/TiO2 system in the visible
region was improved [208] by the addition of the sensitizer
([Ru(dcbpy)2 (dpq)]2+  [where dcbpy = 4,4 -dicarboxy 2,2 bipyridine and dpq = 2,3-bis-(2 -pyridyl)-quinoxaline] leading to an efcient water reduction.
Photocatalytic properties for hydrogen production were
investigated [209] on layered titanium compounds intercalating CdS in the interlayer, which were prepared by direct
cation exchange reactions and sulfurization processes. The
photocatalytic activity of the compounds intercalating CdS
was superior to those of simple CdS and the physical mixture of CdS and metal oxides. The improvement might be
attributed to the formation of microheterojunctions between
the CdS nanoparticles and the layers of oxides.

3.6. Novel Deposition Methods


and Structures
In recent years, there has been increased interest in studying
and manufacturing nanoscaled TiO2 materials as nanoparticles [210], nanowires [211], nanorods [212], whiskers [213],
and nanotubes [214217].
There are many synthetic routes for the creation of
nanocrystals of oxides and the controlled hydrolysis of metal
alkoxides is the most generalized solution-phase synthetic
strategy [218]. Increased photocatalytic activity was reported
recently for TiO2 prepared by ultrasonic irradiation and glycothermal methods. This novel method [219] for preparing highly photoactive nanometer-sized TiO2 photocatalysts

Nanocrystalline TiO2 for Photocatalysis

with anatase and brookite phases has been developed by


hydrolysis of titanium tetraisopropoxide in pure water or
a 1 + 1 EtOHH2 O solution under ultrasonic irradiation;
the photocatalytic activity of TiO2 particles prepared by this
method exceeded that of Degussa P25 and was the rst
report that showed high photocatalytic activity of a photocatalyst containing an 80% anatase and 20% brookite phase.
A novel and convenient nonhydrolytic approach to the
preparation of uniform, quantum conned TiO2 nanocrystals, using an intramolecular adduct stabilized alkoxide
precursor, was described recently [220]. In contrast to established aqueous solgel-techniques, the processing in hydrocarbon solvents at high temperatures allows access to very
small free-standing crystallites, and opens up new possibilities for control over size distribution, surface chemistry, and
particle agglomeration.
It has been reported that the columnar morphologies
in sputtered TiO2 lms enhances the photocatalytic [221]
and photovoltaic [222] efciency. Following these studies,
enhanced surface-reaction efciency has been demonstrated
in the photocatalysis of sculptured thin lms of TiO2 [223].
In obliquely deposited lms with variously shaped columns
such as zigzag, cylinder, and helix, the columnar thickness
and spacing play an important role in the enhancement of
the effective surface area, while the columnar shape is less
important. The optimum morphology for a surface reaction
has been obtained at the deposition angle of 70 , where the
photocatalytic activity is 2.5 times larger than that at 0 . The
morphology of these obliquely deposited thin lms appears
well suited for application as solar cells, electro- and photochromic devices, and photocatalysts.
The template method for synthesizing nanostructures
involves the synthesis of the desired material within the
pores of a nanoporous membrane or other solid. This
approach has been used in several experiments [224229] for
the preparation of TiO2 nanotubes and nanorods; typically,
porous aluminum oxide (PAO) nano-templates were used.
Compact, continuous, and uniform anatase nanotubules
with diameters in the range 5070 nm were produced
inside PAO nano-templates by pressure impregnating the
PAO pores with titanium isopropoxide and then oxidatively
decomposing the reagent at 500  C [230]. Cleaning the surface of the template and repeating the process several times
produced titania nanotubules with a wall thickness of 3 nm
per impregnation. The tube exteriors appeared to be faithful
replicas of the pores in which they were formed.
Nano-TiO2 whiskers were prepared by various techniques
[213, 231]; using, for example, controlled hydrolysis of
titanium butoxide [231] it was found that the nano-TiO2
whiskers obtained were anatase and grew selectively in the
[001] direction with a diameter of about 4 nm and a length
of about 40 nm. Acetic acid played an important role in the
oriented growth of nano-TiO2 whiskers.
Highly ordered TiO2 nanowire (TN) arrays were prepared
[211] in anodic alumina membranes by a solgel method.
The TNs are single crystalline anatase phase with uniform
diameters around 60 nm. At room temperature, photoluminescence measurements of the TN arrays show a visible
broadband with three peaks, which are located at about 425,
465, and 525 nm that are attributed to self-trapped excitons,
F , and F + centers, respectively.

517

Nanocrystalline TiO2 for Photocatalysis

4. STRUCTURAL PROPERTIES
OF NANOCRYSTALLINE TiO2 FILMS
4.1. The Lattice Structure of Rutile
and Anatase
Three different crystal structures of TiO2 exist [18]: rutile,
anatase, and brookite; only the former two of them are commonly used in photocatalysis, with anatase typically exhibiting the higher photocatalytic activity [232]. The structure
of rutile and anatase can be described in terms of chains
of TiO6 octahedra. The two structures differ by the distortion of each octahedron and the actual pattern of the
chains. Figure 7 depicts the unit cell structures of rutile and
anatase crystals [233235]. Each Ti4+ ion is surrounded by an
octahedron of six O2 ions. The octahedron in rutile shows
a slightly orthorhombic distortion, whereas the respective
octahedra in anatase are signicantly 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 compared to 0.357 and 0.296 nm in rutile) while the Ti-O distances are shorter than in rutile (0.1934 and 0.1980 nm in
anatase versus 0.1949 and 0.1980 nm in rutile). In the rutile
structure, each octahedron is in contact with 10 neighboring
ones (two sharing edge oxygen pairs and eight sharing corner oxygen atoms), whereas in the anatase crystal each octahedron is in contact with eight neighbors (four sharing an
edge and four sharing a corner). These differences in lattice
structure result in different mass densities ( = 4250 g/cm3
for rutile and  = 3894 g/cm3 for anatase) and electronic
band structures for the two forms of TiO2 .
Synthetic titanium oxide crystallizes in two polymorphs:
anatase and rutile. Anatase is metastable and transforms
exothermally and irreversibly to rutile. Some properties of
TiO2 may strongly depend on its polymorphic phase. The
anatase-rutile transformation is strongly inuenced by the
synthesis method, atmosphere, grain growth, and impurities. Some additives, such as ZrO2 and Al2 O3 , retard the
anatase-rutile transformation, whereas others, such as CoO
and ZnO, accelerate such a process [236]. The anatase-torutile 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

90
81.21

On the other hand, La retarded both transformation and


densication.
Transmission electron microscopy (TEM) is typically used
to investigate the crystal size distribution, grain-boundary
disorder, and defect structure in nanocrystalline TiO2 materials prepared by the various techniques outlined in the previous section. In a recent study of lms with an average grain
size of 15 nm prepared by reactive sputtering [238], evidence
of both ordered and disordered grain-boundary regions was
found and planar defects were observed in grain interiors
identied as (011) deformation twins. Also, crystallographic
shear defects can occur as a result of aggregation of oxygen
vacancies in understoichiometric titanium oxide. Figure 8
shows results of TEM investigations [239] of nanocrystalline
TiO2 lms prepared from colloidal suspensions; the TiO2
crystallites were nominally pure anatase phase with a size of
16 nm.

4.2. Structure of Crystalline TiO2 Surfaces


The geometric structure of crystalline surfaces of TiO2 has
been studied predominantly on (macroscopically) large single crystals; in particular, the (110) surface of rutile, being
thermodynamically the most stable one [240], has been
investigated extensively by a broad variety of surface science
tools [241243]. Among the different questions addressed
therein, a prominent one was related to the possible types
of oxygen defects at the surface; in fact, three distinct oxygen vacancy sites were tentatively identied: lattice, singlebridging, and double-bridging vacancies [244]. With the
widespread use of scanning-tunneling microscopy (STM),
atomically resolved investigations of TiO2 surfaces became
feasible and a more detailed picture of the rutile (110) surface structure emerged [245248]. Consistently, these studies
corroborate the longstanding notion about the prominent
importance of surface defects as the active sites for various types of chemical reactions [249251], for example, for

Oxygen

78.12
92.43

Figure 7. Crystal structures of rutile (left) and anatase (right) TiO2 .

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


nanocrystalline TiO2 anatase lm. The nominal crystallite size is 16 nm.

518
the dissociation of water [252, 253]. The number of studies on the structural properties of anatase single crystals
is considerably smaller, which appears to be largely due to
the difculties encountered in preparing such surfaces in a
defect-free state. Nevertheless, a rather distinct picture of
the anatase TiO2 surface structure [254257] and its properties in terms of adsorption/desorption reactivity [258260]
has been achieved.
The structure and composition of a nanocrystalline surface may have a particular importance in terms of chemical and physical properties because of their small size. For
instance, nanocrystal growth and manipulation relies heavily on surface chemistry [261]. The thermodynamic phase
diagrams of nanocrystals are strongly modied from those
of the bulk materials by the surface energies [262]. Moreover, the electronic structure of semiconductor nanocrystals
is inuenced by the surface states that lie within the bandgap
but are thought to be affected by the surface reconstruction process [263]. Thus, a picture of the physical properties
of nanocrystals is complete only when the structure of the
surface is determined.
To understand and improve the applications of titaniumoxide nanoparticles, it is extremely important to perform a
detailed investigation of the surface and the interior structural properties of nanocrystalline materials, such as rutile
and anatase with diameter of a few nanometers. Detailed
experiments using X-ray absorption near-edge structure
spectroscopy (XANES) demonstrated [264] that the presence of both defects and surface states strongly inuence
the X-ray absorption spectra, even though the rst nearestneighbor geometrical arrangement around the central Ti
atom in both rutile and anatase is quite similar: the differences in the XANES spectra arise from the outer-lying
atomic shells, indicating that medium to long-range
effects play an important role to the near-edge features.
In another study of this kind [265], a shorter Ti-O distance for surface TiO2 , resulting from Ti-OH bonding was
observed together with a minor disorder of the lattice
in smaller nanoparticles. Nevertheless, the Ti sites largely
remain octrahedral even in particles with diameters of 3 nm.
Because the interfacial electron/hole transfer occurs via surface Ti or O atoms, the observed structural changes around
the surface Ti atoms in small TiO2 particles could be responsible [265] for the unique photocatalytic properties.
A qualitative analysis of opaque nanostructured glasssupported TiO2 lms was carried out [266] using scanning
force microscopy (SFM), and surface parameters such as
average grain diameter, roughness exponent, and fractal
dimension [267] were determined. The TiO2 surfaces exhibit
distinct roughness due to the large aggregates formed by the
interconnected TiO2 particles. Fractal dimension was found
to range between 2.10 and 2.45, depending on the scanned
range and the preparation method. The surface morphology of nanocrystalline materials prepared by compacting
nanometer-sized clusters was investigated by SFM [268];
these materials had a grain size of 515 nm and contained
about 1019 interfaces per cubic centimeter. Upon heat treatment, grains were found to fuse together forming bamboolike structures and then lined up as tubular structures.
The inuence of the iron concentration in mixedoxide (TiO2 and Fe2 O3  thin lms prepared by reactive

Nanocrystalline TiO2 for Photocatalysis

radio-frequency sputtering on the structural properties of


the layers has been studied [269]. This characterization
allowed the correlation of the inhibition of the grain growth
of titania to the presence of iron oxide and its segregation
at grain boundaries. This behavior could be ascribed to a
supercial-tension phenomenon. As a possible application
of these thin lms, it was observed that they were able to
sense CO down to the level requested for environmental
monitoring.
A study [270] of the structure and morphology of a titanium dioxide photocatalyst (Degussa P25) reveals multiphasic material consisting of an amorphous state, together with
the crystalline phases anatase and rutile in the approximate
proportions 80/20. Transmission electron microscopy provides evidence that some individual particles are a mixture
of the amorphous state with either the anatase phase or
the rutile phase, and that some particles, which are mostly
anatase, are covered by a thin overlayer of rutile that manifests its presence by the appearance of Moir fringes. The
photocatalytic activity of this form of titanium dioxide is
reported as being greater than the activities of either of the
pure crystalline phases, and an interpretation of this observation has been given in terms of the enhancement in the
magnitude of the space-charge potential, which is created
by contact between the different phases present and by the
presence of localized electronic states from the amorphous
phase.

5. ELECTRICAL PROPERTIES
OF NANOCRYSTALLINE TiO2 FILMS
Most studies into the carrier transport in nanocrystalline
TiO2 were carried out with the lms in contact with electrolytes, mostly due to their use in highly efcient electrochemical solar cells, often called Grtzel cells [38, 71]. In
this application, the pore surface is covered with an ultrathin
organic dye layer and contacted with an electrolytic solution
that penetrates the pore structure. The experimental work
indicates [271274] that in this conguration the electrolyte
screens any electric eld within the porous structure and
establishes diffusion conditions for the carrier propagation.
On the other hand, investigations in which the nanopores
are in contact with an insulating medium (a gas or vacuum)
may allow one to obtain quantitative insight into the electronic properties of the material and the basic feature of carrier transport. In a series of measurements [275277] on the
electron transport in nanoporous TiO2 lms with gas-lled,
insulating pores employing Pt/TiO2 , Schottky barrier structures indicate a barrier height of 1.7 eV, compatible with an
electron afnity of 3.9 eV for the TiO2 lms. Below 300 K,
tunneling transport through the barrier occurs, resulting in
barrier lowering effects. Carrier drift mobilities, recombination lifetimes, and their dependence on injection level
in TiO2 are reported. It is found that the mobility-lifetime
product is independent of injection level, while drift mobility
and recombination lifetime change strongly with injection.
A trap-lling model appears as the most plausible model
compatible with the experimental ndings [277]. Comparison with recent experiments on nanoporous lms in contact
with electrolytes indicate that the transport and recombination mechanism is qualitatively similar for the two cases.

Nanocrystalline TiO2 for Photocatalysis

Various observations indicate that electron transport in


the nanocrystalline TiO2 dominates the transient response
of the system. Transient photocurrent measurements reveal
a very slow (millisecond), multiphasic response to both
continuous wave [278, 279] and pulsed [280282] illumination. The characteristic rise or decay time of the response
is dependent upon the intensity of the light source [278,
281283]. Comparison with the transient response without
electrolyte indicates that it is the TiO2 , and not the electrolyte, which is responsible for the very long tail [284, 275].
A slow and multiphasic time dependence has also been
observed in the rereduction of oxidized dye molecules in a
redox inactive environment [285]; the same work indicates
that the rate of dye reduction is controlled by the concentration of electrons introduced into the TiO2 by externally
applied bias. The wide range of time scales is consistent with
the assumption that electron transport within the TiO2 is the
rate limiting step.
The slow processes are attributed to the trapping of electrons by a high density of localized states in the TiO2 . Since
the TiO2 grains are normally crystalline [286], the localized
states are believed to be concentrated at the grain boundaries and on the very large surface. There is evidence for
intraband-gap states in bias-dependent optical absorbance
spectra [287] and surface photovoltage spectra. It would,
therefore, be extremely useful to correlate the density and
nature of those states with the electronic transport properties of the material.
Investigating electron migration in nanostructured
anatase TiO2 lms with intensity-modulated photocurrent
spectroscopy [288], it was found that, upon illumination, a
fraction of the electrons accumulated in the nanostructured
lm is stored in deep surface states, whereas another
fraction resides in the conduction band and is free to move.
These data indicate that the average concentration of the
excess conduction band electrons equals about one electron
per nanoparticle, irrespective of the type of electrode, the
lm thickness, or the irradiation intensity.
The photocurrent in thin lm TiO2 electrodes prepared
by solgel methods was studied in [289] as a function of
lm thickness. Films with thickness smaller than the space
charge layer were found to show a larger photocurrent than
lms with thickness larger than the space charge layer. It
was concluded that the increase in photocurrent is due to
the effective electron-hole separation throughout the whole
lm thickness and the reduction of bulk recombination. The
use of TiO2 thin-lm electrodes for photocatalytic devices
might therefore be useful to gain high device performance.
In a series of papers, Dittrich et al. carried out extensive
investigations of the electrical conductivity in nanoporous
TiO2 lms [290296]. Studying the temperature- and oxygen partial pressure-dependent conductivity of rutile and
anatase, they noted [292] that is thermally activated with
EA = 085 eV, independent of the absolute value of and
depends on p(O2  by a power law for p(O2  < 110 mbar.
The electrical properties of reduced nanoporous TiO2 are
determined by surface chemical reactions which lead to the
formation of shallow donor and deep trap states. Furthermore, this group examined in detail the photovoltage in
nanocrystalline TiO2 [293, 295] and the injection currents in
these porous specimens [294, 296].

519
Such a power dependence of on the oxygen partial pressure was noted also in recent work [297] investigating electrical and defect thermodynamic properties of nanocrystalline
titanium oxide. At high O2 pressures, p(O2  > 1 mbar,
the conductivity is constant, whereas at values p(O2  <
1014 mbar a steep increase of with decreasing pressure
was found, following a power dependence p(O2 n with
n = 1/2 [297]. The plateau of conductivity at high oxygen pressures can be interpreted as being a domain of ionic
conductivity, an unexpected behavior for titanium dioxide.
In a coarse-grained material, dominant hole conductivity is
observed in this partial pressure range. This difference may
be due to the high density of grain boundaries in nanocrystalline ceramics, which can be preferred paths for diffusion at reduced temperatures. Furthermore, an increase in
ionic conductivity is also expected due to enhanced defect
formation in the space charge regions adjacent to grain
boundaries [298]. At low oxygen partial pressure, nanocrystalline TiO2 exhibits enhanced electronic conductivity as
compared to coarse-grained TiO2 . The power exponent n =
1/2 can be explained under the assumption that doubly
charged titanium interstitials are formed. The intrinsic disorder of titanium dioxide is reputedly of the cationic Frenkeltype [299302], although alternative defect models based
on Schottky disorder are also described in the literature
[303, 304]. In the domain of ionic conductivity, the activation energy of conduction is 10 01 eV [297], a value
typical of migration enthalpies for ionic defects. By contrast,
the activation energy in the reduction-controlled regime was
found to be 39 02 eV.
In titanium oxide thin lms prepared by a d.c. sputtering technique onto glass substrates with average grain sizes
of 100200 nm, the surface structure and phase morphology
of the lms was found [305] to depend on the deposition
conditions. The current-voltage characteristics of these lms
are ohmic for values of applied voltage lower than 0.5 V.
For higher values, the mechanism of electrical conduction
is determined by space-charge-limited currents [306]; then,
a power-law dependence was observed with n 2.32.9. In
much thicker Ti oxide lms (1.98 m) deposited by sputtering [307], both the surface roughness and the internal
surface area increased with lm thickness; this resulted in an
enhancement of the incident photon-to-current efciency.
Electrical and optical spectroscopic studies of TiO2
anatase thin lms deposited by sputtering showed [308, 309]
that the metastable phase anatase differs in electronic properties from the well-known, stable phase rutile. (From the
broadening of the X-ray diffraction peaks, the average grain
size of the lms is estimated to be in the range of 3040 nm
[308].) Resistivity and Hall-effect measurements revealed an
insulator-metal transition in a donor band in anatase thin
lms with high donor concentrations. Such a transition is
not observed in rutile thin lms with similar donor concentrations. This indicates a larger effective Bohr radius
of donor electrons in anatase than in rutile, which in turn
suggests a smaller electron effective mass in anatase. The
smaller effective mass in anatase is consistent with the
high-mobility, bandlike conduction observed in anatase crystals. It is also responsible for the very shallow donor energies in anatase. Luminescence of self-trapped excitons was

520
observed in anatase thin lms, which implies a strong lattice relaxation and a small exciton bandwidth in anatase.
Optical absorption and photoconductivity spectra show that
anatase thin lms have a wider optical absorption gap than
rutile thin lms. The extrapolated optical absorption gaps of
anatase and rutile lms were found to be 3.2 and 3.0 eV,
respectively, at room temperature.
The observation of space charge limited currents (SCLC)
in nanoscaled pure and chromium-doped titania was
reported [310] and both the free-charge carrier density and
the trapped-charge carrier density were given.
Photoconductivity was also studied in compound systems;
for example, in a TiO2 -C60 bilayer system the conductivity
increases signicantly in the fullerene upon irradiation at
wavelengths <300 nm [311].
Although being an efcient photocatalyst for the detoxication of organically charged waste water, titanium dioxide suffers from the drawback of poor absorption properties
because of a bandgap of 3.2 eV. Thus, wavelengths shorter
than 400 nm are needed for light-induced generation of
electron-hole pairs. That is the reason why doping with transition metal ions is interesting for inducing a reduction of the
bandgap. However, this doping changes other physical properties such as lifetime of electron-hole pairs and adsorption
characteristics [312].

6. PHOTOCATALYTIC PROPERTIES
OF NANOCRYSTALLINE TiO2
Most experimental investigations reported a higher photocatalytic efciency in the anatase TiO2 phase; as a possible reason it was suggested that the recombination of the
electron-hole pairs produced by UV irradiation occurs more
rapidly on the surface of the rutile phase, and the amount of
reactants and hydroxides attached to this surface is smaller
than on the surface of the anatase phase.
The study of the photocatalytic activity of nanocrystalline
TiO2 materials is a longstanding research effort [313315];
in most lab-scale experiments it was evaluated by means
of the degradation observed for typical substances (e.g.,
aqueous methyl orange, methylene blue, etc.) upon exposure of the specimen to UV irradiation. In such a way,
the possible inuence of light intensity, structural properties, surface morphology, phase and chemical composition,
resulting from various deposition or preparation methods,
could conveniently be explored [316320]. Furthermore,
any correlation with the optical or electrical properties of
the nanocrystalline lms could thereby be investigated. In
addition, alternative approaches have come under scrutiny.
A new simple method for characterizing photocatalytic
activity by measuring photo-generated transient charge separation at the surface of semiconductor photocatalysts was
proposed [321]. In this technique, the charge separation generated by a pulse dye laser is obtained as a function of the
incident laser energy; thereby, the photocatalytic activity and
the type of surface reaction (reduction or oxidation) in titanium dioxide lms were rapidly determined. In the following
sections some examples of such studies will be given.

Nanocrystalline TiO2 for Photocatalysis

6.1. Dependence of Photocatalytic Activity


on Film Structure and Phase
The photoactivities of ultrane TiO2 nanoparticles in
anatase, rutile, or mixed phases were tested in the photocatalytic degradation of phenol [322]. For TiO2 nanoparticles,
mainly in the anatase phase and mixed-phases, the photocatalytic activities increased signicantly with the content of
the amorphous part decreasing. The completely crystallized
rutile nanoparticles exhibited size effects in this photocatalytic reaction and the photocatalytic activity of rutile-type
TiO2 nanoparticles with a size of 7.2 nm was much higher
than that with 18.5 nm or 40.8 nm and was comparable to
that of anatase nanoparticles.
A modied solgel process was used [323] to prepare
nano-structured TiO2 catalysts of controlled particle size
(i.e., 6, 11, 16, and 20 nm). The effect of TiO2 particle size on gas-phase photocatalytic oxidation of toluene
was examined under dry and humid conditions. Main reaction products were CO2 and water, although small amounts
of benzaldehyde were also detected. The smaller particle
size (i.e., 6 nm) led to higher conversion and complete
mineralization of toluene into CO2 and H2 O. Electronic
and structural effects (i.e., size and ensemble effects) were
responsible for the excellent performance of a 6 nm TiO2
catalyst for toluene photodegradation. The dependency of
the photoactivity on the crystallite size of anatase titania for
the decomposition of trichloroethylene (TCE) was investigated [324]. It was found that the photoactivity of all titania samples was linearly increased as the crystallite size of
the anatase phase increased, regardless of the preparation
method, as long as there was no signicant rutile phase.
The enhancement of the photocatalytic activity of TiO2
was investigated as a function of added amount of other
oxides to promote desired oxidation or reduction reactions [325]. Mixed oxides of Nb2 O5 or Li2 O with TiO2 were
prepared by the solgel process. The target material of
dichloroacetic acid (DCA) was chosen for oxidation reactions and K2 Cr2 O7 for reduction. While the Nb-oxide had
a deleterious effect on the decomposition rate of DCA, the
excess electrons due to the doping of Nb2 O5 into TiO2 promoted the reduction process for Cr6+ . Li2 O (1 wt%) with
TiO2 was found to be the most efcient photocatalyst for
DCA oxidation, resulting in photocatalytic activity of 50%.
A highly sensitive biochemical oxygen demand (BOD)
sensor using a commercial TiO2 semiconductor and photocatalytic pretreatment was developed to evaluate low BOD
levels in river waters [326]. The photocatalytic oxidation was
investigated as a function of irradiation times, TiO2 concentrations, and pH. The optimal irradiation time was 4 min.
The sensor response was increased with increasing pH and
the responses obtained by photocatalysis to river samples
were higher than those obtained without photocatalysis.

6.2. Inuence of Surface Morphology


and Defects
Photocatalytic reduction of CO2 to organic compounds was
carried out [327] in a semiconductor suspension system
under simulated solar power using a TiO2 catalyst. Experimental results show that the photocatalytic activity can be

Nanocrystalline TiO2 for Photocatalysis

improved by depositing Pd or Ru on the TiO2 surface. Films


of TiO2 dispersed or coated with platinum were deposited
on glass and Pt-buffered polyamide substrates, respectively,
by magnetron sputtering [328]. The photocatalytic activity of
the lms was evaluated through the decomposition of acetic
acid under UV irradiation. The Pt-dispersed TiO2 lm with
approximately 1.5 wt% platinum shows a maximum activity
due to the formation of anatase phase with a ne grain size.
Platinum particles 2 nm in thickness coated on anatase
lm greatly improves activity. The activity shows a steplike
dependence of lm thickness, where the critical thickness
varies between 150 and 200 nm depending on the deposition
temperatures. The correlation between defects and activity
was veried by measuring either the temperature dependence of electric resistance or the shift of binding energy
from X-ray photoelectron spectroscopy (XPS).
In crystalline TiO2 lms deposited by reactive RF magnetron sputtering on glass substrates without additional
external heating, the photocatalytic activity was evaluated by
the measurement of the decomposition of methylene blue
under UV irradiation [162]. The results showed that crystalline anatase, anatase/rutile, or rutile lms can be successfully deposited on unheated substrates. Anatase TiO2
lms with a more open surface, a higher surface roughness,
and a larger surface area, formed at higher total pressures,
exhibit the best photocatalytic activity. The photocatalytic
activity of polycrystalline anatase TiO2 lms was found [329]
to be affected by the crystalline orientation that depends
on the deposition temperature; it was greater on the (112)oriented than on the (001)-oriented lm. The former lm
exhibited a columnar structure resulting in a larger surface area for photocatalytic reaction than the lms with
the (001)-preferred orientation. Furthermore, the introduction of structural defects associated with oxygen vacancies
was found [169] to create some energy levels around the
mid-gap, indicating that they could work as recombination
centers of photo-induced holes and electrons, causing the
decrease in photocatalytic activity. Therefore, the decrease
in the structural defects associated with oxygen vacancies
appears to be important for improving the photocatalytic
activity of the lms.
A marked difference of the photocatalytic activity
between the TiO2 lms coated on quartz and glass substrates
was conrmed [330], which would be interpreted in terms of
the difference in the photocarriers diffusion length induced
by impurity Na+ ions. These results lead to a conclusion
that the crystallinity and defects of TiO2 as well as the lm
thickness and surface area have a great inuence on the
photocatalytic activity. An enhanced photocatalytic activity
of TiO2 could be achieved also by deposition of the lms on
sulfonated glass substrates [331] or by using special support
materials [332, 333].
Photo-oxidative self-cleaning and antifogging effects of
transparent titanium dioxide lms has attracted considerable
attention for the past decade [334, 335]. In order to understand the photo-induced hydrophilic conversion on titanium
dioxide coatings in details, it is inevitably necessary to understand the relationship between the photo reaction and the
surface crystal structure; this can be done, for example, by
an evaluation of the photo-induced hydrophilic conversion
on the different crystal faces of rutile single crystals and also

521
polycrystalline anatase titanium dioxide [336]. Self-cleaning
and antifogging effects of TiO2 lms prepared by magnetron
sputtering were investigated [161, 163] in terms of the photocatalytic behavior by measuring the decomposition of methylene blue and the reduction of the contact angle between
water and TiO2 under ultraviolet irradiation. The phase conversion from the rutile to the anatase TiO2 lm leads to an
enhancement of the activity; the anatase lms with the best
photocatalytic behavior are prepared at higher total pressures (>1.50 Pa) and characterized by a high decomposition
efciency, a contact angle about 10 after irradiation, and a
good stability in darkness.
Titanium dioxide thin lms prepared with various surface morphologies by metalorganic chemical vapor deposition were found to exhibit reversible wettability control by
light irradiation [337]. These TiO2 surfaces became highly
hydrophilic by UV irradiation, and returned to the initial
relatively hydrophobic state by visible-light (VIS) irradiation. The hydrophobic-hydrophilic conversion induced by
UV light was ascribed to the increase in dissociated water
adsorption on the lm surface. By contrast, the conversion from hydrophilic to hydrophobic by VIS irradiation was
caused by the elimination of water adsorbed on the surface due to the heat generated. Changes of the water contact angle between hydrophilic states and hydrophobic ones
strongly depended on the roughness of the lm surface.
The self-cleaning property of thin transparent TiO2 coatings on glass under solar UV light was studied [338] for two
compounds: palmitic (hexadecanoic) acid and uoranthene,
which are both present in the atmospheric solid particles.
The removal rates of layers of these compounds sprayed on
the self-cleaning glass were found to be sufcient for the
expected application. The identied intermediates (about
40 for each compound) show the gradual splitting of the
palmitic acid chain and the oxidative openings of the aromatic rings of uoranthene. In the case of palmitic acid,
the products give some indications about the photocatalytic
mechanism. About 20% of the organic carbon contained in
the initial compounds was transformed into volatile carbonyl
products.
An extreme photo-induced hydrophilicity was achieved
[339, 340] when TiO2 lms were covered by SiO2 overlayers (with 1020 nm in thickness). These multilayer lms
exhibited much more extreme hydrophilicity than a TiO2
lm without overlayer. The surface analyses revealed that
the enhanced photo-induced hydrophilic surface of the multilayer lms exhibited an improved photocatalytic activity
towards decomposition of organic substances on their surfaces. An extreme light-induced superhydrophilicity was also
reported [341] for mesoporous TiO2 thin lms (crystallite
size 15 nm, surface area 50 m2 /g, pore size 3.6 nm).
For such lms, the water contact angle was found to be
reduced essentially to zero upon UV-irradation for a duration of about 60 min. In addition, the photocatalytic activity
of those lms could be enhanced by treating the substrate
surfaces with an H2 SO4 solution [342].
In order to investigate the cathodic photoprotection of
the steel from corrosion, stainless steel was coated with
TiO2 thin lms, applied by a spray pyrolysis [343]. It
was concluded that these coatings exhibit both a cathodic

522

Nanocrystalline TiO2 for Photocatalysis

photo-protection effect against corrosion and the frequently


reported photocatalytic self-cleaning effect.

h
CB

VB

6.3. Inuence of Electronic Properties


Detailed spectroscopic investigations of the processes occurring upon bandgap irradiation in colloidal aqueous TiO2
suspensions in the absence of any hole scavengers showed
[344] that while electrons are trapped instantaneously, that
is, within the duration of the laser ash (20 ns), at least
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
ions of model compounds for photocatalytic oxidation like
dichloroacetate or thiocyanate. Shallowly trapped holes, on
the other hand, are in a thermally activated equilibrium with
free holes which exhibit a very high oxidation potential. The
overall yield of trapped holes can be considerably increased
when small platinum islands are present on the TiO2 surface, which act as efcient electron scavengers competing
with the undesired e /h+ recombination. While molecular
oxygen, O2 , reacts in a relatively slow process with trapped
electrons, the adsorption of the model compounds on the
TiO2 surface prior to the bandgap excitation appears to be
a prerequisite for an efcient hole scavenging.

6.4. Enhanced Photocatalytic Activity


Via Surface Modications
A driving force for research in heterogeneous photocatalysis using TiO2 (and semiconductors in general) is the creation and application of systems capable of using natural
sunlight to degrade a variety of organic and inorganic contaminants in wastewater or polluted air. As mentioned, the
photocatalytic activity depends strongly, among other factors, on the wavelength range response. Since the bandgap
of TiO2 is 3.2 eV, it is active only in the ultraviolet region
which amounts to <10% of the overall solar intensity. Principally, there are several remedies to circumvent (at least
partially) this limitation: (i) Deposition of metals on the
semiconductor; (ii) using multicomponent semiconductors;
(iii) surface modication with sensitizing dyes. The merits of
these options will be outlined briey in the following.

Schottky
Barrier
metal

semiconductor

Figure 9. Metal-modied semiconductor photocatalyst particle.

activity. The Pt/TiO2 system is probably the most frequently


studied metal-semiconductor pair (see, e.g., [350, 351]);
Figure 10 exemplies the enhancement of the photocatalytic activity of nanocrystalline TiO2 by platinization
[350]: three commercially available TiO2 -catalysts, namely,
Degussa P25, Sachtleben Hombikat UV100, and Millennium
TiONA PC50, were platinized by a photochemical impregnation method with two ratios of platinum deposits (0.5 and
1 wt%). The photocatalytic activities of these samples were
determined using three different model compounds, EDTA,
4-chlorophenol (4-CP), and dichloroacetic acid (DCA). Platinization resulted in all cases in an enhancement of the
activity; Figure 10 shows the degradation of DCA as a function of illumination time (light intensity 23 W/m2 at a wavelength of 320400 nm) for pure TiO2 and the two platinized
specimens. After 2 h of illumination, the initial concentration of 120 ppm total organic carbon (TOC) was reduced
to 2.3 ppm at pH 3 employing the best photocatalyst, in
this case, Hombikat UV100 with 0.5 wt% Pt. For this system, an initial photonic efciency (i.e., number of degraded
molecules per number of incident photons) of 43% was
obtained [350]. Apart from platinum, an enhanced photocatalytic activity has been also noted for other metals
and semiconductors. Their inuence on the photocatalytic
activities has been studied in detail, for example, utilizing

1.0

without Pt
0.5 wt% Pt
1 wt% Pt

0.8

TOC/TOC0

6.4.1. Deposition of Metals on the Surface


The selectivity and efciency of a photochemical reaction
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 [345348] and
therefore, there is, no need to duplicate it here. In terms
of photocatalytic activity, a drastic enhancement has, for the
rst time, been observed for the photocatalytic conversion
of water into hydrogen and oxygen upon a fractional coverage of the TiO2 surface with platinum [349]. After excitation
the electron migrates to the metal where it becomes trapped
and electron-hole recombination is suppressed. The hole is
then free to diffuse to the semiconductor surface where oxidation of organic species can occur. These processes are
schematically depicted in Figure 9. The presence of the
metal can be benecial also because of its own catalytic

0.6

0.4

0.2

0.0

20

40

60

80

100

120

illumination time (min)


Figure 10. Degradation of dichloroacetic acid (expressed as the relative
change of TOC) with platinized TiO2 in comparison to pure TiO2 as a
function of illumination at pH 3. Data from [350], D. Hufschmidt et al.,
J. Photochem. Photobiol. A: Chem. 148, 223 (2002).

523

Nanocrystalline TiO2 for Photocatalysis

the following systems: Au-modied TiO2 [352, 353], silvermodied titanium particles [354], transition-metal doped
TiO2 photocatalysts [355357], or rare-earth-doped TiO2
nanoparticles [358].
It should be noted in this context that various analytical
techniques like transmission electron microscopy or scanning force microscopy are often very useful in determining
the size of the particles and their distribution in the bulk
and at the surface of these nanocrystalline materials.

6.4.2. Composite Semiconductors


The advantage of composite semiconductors is usually
twofold: rst, to extend the photo-response by coupling
a large bandgap semiconductor with another featuring a
smaller one and, second, to retard the recombination of
photo-generated charge carriers by injecting electrons into
the lower lying conduction band of the large bandgap material. Two types of geometries have been employed: capping the nanocrystallites of one semiconductor with those
of the second or bringing the nanocrystalline particles of
the two materials into intimate contact. The principle of
charge exchange and separation for both arrangements is
illustrated in Figure 11. Let us consider the case of coupling CdS with TiO2 ; the energy of the excitation light is
too small to directly excite the TiO2 particle, but it is large
enough to excite an electron from the valence band across
the bandgap of CdS (Eg = 25 eV) to the conduction band.
According to the energetics, the hole produced in the CdS
valence band remains in the CdS particle, whereas the electron transfers to the conduction band of the TiO2 particle;
this increases the charge separation and efciency of the
photocatalytic process. The separated electrons and holes
are then free to undergo electron transfer with adsorbates at
the surface. While the mechanisms of charge separation in a
capped system are similar, in a capped semiconductor only
one of the charge carriers is accessible at the surface for catalytic reactions. Several semiconductors have been studied
CB
TiO2

CB SnO2

VB

h'

VB +

A
A+

(a)

TiO2

CdS
h

CB -

thoroughly in combination with TiO2 : TiO2 CdS [359361],


TiO2 CdSe [362], TiO2 coupled SnO2 [363], TiO2 capped
SiO2 [364], and some mixed-oxide systems like Fe2 O3 TiO2
[365] or ZrO2 TiO2 [366].

6.4.3. Surface Modication


with Sensitizing Dyes
Surface sensitization of a wide bandgap semiconductor photocatalyst like TiO2 via adsorbed dyes can increase the efciency of the excitation step and, as a consequence, the
activity. This process can also expand the wavelength range
of excitation. Some common dyes that are used as sensitizers include erythrosin B, eosin, rhodamines, cresyl violet, thionine, porphyrins, [Ru(bpy)3 ]2+ and its analogues,
and many others (see, e.g., [102] for a more comprehensive list). The individual charge-transfer and excitation steps
involved in the dye sensitizer surface process are exemplied
in Figure 12. The primary excitation of an electron in the
dye molecule occurs in either the singlet or triplet excited
state of the molecule; if the oxidative energy level of the
excited state of the dye molecule, with respect to the conduction band energy level of the semiconductor, is more
negative, then the dye molecule can transfer the electron
to the conduction band of the semiconductor. The surface
accepts electrons from the dye molecules which, in turn,
can be transferred to reduce an organic acceptor molecule
adsorbed on the surface. The dye-sensitized semiconductor can also be used in oxidative degradation of the dye
molecule itself after charge transfer; this appears to be an
important process in view of the large quantities of dye substances in wastewater from the textile industries and others.
In passing, it can be noted that the process of utilizing subbandgap excitation with dyes that absorb strongly in the visible for photosensitization is frequently employed in color
photography and other imaging science applications. This
approach of light-energy conversion is also similar to plant
photosynthesis, in which chlorophyll molecules act as light
harvesting antenna molecules.
Generally, the high porosity and strong bonding character of nanostructured TiO2 (and other) semiconductor
lms facilitate the surface modication with organic dyes
or organometallic complexes. For example, photoconversion
efciencies in the range 1015% in diffused daylight have
been reported [367] for nanostructured TiO2 lms modied
with a ruthenium complex. The charge injection from a singlet excited sensitizer into the conduction band of a large
bandgap semiconductor is thought to be an ultrafast process

B
-

CdS CB

S*

B-

CB

h
S

h'

VB +

VB

A+

CB

S+

CB

A
VB

S+

AVB

VB
(b)

Figure 11. Photo-induced charge separation in composite semiconductor particles: (a) capped and (b) coupled semiconductor nanocrystallites. Photo-generated charge carriers move in opposite directions.

Figure 12. Sequence of excitation and charge-transfer steps using a dyemolecule sensitizer. In the rst step, the sensitizer (S) is excited by an
incident photon of energy h and an electron is transferred into the
conduction band of the semiconductor particle; the electron then can
be transferred to reduce an organic acceptor molecule (A) adsorbed on
the surface.

524

Nanocrystalline TiO2 for Photocatalysis

occurring at a picosecond time scale; in the case of different


organic dyes, it has been shown [368, 369] that charge injection takes place within 20 ps. A similar fast electron transfer
has also been noted for [Ru(H2 O)2 ]2 on a TiO2 surface at
very low coverage [370]. Progress in femtosecond laser spectroscopy opened a venue for investigations on even much
shorter time scales. In fact, in recent studies charge carrier injection times in the range of 20200 fs were reported
[371375] for various dyes on nanocrystalline TiO2 particles.
Signicant enhancement effects of electron acceptors
(additives) such as hydrogen peroxide, ammonium persulfate, potassium bromate, and potassium peroxymonosulfate (oxone) on the TiO2 photocatalytic degradation
of various organic pollutants were observed already in
early investigations [376]. The results showed that these
additives markedly improved the degradation rate of 2,4dichlorophenol. The enhanced photocatalytic oxidation of
sulde ions on phthalocyanine modied titania was ascribed
[377] to the additional formation of superoxide radicals.
Sensitization of wide bandgap semiconductor electrodes
by dyes absorbing visible light are routinely used also in dyesensitized photoelectrochemical cells, in order to achieve
high photon-to-current conversion efciencies [378, 379].
The preparation and dynamics of interfacial photosensitized charge separation in metal oxides such as TiO2 lms
has been reviewed [70, 71, 380]. Principally, the photophysical reactions occurring in those dye-sensitized injection solar cells, which are based on a dye adsorbed onto
a porous TiO2 layer, are very similar to those relevant to
photocatalysis. Because of their importance as an energy
source, Figure 13 presents a schematic drawing [69] of such
TCO glass
E

dye
e-

S+/ S*

TCO glass with Pt


h

4
e-

S+/ S

e-

I- / I3-

5
e-

por-TiO2 electrolyte I- / I3e-

load

e-

Figure 13. Schematic outline of a dye-sensitized photovoltaic cell,


showing the electron energy levels in the different phases. The system
consists of a semiconducting nanocrystalline TiO2 lm onto which a
Ru-complex is adsorbed as a dye and a conductive counterelectrode,
while the electrolyte contains an I /I3 redox couple. The cell voltage observed under illumination corresponds to the difference, V ,
between the quasi-Fermi level of TiO2 and the electrochemical potential
of the electrolyte. S, S , and S+ designate, respectively, the sensitizer,
the electronically excited sensitizer, and the oxidized sensitizer. See text
for details. Adapted from [69], A. Hagfeldt and M. Grtzel, Chem Rev.
95, 49 (1995). 1995, American Chemical Society.

a dye-sensitized nanocrystalline TiO2 solar cell, depicting


the relevant energy levels and the pathway for the photoexcited electrons. In this specic example, a ruthenium complex [367] was adsorbed as a dye onto the TiO2 and an
I /I
3 redox couple was used in the electrolyte. Contrary to
conventional semiconductor devices, in the dye-sensitized
cells the function of light absorption is separated from the
charge-carrier transport. The Ru-complex has to absorb the
incident sunlight and to effect, via this energy, the electron1 and 
2 in Fig. 13). Apart from
transfer reaction (numbers 
supporting the dye, the TiO2 lm acts as an electron acceptor and electronic conductor: the electrons injected into the
TiO2 conduction band drift across the nanocrystalline lm
to the conducting glass support which functions as current
3 in Fig. 13). At the counterelectrode, the eleccollector (
4 )
tron is transferred to the redox couple in the electrolyte (
5 ). The cell
which, in turn, serves to regenerate the dye (
voltage observed under illumination, V , is determined by
the difference between the Fermi-level of TiO2 and the electrochemical potential of the electrolyte (cf. Fig. 13 [69]).

7. PHOTOCATALYTIC APPLICATIONS
OF NANOCRYSTALLINE TiO2
While many examples of the photocatalytic activity of
nanocrystalline TiO2 have been presented already in the
foregoing sections, the following discussions will focus on
novel and important (and sometimes large-scale) applications. A very prominent area appears to be environmental
catalysis [31, 32, 45] which, in recent years, has expanded
considerably beyond the traditional elds like NOx removal
from stationary and mobile sources, or the conversion of
volatile organic compounds (VOC). According to [381],
these potential new areas include:
(i) catalytic technologies for liquid or solid waste
reduction or purication;
(ii) use of catalysts in energy-efcient catalytic technologies and processes;
(iii) reduction of the environmental impact in the use or
disposal of catalysts;
(iv) new ecocompatible renery, chemical or nonchemical catalytic processes;
(v) catalysis for greenhouse gas control;
(vi) use of catalysts for user-friendly technologies and
reduction of indoor pollution;
(vii) catalytic processes for sustainable chemistry;
(viii) reduction of the environmental impact of transport.
Hence, (photo)catalysis in environmental applications can
be instrumental in promoting the quality of life and environment, in promoting a more efcient use of resources, and in
promoting sustainable processes and products.
Because of the tremendous importance of those environmentally related areas, the use of nanocrystalline TiO2 for
such photocatalytic applications is illustrated in this section
by means of selected examples. Before doing so, it is stressed
that there exists at least one other rather important issue
in this context. In fact, hydrogen production from aqueous
solutions using semiconductor particles such as, CdS, TiO2 ,
WO3 as photocatalysts is envisaged to become a potential

525

Nanocrystalline TiO2 for Photocatalysis

major application of these materials, and new concepts and


approaches are developed continuously. For example, a new
photocatalytic reaction that splits water into H2 and O2
was designed [382] by a two-step photo-excitation system

composed of a IO
3 /I shuttle redox mediator and two different TiO2 photocatalysts, Pt-loaded TiO2 -anatase for H2
evolution and TiO2 -rutile for O2 evolution. Simultaneous
gas evolution of H2 (180 mol/h) and O2 (90 mol/h) was
observed from a basic (pH = 11) NaI aqueous suspension
of two different TiO2 photocatalysts under UV irradiation
( > 300 nm, 400 W high-pressure Hg lamp). An extensive review [383] assesses photocatalytic efciencies with reference to hydrogen production by means of light energy
in the presence and absence of loaded metals, electrondonors/acceptors, and hole scavengers.

whereas it inhibited that of acetone. As for the effect of photon ux, it was found that photocatalytic degradation occurs
in two regimes with respect to photon ux: for illumination levels distinctly blow 10002000 W/cm2 , the photocatalytic degradation rate increased linearly with photon ux,
whereas for power densities above that value, the rate was
found to scale with the square root of the ux. Figure 14
shows some of those data [388], depicting in panel (a), the
degradation of methanol as a function of UV illumination
time for ve different initial concentrations. (Using TiO2
anatase nanocrystallites with 7 nm diameter in a solution,
in this work photocatalytic TiO2 lms were deposited onto
glass substrates by dip-coating.) The reaction kinetics were
found to follow the L-H model, in which the reaction rate r
varies proportionally with the surface coverage  according
to

7.1. Reduction/Removal of Toxic Gases


r = k =

kKc
1 + Kc

(9)

concentration (103 mol/m3)

where c is the concentration of the VOC and k and K are,


respectively, the reaction rate constant and the adsorption
equilibrium constant. Figure 14(b) exemplies this nding,
showing the initial reaction rates r0 derived from data like
those in Figure 14(a) as a function of the respective initial
methanol concentrations c0 . The solid line in Figure 14(b)
is a t to the data according to Eq. (9).

initial reaction rate r0 (103 mol/m3min)

The conversion of nitrogen oxides to less toxic compounds is


important both because of their toxicity and the global atmospheric pollution. NOx can be converted to N2 and other
nitrogen compounds by reduction. TiO2 -loaded zeolites and
the vanadium silicate-1 were found [384] to decompose NO
under irradiation, in particular, TiO2 included in zeolite cavities results in complete decomposition into N2 and O2 . Titanium oxide catalysts prepared within the Y-zeolite cavities
via an ion-exchange method exhibit [385] high and unique
photocatalytic reactivities for the decomposition of NO into
N2 and O2 , as well as the reduction of CO2 with H2 O showing a high selectivity for the formation of CH3 OH. It was
also found that the charge transfer excited state of the titanium oxide species, (T3+ -O ) , plays a vital role in these
unique photocatalytic reactions. In yet another approach, an
efcient catalytic reduction of NO at low temperature by
means of NH3 could be achieved using Mn-, Cr-, or Cuoxides on a nanocrystalline TiO2 support [386].
The NOx removal process was studied experimentally in
a pulsed corona discharge combined with the TiO2 photocatalytic reaction [387]. NO2 was found to adsorb easily on
the photocatalyst surface, whereas NO was hardly adsorbed.
Addition of water vapor enhanced the NO2 adsorption. It
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
surface, and is transformed to HNO3 through photocatalytic
reaction with OH.
The photocatalytic degradation of VOCs in the gas phase
constitutes another very important example in this range of
applications. Utilizing variously prepared TiO2 photocatalysts (e.g., deposited on glass ber cloth, as pellets or as
thin lms), the photo-induced reactions of trichloroethylene,
acetone, methanol, and toluene were investigated [388390].
The photocatalytic degradation rate was observed [388] to
increase with increasing initial concentration of the VOCs,
but remained almost constant beyond a certain concentration. It matched well with the LangmuirHinshelwood
(L-H) kinetic model [11]. For the inuence of water vapor
in a gas-phase photocatalytic degradation rate, there was an
optimum concentration of water vapor in the degradation of
trichloroethylene and methanol. Furthermore, water vapor
enhanced the photocatalytic degradation rate of toluene,

20
(a)

methanol
15
10
5
0

10

illumination time (min)

2.0
1.5
1.0
0.5
(b)

0.0

10

15

20

25

initial concentration c0 (103 mol/m3)

Figure 14. (a) Photocatalytic degradation of methanol with different


initial concentrations as a function of UV illumination time (light intensity 2095 W/cm2 at a wavelength of 254 nm) at a H2 O concentration
of 0.38 mol/m3 and a reaction temperature of 45  C. (b) Initial reaction
rates r0 as derived from the data in (a) versus the initial methanol concentrations; the solid line is a t according Eq. (9). Data from [388],
S. B. Kim and S. C. Hong, Appl. Catal. B: Environ. 35, 305 (2002).

526

The degradation of organic compounds is probably the


most widely used photocatalytic application of nanocrystalline TiO2 and other semiconductor materials. In an aqueous environment, the holes created under UV irradiation
are scavenged by surface hydroxyl groups to generate OH
radicals that then promote the oxidation of organics. This
radical-mediated oxidation has been successfully employed
in the mineralization of several hazardous chemical contaminants such as hydrocarbons, haloaromatics, phenols, halogenated biphenyls, surfactants, and textile and other dyes
[102].
The possible photocatalytic decomposition of a broad
range of organic compounds has been investigated using
nanocrystalline TiO2 particles. Detailed studies reported
the oxidation of dissolved cyanide [391], the degradation
of various kinds of acids [392398], and of several herbicides [399402], for the photocatalytic oxidation of toluene,
benzene, cyclohexene, and benzhydrol [403406] or for
the 1,1 -dimethyl-4,4 -bipyridium dichloride decomposition
[407]. In another application, a titanium oxide photocatalyst of ultra-high activity has been employed for the selective
N-cyclization of an amino acid in aqueous suspensions [408].
Anatase crystallites of average diameter of 15 nm were
platinized by impregnation from aqueous chloroplatinic acid
solution followed by hydrogen reduction. The catalyst was
suspended in an aqueous L-lysine (Lys) solution and photoirradiated under argon at ambient temperature to obtain
L-pipecolinic acid.
The photocatalytic degradation and oxidation of phenol
and phenol-based compounds has been examined quite frequently [409414]. The decomposition of aqueous phenol
solutions to carbon dioxide have been studied using natural
sunlight in geometries simulating shallow ponds [415]. The
photocatalyst was titanium dioxide freely suspended in the
solution or immobilized on sand or silica gel. Photodegradation rates were approximately three times faster with the
free suspension than with the immobilized catalyst under
the same conditions, and were dependent on the time of
year and the time of day. The seasonal variation correlated
roughly with seasonal solar irradiance tabulations for the
UV component of the spectrum. For 10 ppm of phenol, the
maximum rate of solar degradation resulted in a decrease
in concentration to 10 ppb in less than 80 min with total
mineralization in 110 min.
An efcient degradation of aqueous phenol was achieved
[416] by a new rotating-drum reactor coated with a TiO2
photocatalyst, in which TiO2 powders loaded with Pt are
immobilized on the outer surface of a glass drum. The reactor can receive solar light and oxygen from the atmosphere
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 concentration of 22.0 mg/dm3 was decomposed within 60 min
and was completely mineralized through intermediate products within 100 min.
The photocatalytic degradation of various types of dyes
appears to be another prominent and extensively explored
application of nanocrystalline TiO2 in environmental catalysis [417423].

In a recent study [424], the photocatalytic degradation of


ve dyes in TiO2 aqueous suspensions under UV irradiation
has been investigated; it was attempted to determine the
individual steps of such a degradation process by varying the
aromatic structures, using either anthraquinonic (Alizarin
S (AS)), or azoic (Crocein Orange G (OG), Methyl Red
(MR), Congo Red (CR)) or heteropolyaromatic (Methylene Blue (MB)) dyes. Figure 15 exemplies the photocatalytic degradation of three of these dyes (CR, OG, and MR)
as a function of UV irradiation. The initial reaction rates
were found to fall in the range from 1.9 mol/l min (for
CR) to 3.6 mol/l min (for MR). In addition to a prompt
removal of the colors, TiO2 /UV-based photocatalysis was
simultaneously able to fully oxidize the dyes, with a complete mineralization of carbon into CO2 . Sulfur heteroatoms
were converted into innocuous SO2
4 ions. The mineralization of nitrogen was more complex. Nitrogen atoms in the
3-oxidation state, such as in amino groups, remain at this
reduction degree and produced NH+
4 cations, subsequently
and very slowly converted into NO
3 ions. For azo-dye (OG,
MR, CR) degradation, the complete mass balance in nitrogen indicated that the central
N N
azo group was
converted into gaseous dinitrogen, which is the ideal issue
for the elimination of nitrogen-containing pollutants. The
aromatic rings were submitted to successive attacks by photogenerated OH radicals leading to hydroxylated metabolites before the ring opening and the nal evolution of CO2
induced by repeated reactions with carboxylic intermediates.
The photocatalytic degradation of acid derived azo dyes
in aqueous TiO2 suspensions follows apparently rst-order
kinetics [425, 426]. The site near the azo bond (C N Nbond) is the attacked area in the photocatalytic degradation process, while the TiO2 photocatalytic destruction of
the C N( ) bond and N N bonds leads to fading
of the dyes. The pH effect on the TiO2 photocatalytic degradation of the acid-derived azo dyes varies with dye structure. Hydroxyl radicals play an essential role in the ssion
of the C N N conjugated system in azo dyes in TiO2
photocatalytic degradation. Metalized azo dyes were studied
80

concentration (mol/l)

7.2. Degradation of Organic Compounds

Nanocrystalline TiO2 for Photocatalysis

CR
OG
MR

60

40

20

50

100

150

200

illumination time (min)


Figure 15. Photocatalytic degradation of three different dyes, Congo
Red (CR), Crocein Orange (OG), and Methyl Red (MR), given in
terms of the concentration versus the time of illumination. Data from
[424], H. Lachheb et al., Appl. Catal. B: Environ. 39, 75 (2002).

Nanocrystalline TiO2 for Photocatalysis

[427] under TiO2 photocatalytic and photosensitized conditions in aqueous buffering solutions. The size and strength of
intramolecular conjugation determines apparently the lightfastness of the dyes; the more powerful OH radicals in TiO2
photocatalytic process are highly reactive towards the azo
dyes.

7.3. Wastewater and Soil Remediation


The major causes [428] of surface water and groundwater contamination are industrial discharges, excess use of
pesticides, fertilizers (agrochemicals), and landlling domestic wastes. Typically, the wastewater treatment is based
upon various mechanical, biological, physical, and chemical
processes. After ltration and elimination of particles in suspension, the biological treatment is the ideal process (natural decontamination). Unfortunately, organic pollutants are
not always biodegradable; a promising approach then relies
on the formation of highly reactive chemical species, which
degraded the more recalcitrant molecules into biodegradable compounds. These are called the advanced oxidation
processes (AOPs). Although there exist differences in their
detailed reaction schemes, their common feature is the production of OH radicals ( OH); these radicals are extraordinarily reactive species (oxidation potential 2.8 V). They are
also characterized by a low selectivity of attack, which is a
useful attribute for an oxidant used in wastewater treatment
and for solving pollution problems. These photocatalytic
degradation of wastewater employing nanocrystalline TiO2
has been examined in various set-ups [429] and pilot-plant
scale solar photocatalytic experiments have been realized
[428].
Several recent studies reported on the removal or reduction of metals or metal-containing contaminants in wastewater, based on the principles outlined in the foregoing
paragraph. Those investigations examined, for example, the
removal of cadmium and mercury from water using modied TiO2 nanoparticles [430, 431], the radical, mediated
photo-reduction of manganese ions in UV-irradiated titania
suspensions [432], the simultaneous photocatalytic Cr(VI)
reduction and dye oxidation in a TiO2 slurry reactor [433],
or the removal of iron(III) cyanocomplexes [434].
While the efcient use of a photocatalytic process in the
presence of TiO2 to degrade many different types of pollutants in wastewater has been conrmed repeatedly, the
question of how to efciently separate and reuse TiO2 from
treated wastewater became a notable problem in the application of a TiO2 photo-oxidation process. A recent study
[435] aimed to develop an advanced process for dyeing
wastewater treatment, in which dyeing wastewater was initially treated by an intermittently decanted extended aeration (IDEA) reactor to initially remove biodegradable
matters and further treated in a TiO2 photocatalytic reactor for complete decolorization and high chemical oxygen
demand (COD) removal. Suspended TiO2 powder used in
the photo-oxidation was separated from slurry by a membrane lter and recycled to the photo reactor continuously.
Photocatalytic destruction of chlorinated solvents in water
with solar energy was investigated [436] using a nearcommercial scale, single-axis tracking parabolic trough system with a glass pipe reactor mounted at its focus. In

527
the photocatalytic degradation of industrial residual waters,
the use of peroxydisulfate (S2 O2
8 ) as an additional oxidant
(electron scavenger) was observed to have an outstanding
effect, producing an important increase in the degradation
rate [437]. The impact of pH and the presence of inorganic
ions and organic acids commonly found in natural waters on
rates of TiO2 photocatalyzed trinitrotoluene (TNT) transformation and mineralization was examined [438]. Raising
the pH slightly increased the rate of TNT transformation,
primarily as a result of an increased rate of TNT photolysis, but signicantly reduced rates of mineralization due to
increased electrostatic repulsion between the catalyst surface
and anionic TNT intermediates. The presence of inorganic
anions did not substantially hinder TNT transformation at
alkaline pH, but mineralization rates were diminished when
the anion either adsorbed strongly to the photocatalyst or
was an effective hydroxyl radical scavenger.
Immobilized TiO2 photocatalysts were used to sterilize
and reclaim the wastewater of bean sprout cultivation from
a continuous hydrocirculation system [439]. The photocatalysts effectively killed bacteria and degraded organic pollutants in the wastewater. Stimulation of bean sprout growth
and suppression of decaying pathogens were also induced by
the TiO2 photocatalytic activity.
Photocatalytic decomposition of seawater-soluble crude
oil fractions using high surface area colloid nanoparticles
of TiO2 under UV irradiation was explored [440]; although
no mineralization occurred due to photolysis, important
chemical changes were observed in the presence of TiO2 ,
with the degradation reaching 90% (measured as dissolved
organic carbon, (DOC)) in waters containing 945 mg C/l of
seawater-soluble crude oil compounds after 7 days of articial light exposure. During light exposure, transient intermediates that showed higher toxicity than the initial compounds
were observed, but were subsequently destroyed. Heterogeneous photocatalysis using TiO2 was considered to be a
promising process to minimize the impact of crude oil compounds on contaminated waters.
TiO2 -photocatalytic degradation of a cellulose efuent
was evaluated [441] using multivariate experimental design.
The efuent was characterized by general parameters such
as adsorbable organic halogens (AOX), TOC, COD, color,
total phenols, acute toxicity, and by the analysis of chlorinated low molecular weight compounds using GC/MS. The
optimal concentration of TiO2 was found to be around 1 g/l.
After 30 min of reaction more than 60% of the toxicity was
removed and after 420 min of reaction, none of the initial
chlorinated low molecular weight compounds were detected,
suggesting an extensive mineralization.
Photocatalysts, based on titanium dioxide, were used for
the purication of contaminated soil polluted by oil [442].
Commercially produced slurry of titanium dioxide was modied with barium, potassium, and calcium. The experiments
were performed under natural conditions in summer months
(July and August) applying direct solar-light irradiation. The
most active photocatalyst for soil purication was titanium
dioxide modied with calcium.
Two different photocatalysts, namely, Hombikat UV100
(Sachtleben Chemie) and P25 (Degussa) have been used in
batch experiments [443] to compare their ability to degrade
the toxic components of a biologically pretreated landll

528

Nanocrystalline TiO2 for Photocatalysis

leachate. A strong adsorption of the pollutant molecules was


observed for both TiO2 -powders, with a maximum of almost
70% TOC reduction for Hombikat UV100.

7.4. Purication of Drinking Water

1000

100

800

80

600

60

400

40

200

20

20

40

60

protein phosphatase inhibitor (%)

concentration microcystin-LR (g/ml)

Pathogens in drinking water supplies can be removed by


sand ltration followed by chlorine or ozone disinfection.
These processes reduce the possibility of any pathogens
entering the drinking water distribution network. However, there is doubt about the ability of these methods
to remove chlorine-resistant microorganisms including protozoan oocysts. Titanium dioxide (TiO2 ) photocatalysis is
a possible alternative/complementary drinking water treatment method and several studies [444, 445] reported a
strong and swift photocatalytic inactivation of bacteria and
bacteriophages in aqueous solutions. For example, the rate
of disinfection was explored using TiO2 electrodes prepared
by the electrophoretic immobilization of TiO2 powders with
different crystallinity. These electrodes were tested for their
photocatalytic bactericidal efciency with E. coli K12 as a
model test organism [446]. Similar studies were reported for
natural water from a river [447].
Cyanobacterial toxins produced and released by
cyanobacteria in freshwater around the world pose a
considerable threat to human health if present in drinking
water sources. Therefore, various treatments have been
applied to remove these toxins. The effectiveness of TiO2
photocatalysis for the removal of microcystin-LR from
water has been established [448]. Not only does the process
rapidly remove the toxin but also the by-products appear to
be nontoxic. The photocatalytic process has also signicantly
reduced the protein phosphatase 1 (PP1) inhibition. Protein
phosphatase 1 inhibition is potentially one of the most
serious harmful effects to humans who may consume water
contaminated by microcystins. Figure 16 shows some of
these data, namely, the reduction of the microcystin-LR
concentration and the PP1 inhibition as a function of the
illumination time. The results indicate that about 86% of

illumination time (min)


Figure 16. Destruction and protein phosphatase (PP1) inhibition of
microcystine-LR via TiO2 photocatalysis as a function of the duration
of UV illumination (xenon lamp with 480 W at a wavelength of 330
450 nm). Data from [448], I. Liu et al., J. Photochem. Photobiol. A:
Chem. 148, 349 (2002).

microcystin-LR was destroyed within the rst 5 min of


photocatalysis, with 97% of the toxin removed in 20 min.
The addition of 0.1% H2 O2 to the photocatalytic system
was found [448] to further enhance the degradation rate:
99.6% of microcystin-LR was destroyed within 5 min and
no toxin was left after 10 min of photocatalysis.
Photocatalytic inactivation of different bacteria and bacteriophages in drinking water at different TiO2 concentrations
with or without concurrent exposure to O2 was studied in
[449] using UV irradiation (5.5 mW/cm2 at 365 nm). For
example, for this light intensity, the most effective inactivation of Escherichia coli CN13 was obtained at 1 g/l suspension of TiO2 , resulting in a reduction by ve orders of
magnitude in 5 min. Under the same experimental conditions, MS2 bacteriophage was reduced by four orders of
magnitude, also in 5 min. The addition of O2 into the experimental environment increased the inactivation of Deinococcus radiophilus by four orders of magnitude in 60 min.

7.5. Miscellaneous Photocatalytic


Applications
It may have become apparent from the foregoing discussions and examples that the solution of environmental
problems constitutes one of the (if not the) major driving
forces in research and development in photocatalysis using
nanometer-sized TiO2 (and other semiconductor) particles.
Another one, of course, is the production of hydrogen from
water splitting. Apart from these main applications, there
exist, on the other hand, many attempts to explore novel
areas for the photocatalytic use of nanocrystalline TiO2
materials. To give a avor of the diversity of these efforts,
some selected (and mostly recent) examples follow.
Nano-sized titanium oxide (TiO2  thin lms have been
explored for alcohol-sensing applications. TiO2 thin lms
with different doping concentrations were prepared on alumina substrates [450] using the solgel process using the
spin-coating technique for ethanol and methanol alcohol.
Experimental results indicated that the sensor is able to monitor alcohols selectively at ppm levels; the lms are stoichiometric with carbon as the dominant impurity on the surface.
The morphologies and crystalline structures of the lms were
studied by scanning electron microscopy (SEM) and XRD.
X-ray diffraction patterns showed that the lms are pure
anatase phase up to an annealing temperature of 600  C.
As the annealing temperature increased to 800  C, a small
amount of rutile phase formed along with the anatase phase.
Optical waveguides were prepared by depositing a solgelderived titania lm onto a silica substrate [451]. The titania
lm is mesoporous, with pore sizes ranging from 3 to 8 nm.
Deposition of the titania does not change the critical angle
of total internal reection. Thus, the titania-coated waveguides propagate light in an attenuated total reection mode,
despite the relatively high refractive index (n = 1.8 in air) of
the titania lm relative to the silica substrate (n = 05).
The light output of electric lighting gradually decreases
due to stain buildup on lamps and covers during operation.
Roadway, and especially tunnel lighting, experiences a large
amount of contamination due to dust, carbon particles found
in vehicle engine exhausts and other airborne contaminants,

529

Nanocrystalline TiO2 for Photocatalysis

which results in the rapid deterioration of the light output.


Photocatalytic reactions caused by TiO2 are known to decompose such stains. This reaction is caused by the absorption
of UV light ( 400 nm, corresponding to the bandgap
of 3 eV) irradiated from lamps or the sun, and followed
by oxidation. Extensive eld tests revealed [452] that a ne
lm coating of TiO2 on lamps and luminaires can effectively decompose various organic compounds such as vehicle
exhaust gases, oil, nicotine, etc. This leads to an improvement
of the luminous performance of installed lighting systems and
reduces the cost of maintenance by approximately one-half.
It has been recently found [453] that photocatalytic TiO2
coated with polycarbonate (PC) releases a huge amount of
exothermic energy in the temperature range between 200
and 400  C (ca. 1.85 kJ/g). The strong interaction between
oxygen-decient sites in TiO2 and carbonyl groups of PC
mediated by a good PC solvent is found to be a prerequisite for a release of the enormous amount of exothermic
energy. This nding suggests that PC-coated TiO2 powders
or related oxides work as a combustion-assisting agent in a
relatively lower temperature range and can be utilized for
incineration applications in order to suppress the formation
of extremely toxic dioxins.
A somewhat unusual application reported [454] the photocatalytic 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
a solution including gold ions is subject to optical exposure, the excited electrons in the conduction band reduce
gold ions into gold metal. Illumination by an evanescent
wave generated with a total reection conguration limits
the deposition region to the very tip. In the experiments,
100300 nm gold particles on SiN cantilever tips for atomic
force microscopes were obtained. In a related vein, photoinduced deposition of copper on nanocrystalline TiO2 lms
was proposed [455].
Solar photocatalytic oxidation processes (PCO) for degradation of water and air pollutants have received increasing
attention. In fact, some eld-scale experiments have demonstrated the feasibility of using a semiconductor (TiO2  in solar
collectors and concentrators to completely mineralize organic
contaminants in water and air [456]. Although successful preindustrial solar tests have been carried out, there are still discrepancies and doubt concerning process fundamentals such
as the roles of active components, appropriate modelling of
reaction kinetics, or quantication of photo-efciency. Challenges to development are catalyst deactivation, slow kinetics, low photo-efciency and unpredictable mechanisms. The
development of specic nonconcentrating collectors for detoxication and the use of additives such as peroxydisulfate have
made competitive use of solar PCO possible.

GLOSSARY
Charge transfer The transfer of a charge carrier (electron or
hole) from an excited semiconductor to an adsorbed species on
its surface. This transfer may initiate a reaction (oxidation or
reduction) in the adsorbed molecule.
Dye-sensitized semiconductor Adsorbing a suitable dye on
the surface of a wide band gap semiconductor (like TiO2 ) can

enhance the efciency of the excitation step and, hence, the


catalytic activity.
Electron-hole pair The absorption of a photon of sufcient
energy may excite in a semiconductor an electron from the
valence band to the conduction band, thereby creating a hole
in the valence band.
Nanocrystalline A material composed of individual crystallites which have a size in the range of nanometer (nm);
1 nm = 109 m.
Photocatalysis A catalytic reaction triggered or enhanced by
illuminating the system with visible or ultraviolet irradiation.
This reaction involves normally the electronic excitation of the
catalyst via the absorption of photons and an interfacial charge
transfer to an adsorbed species. Typically, the photocatalyst is
not consumed in the reaction.
Photocatalytic degradation The removal or reduction of
(usually unwanted) substances via a photocatalytic reaction.
Quantum yield The probability of product formation per
adsorbed photon in a photocatalytic reaction.
Titanium oxide Titanium dioxide with the nominal composition TiO2 is a semiconductor with a band gap of 3.2 eV; it
exists in three different crystalline modications, two of which
(anatase and rutile) are commonly employed in photocatalysis.

REFERENCES
1. H. S. Nalwa, Ed., Encyclopedia of Nanoscience and Nanotechnology. American Scientic Publisher, Stevenson Ranch CA, 2003, in
press.
2. J. H. Fendler, Ed., Nanoparticles and Nanostructured Films.
Wiley-VCH, Weinheim, 1998.
3. R. W. Siegel, in Physics of New Materials (E. Fujita, Ed.), p. 66.
Springer, Berlin, 1998.
4. G. L. Timp, Ed., Nanotechnology. Springer, New York, 1999.
5. V. A. Shchukin and D. Bimberg, Rev. Mod. Phys. 71, 1125 (1999).
6. H. S. Nalwa, Ed., Handbook of Nanostructured Materials and
Nanotechnology, Vols. 15. Academic, San Diego, 2000.
7. C. Binns, Surf. Sci. Rep. 44, 1 (2001).
8. J. Y. Ying, Ed., Nanostructured Materials, Academic, San Diego,
2001.
9. M. Khler, Nanotechnologie. Wiley-VCH, Weinheim, 2001.
10. H. S. Nalwa, Ed., Nanostructured Materials and Nanotechnology.
Academic, San Diego, 2001.
11. J. M. Thomas and W. J. Thomas, Principles and Practice of Heterogeneous Catalysis. VCH, Weinheim, 1997.
12. J. H. Sinfelt, Surf. Sci. 500, 923 (2002).
13. F. Zaera, Surf. Sci. 500, 947 (2002).
14. J. M. Greenberg, Surf. Sci. 500, 793 (2002).
15. D. A. Williams and E. Herbst, Surf. Sci. 500, 823 (2002).
16. E. Zinner, Ann. Rev. Earth Planet. Sci. 26, 147 (1998).
17. L. Grossman, Geochim. Cosmochim. Acta 36, 597 (1972).
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.
20. B. Levy, J. Electroceram. 1, 239 (1997).
21. M. A. Fox and M. T. Dulay, Chem. Rev. 93, 341 (1993).
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).
24. A. Fuijshima, K. Kobayakawa, and K. Honda, J. Electrochem. Soc.
122, 1487 (1975).
25. A. Mills and S. LeHunte, J. Photochem. Photobiol. A: Chem. 108, 1
(1997).

530
26. A. Fujishima and T. N. Rao, Pure Appl. Chem. 70, 2177 (1998).
27. A. Fujishima, K. Hashimoto, and T. Watanabe, TiO2 Photocatalysis:
Fundamentals and Applications. BKC, Tokyo, 1999.
28. A. J. Bard, J. Phys. Chem. 86, 172 (1982).
29. M. Grtzel, Ed., Energy Resources Through Photochemistry and
Catalysis. Academic, New York, 1983.
30. E. Pelizzetti and M. Schiavello, Eds., Photochemical Conversion
and Storage of Solar Energy. Kluwer Academic Publishers, Dordrecht, 1991.
31. M. Schiavello, Ed., Photocatalysis and Environment. Kluwer Academic Publishers, Dordrecht, 1988.
32. D. F. Ollis and H. Al-Ekabi, Eds., Photocatalytic Purication and
Treatment of Water and Air. Elsevier, Amsterdam, 1993.
33. J. Peral, X. Domnech, and D. F. Ollis, J. Chem. Technol. Biotechnol.
70, 117 (1997).
34. J. H. Fendler, J. Phys. Chem. 89, 2730 (1985).
35. M. Pessarakli, Ed., Handbook of Photosynthesis. Marcel Dekker,
New York, 1997.
36. K. Kalayanasundaram, M. Grzel, and E. Pelizzetti, Coord. Chem.
Rev. 69, 57 (1986).
37. B. A. Perkinson and M. T. Spitlet, Electrochem. Acta 37, 943 (1992).
38. B. ORegan and M. Grtzel, Nature 353, 737 (1991).
39. H. Weller, Angew. Chem. Int. Ed. Engl. 32, 41 (1993).
40. L. N. Lewis, Chem. Rev. 93, 2693 (1993).
41. A. Henglein, J. Phys. Chem. 97, 5457 (1993).
42. C. Bechinger, E. Wirth, and P. Leiderer, Appl. Phys. Lett. 68, 2843
(1996).
43. S. N. Frank and A. J. Bard, J. Am. Chem. Soc. 99, 303 (1977).
44. S. N. Frank and A. J. Bard, J. Phys. Chem. 81, 1484 (1977).
45. M. R. Hoffmann, S. T. Martin, W. Choi, and D. W. Bahnemann,
Chem. Rev. 95, 69 (1995).
46. A. Fujishima, T. N. Rao, and D. A. Tryk, J. Photochem. Photobiol. C:
Photochem. Rev. 1, 1 (2000).
47. J. C. Ireland, P. Klostermann, E. W. Rice, and R. M. Clark, Appl.
Environ. Microbiol. 59, 1668 (1993).
48. J. C. Sjogren and R. A. Sierka, Appl. Environ. Microbiol. 60, 344
(1994).
49. R. X. Cai, Y. Kubota, T. Shuin, H. Sakai, K. Hashimoto, and
A. Fujishima, Cancer Res. 52, 2346 (1992).
50. N. B. Jackson, C. M. Wang, Z. Luo, J. Schwitzgebel, J. G. Ekerdt, J.
R. Brock, and A. Heller, J. Electrochem. Soc. 138, 3660 (1991).
51. H. Gerischer and A. Heller, J. Electrochem. Soc. 139, 113 (1992).
52. E. C. Butler and A. P. Davis, J. Photochem. Photobiol. A: Chem. 70,
273 (1993).
53. Y. Luo and D. F. Ollis, J. Catal. 163, 1 (1996).
54. L. R. Skubal, N. K. Meshkov, and M. C. Vogt, J. Photochem. Photobiol. A: Chem. 148, 103 (2002).
55. C. Wang, J. Rabani, D. W. Bahnemann, and J. K. Dohrmann, J. Photochem. Photobiol. A: Chem. 148, 169 (2002).
56. R. Gao, J. Stark, D. W. Bahnemann, and J. Rabani, J. Photochem.
Photobiol. A: Chem. 148, 387 (2002).
57. K. J. Klabunde, J. Stark, O. Koper, C. Mohs, D. G. Park, S. Decker,
Y. Jiang, I. Lagadic, and D. Zhang, J. Phys. Chem. 100, 12142 (1996).
58. K. N. P. Kumer, K. Keizer, A. J. Burggraaf, T. Okubo, H. Nagamoto,
and S. Morooka, Nature 358, 48 (1992).
59. M. Gopal, W. J. Moberly Chan, and L. C. De Jonghe, J. Mater. Sci.
32, 6001 (1997).
60. S. Ito, S. Inoue, H. Kawada, M. Hara, M. Iwasaki, and H. Tada,
J. Colloid Interface Sci. 216, 59 (1999).
61. K. Terabe, K. Kato, H. Miyazaki, S. Yamaguchi, A. Imai, and
Y. Iguchi, J. Mater. Sci. 29, 1617 (1994).
62. K. S. Suslick, Science 247, 1439 (1990).
63. H. Tada, K. Teranishi, Y. Inubushi, and S. Ito, Chem. Comm. 2345
(1998).
64. S. Kambe, K. Murakoshi, T. Kitamura, Y. Wada, S. Yanagida,
H. Kominami, and Y. Kera, Solar Energy Mater. Solar Cells 61, 427
(2000).

Nanocrystalline TiO2 for Photocatalysis


65. R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima,
A. Kitamura, M. Shimohigoshi, and T. Watanabe, Nature 388, 431
(1997).
66. R. Wang, N. Sakai, A. Fujishima, T. Watanabe, and K. Hashimoto, J.
Phys. Chem. B 103, 2188 (1999).
67. C. Anderson and J. A. Bard, J. Phys. Chem. 99, 9882 (1995); ibid. 101,
2611 (1997).
68. H. Tada, Y. Kubo, M. Akazawa, and S. Ito, Langmuir 14, 2936 (1998).
69. A. Hagfeldt and M. Grtzel, Chem. Rev. 95, 49 (1995).
70. T. Gern, M. Grtzel, and L. Walder, Prog. Inorgan. Chem. 44, 345
(1997).
71. M. Grtzel, Nature 414, 338 (2001).
72. E. Becquerel, Compt. Rend. 9, 145 (1839).
73. U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissrtel, J. Salbeck,
H. Spreitzer, and M. Grtzel, Nature 395, 583 (1998).
74. S. Sdergren, A. Hagfeldt, J. Olsson, and S. E. Lindquist, J. Phys.
Chem. 98, 5552 (1994).
75. G. K. Boschloo and A. Goossens, J. Phys. Chem. 100, 19489 (1996).
76. G. K. Boschloo, A. Goossens, and J. Schoonman, J. Electroanal.
Chem. 428, 25 (1997).
77. D. Cahen, G. Hodes, M. Grtzel, J. F. Guillemoles, and I. Riess, J.
Phys. Chem. B 104, 2053 (2000).
78. L. Kavan, K. Kratochvilov, and M. Grtzel, J. Electroanal. Chem.
394, 93 (1995).
79. P. E. de Jong and D. Vanmaekelbergh, J. Phys. Chem. B 101, 2716
(1997).
80. D. Vanmaekelbergh and P. E. de Jongh, J. Phys. Chem. B 103, 747
(1999).
81. G. Schlichthrl, S. Y. Huang, J. Sprague, and A. J. Frank, J. Phys.
Chem. B 101, 8141 (1997).
82. S. Y. Huang, G. Schlichthrl, A. J. Nozik, M. Grtzel, and A. J.
Frank, J. Phys. Chem. B 101, 2576 (1997).
83. G. Franco, J. Gehring, L. M. Peter, E. A. Ponomarev, and I. Uhlendorf, J. Phys. Chem. B 103, 692 (1999).
84. G. K. Boschloo and D. Fitzmaurice, J. Phys. Chem. B 103, 2228
(1999).
85. B. ORegan, M. Grtzel, and D. Fitzmaurice, Chem. Phys. Lett. 183,
89 (1991).
86. A. Kay, R. Humphry-Baker, and M. Grtzel, J. Phys. Chem. 98, 952
(1994).
87. R. J. Ellingson, J. B. Asbury, S. Ferrere, H. N. Ghosh, J. R. Sprague,
T. Lian, and A. J. Nozik, J. Phys. Chem. B 102, 6455 (1998).
88. J. B. Asbury, R. J. Ellingson, H. N. Ghosh, S. Ferrere, A. J. Nozik,
and T. Lian, J. Phys. Chem. B 103, 3110 (1999).
89. W. S. Struve, Fundamentals of Molecular Spectroscopy. Wiley,
New York, 1989.
90. P. W. Atkins, Molecular Quantum Mechanics, 2nd ed. Oxford University Press, Oxford, 1983.
91. Y. Nosaka and M. A. Fox, J. Phys. Chem. 92, 1893 (1988).
92. S. T. Martin, H. Herrmann, and M. R. Hoffmann, Trans. Faraday
Soc. 90, 3315 and 3323 (1994).
93. M. Grtzel, Heterogeneous Photochemical Electron Transfer.
CRC Press, Boca Raton, FL, 1989.
94. P. Y. Yu and M. Cardona, Fundamentals of Semiconductors.
Springer, Berlin, 2001.
95. G. J. Kavernos and N. J. Turro, Chem. Rev. 86, 401 (1986).
96. P. V. Kamat, Chem. Rev. 93, 267 (1993).
97. W. J. Albery and P. N. Bartlett, J. Electrochem. Soc. 131, 315 (1984).
98. H. Lth, Solid Surfaces, Interfaces and Thin Films. Springer,
Berlin, 2001.
99. L. E. Brus, J. Chem. Phys. 80, 4403 (1984).
100. A. Henglein, J. Phys. Chem. 86, 2291 (1982).
101. D. Duonghong, J. Ramsden, and M. Grtzel, J. Am. Chem. Soc. 104,
2977 (1982).
102. P. V. Kamat, in Handbook of Nanostructured Materials and Nanotechnology (H. S. Nalwa, Ed.), Vol. 3, p. 291. Academic, San Diego,
2000.

Nanocrystalline TiO2 for Photocatalysis


103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.

N. S. Lewis, Annu. Rev. Phys. Chem. 42, 543 (1991).


A. Henglein, Chem. Rev. 89, 1861 (1989).
M. L. Steigenwald and L. E. Brus, Acc. Chem. Res. 23, 183 (1990).
L. Banyai and S. W. Koch, Semiconductor Quantum Dots. World
Scientic, Rivers Edge, NJ, 1993.
L. Jacak, P. Hawrylak, and A. Wjs, Quantum Dots. Springer,
Berlin, 1998.
L. E. Brus, J. Chem. Phys. 79, 5566 (1983).
H. M. Schmidt and H. Weller, Chem. Phys. Lett. 129, 615 (1986).
M. G. Bawendi, M. L. Steigenwald, and L. E. Brus, Annu. Rev. Phys.
Chem. 41, 477 (1990).
H. Tang, H. Berger, P. E. Schmid, and F. Lvy, Solid State Comm. 92,
267 (1994).
H. Tang, F. Lvy, H. Berger, and P. E. Schmid, Phys. Rev. B 52, 7771
(1995).
G. Mie, Ann. Phys. 25, 377 (1908).
C. F. Bohren and D. R. Huffman, Absorption and Scattering of
Light by Small Particles. Wiley, New York, 1983.
U. Kreibig and M. Vollmer, Optical Properties of Metal Clusters.
Springer, Berlin, 1995.
U. Kreibig and C. v. Fragstein, Z. Phys. 224, 307 (1969).
M. M. Alvarez, J. T. Khoury, T. G. Schaaff, M. N. Shagullin, I. Vezmar, and R. L. Whetten, J. Phys. Chem. B 101, 3706 (1997).
G. C. Papavassiliou, Prog. Sol. State Chem. 12, 185 (1979).
P. V. Kamat, Prog. Inorgan. Chem. 44, 273 (1997).
G. Hodes, Isr. J. Chem. 33, 95 (1993).
G. J. Meyer and P. C. Searson, Interface 2327 (1993).
P. V. Kamat, Chemtech, June, 2228 (1995).
N. Uekawa, T. Suzuki, S. Ozeki, and K. Kaneko, Langmuir 8, 1 (1992).
D. V. Paranjape, M. Sastry, and P. Ganguly, Appl. Phys. Lett. 63, 18
(1993).
K. Vinodgopal, U. Stafford, K. A. Gray, and P. V. Kamat, J. Phys.
Chem. 98, 6797 (1994).
A. Hagfeldt, S. E. Lindquist, and M. Grtzel, Sol. Energy Mater. Sol.
Cells 32, 245 (1994).
N. A. Kotov, F. C. Meldrum, and J. H. Fendler, J. Phys. Chem. 98,
8827 (1994).
X.-Z. Ding, Z.-Z. Qi, and Y.-Z. He, J. Mater. Sci. Lett. 14, 21 (1995).
L. Su and Z. Lu, Spectrochim. Acta 53A, 1719 (1997).
X. Z. Ding and X. H. Liu, Mater. Sci. Eng. A 224, 210 (1997).
D. Robert and J. V. Weber, J. Mater. Sci. Lett. 18, 97 (1999).
H. Kozuka, Y. Takahashi, G. Zhao, and T. Yoko, Thin Solid Films
358, 172 (2000).
L. Znaidi, R. Seraphimova, J. F. Bocquet, C. Colbeau-Justin, and C.
Pommier, Mater. Res. Bull. 36, 811 (2001).
C. J. Brinker and G. W. Scherer, Sol-Gel Science. Academic, San
Diego, 1990.
L. C. Klein, Ed., Sol-Gel Optics: Processing and Applications.
Kluwer, Boston, 1994.
K. Vinodgopal, S. Hotchandani, and P. V. Kamat, J. Phys. Chem. 97,
9040 (1993).
M.-P. Zheng, M.-Y. Gu, Y.-P. Jin, and G.-L. Jin, J. Mater. Sci. Lett.
20, 485 (2001).
J. Livage, M. Henry, and C. Sanchez, Prog. Solid State. Chem. 18, 259
(1988).
J. Moser, S. Punchihewa, P. P. Infelta, and M. Grtzel, Langmuir 7,
3012 (1991).
H. Kominami, J. Kato, S. Murakami, Y. Kera, M. Inoue, T. Inui, and
B. Ohtani, J. Mol. Catal. A 144, 165 (1999).
S. Ito, S. Inoue, H. Kawada, M. Hara, M. Iwasaki, and H. Tada,
J. Colloid Interface Sci. 216, 59 (1999).
T. Sugimoto, H. Itoh, and T. Mochida, J. Colloid Interf. Sci. 205, 42
(1998).
T. Sugimoto, X. Zhou, and A. Muramatsu, J. Colloid Interf. Sci. 252,
339 (2002).
T. Sugimoto and X. Zhou, J. Colloid Interf. Sci. 252, 347 (2002).

531
145. N. Uekawa, J. Kajiwara, K. Kakegawa, and Y. Sasaki, J. Colloid Interf.
Sci. 250, 285 (2002).
146. S. Yin, Y. Inoue, S. Uchida, Y. Fujishiro, and T. Sato, J. Mater. Res.
13, 844 (1998).
147. C.-C. Wang and J. Y. Ying, Chem. Mater. 11, 3113 (1999).
148. J. Yu, X. Zhao, and Q. Zhao, Thin Solid Films 379, 7 (2000).
149. J. Yu and X. Zhao, Mater. Res. Bull. 35, 1293 (2000).
150. N. Negishi, K. Takeuchi, T. Ibusuki, and A. K. Datye, J. Mater. Sci.
Lett. 18, 515 (1999).
151. N. Negishi and K. Takeuchi, Thin Solid Films 392, 249 (2001).
152. P. Sawunyama, A. Yasumori, and K. Okada, Mater. Res. Bull. 33, 795
(1998).
153. C. K. Chan, J. F. Porter, Y.-G. Li, W. Guo, and C.-M. Chan, J. Am.
Ceram. Soc. 82, 566 (1999).
154. A. V. Vorontsov, A. A. Altynnikov, E. N. Savinov, and E. N. Kurkin,
J. Photochem. Photobiol., A: Chem. 144, 193 (2001).
155. S. Yin, R. Li, Q. He, and T. Sato, Mater. Chem. Phys. 75, 76 (2002).
156. X.-S. Feng, S.-Z. Kang, H.-G. Liu, and J. Mu, Thin Solid Films 352,
223 (1999).
157. K. Shimizu, H. Imai, H. Hirashima, and K. Tsukuma, Thin Solid
Films 351, 220 (1999).
158. S. J. Kim, S. D. Park, C. K. Rhee, W. W. Kim, and S. Park, Scr. Mater.
44, 1229 (2001).
159. L. Shi, C. Li, H. Gu, and D. Fang, Mater. Chem. Phys. 62, 62 (2000).
160. R. W. Siegel, in Materials Science and Technology (R. W. Cahn,
Ed.), Vol. 15, p. 583. VCH, Weinheim, 1991.
161. K. Takagi, T. Makimoto, H. Hiraiwa, and T. Negishi, J. Vac. Sci.
Technol. A 19, 2931 (1999).
162. P. Zeman and S. Takabayashi, Surf. Coat. Technol. 153, 93 (2002).
163. P. Zeman and S. Takabayashi, J. Vac. Sci. Technol. A 20, 388 (2002).
164. W. Zhang, Y. Li, and F. Wang, J. Mater. Sci. Technol. (China) 18, 101
(2002).
165. Y. Hou, D. Zhuang, D. Zhao, J. Zhang, and C. Wang, in Advanced
Photonic Sensors: Technology and Applications. Proc. SPIE, Vol.
4220, 2000, p. 34.
166. O. Banakh, P. E. Schmid, R. Sanjines, and F. Levy, Surf. Coat. Technol.
151152, 272 (2002).
167. T. Wang, H. Wang, P. Xu, X. Zhao, Y. Liu, and S. Chao, Thin Solid
Films 334, 103 (1998).
168. C.-C. Ting, S.-Y. Chen, and D.-M. Liu, J. Appl. Phys. 88, 4628 (2000).
169. S. Takeda, S. Suzuki, H. Odaka, and H. Hosono, Thin Solid Films
392, 338 (2001).
170. H. Ohsaki, Y. Tachibana, A. Mitsui, T. Kamiyama, and Y. Hayashi,
Thin Solid Films 392, 169 (2001).
171. S. G. Springer, P. E. Schmid, R. Sanjines, and E. Levy, Surf. Coat.
Techn. 151152, 51 (2002).
172. J. A. Ayllon, A. Figueras, S. Garelik, L. Spirkova, J. Durand, and L.
Cot, J. Mater. Sci. Lett. 18, 1319 (1999).
173. V. G. Bessergenev, I. V. Khmelinskii, R. J. F. Pereira, V. V. Krisuk,
A. E. Turgambaeva, and I. K. Igumenov, Vacuum 64, 275 (2002).
174. Y. C. Zhu and C. X. Ding, J. Eur. Ceram. Soc. 20, 127 (2000).
175. Y. C. Zhu and C. X. Ding, Nanostruct. Mater. 11, 319 (1999).
176. A. Goossens and J. Schoonman, Eur. J. Sol. State Inorgan. Chem. 32,
779 (1995).
177. B. Major, R. Ebner, P. Zieba, and W. Wolzynski, Appl. Phys. A 69,
921 (1999).
178. T. Deguchi, K. Imai, H. Matsui, M. Iwasaki, H. Tada, and S. Ito, J.
Mater. Sci. 36, 4723 (2001).
179. H. Wang, T. Wang, and P. Xu, J. Mater. Sci., Mater. Electron. 9, 327
(1998).
180. M. Terashima, N. Inoue, S. Kashiwabara, and R. Fujimoto, Appl.
Surf. Sci. 169170, 535 (2001).
181. S. K. Zheng, T. M. Wang, C. Wang, and G. Xiang, Nucl. Instrum.
Methods B 187, 479 (2002).
182. S. K. Zheng, T. M. Wang, W. C. Hao, and R. Shen, Vacuum 65, 155
(2002).

532
183. X. Liu, C. Liang, H. Wang, X. Yang, L. Lu, and X. Wang, Mater. Sci.
Eng. A 326, 235 (2002).
184. H. Yamashita, Y. Ichihashi, M. Takeuchi, S. Kishiguchi, and
M. Anpo, J. Synchrotron Radiat. 6, 451 (1999).
185. K. Oyoshi, N. Sumi, I. Umezu, R. Souda, A. Yamazaki, H. Haneda,
and T. Mitsuhashi, Nucl. Instrum. Methods B 168, 221 (2000).
186. Z. Xu, J. Shang, C. Liu, C. Kang, H. Guo, and Y. Du, Mater. Sci. Eng.
B 63, 211 (1999).
187. T. Ihara, M. Miyoshi, M. Ando, S. Sugihara, and Y. Iriyama, J. Mater.
Sci. 36, 4201 (2001).
188. K. Furusawa, K. Takahashi, S.-H. Cho, H. Kumagai, K. Midorikawa,
and M. Obara, J. Appl. Phys. 87, 1604 (2000).
189. V. Subramanian, E. Wolf, and P. V. Kamat, J. Phys. Chem. B 105,
11439 (2001).
190. T. Ohno, K. Fujihara, S. Saito, and M. Matsumura, Sol. Energy Mater.
Sol. Cells 45, 169 (1997).
191. M. Tsujimoto, S. Moriguchi, S. Isoda, T. Kobayashi, and T. Komatsu,
J. Electron Microsc. 48, 361 (1999).
192. J. Yu, J. C. Yu, B. Cheng, and X. Zhao, J. Sol-Gel Sci. Technol. 24,
39 (2002).
193. H. Wang, P. Xu, and T. Wang, J. Vac. Sci. Technol. B 19, 645 (2001).
194. Y. Wang, H. Cheng, Y. Hao, J. Ma, W. Li, and S. Cai, Thin Solid
Films 349, 120 (1999).
195. H. Deng, Z. Lu, Y. Shen, H. Mao, and H. Xu, Chem. Phys. 231, 95
(1998).
196. A. Dawson and P. V. Kamat, J. Phys. Chem. B 105, 960 (2001).
197. D. Beydoun, R. Amal, G. K. C. Low, and S. McEvoy, J. Phys. Chem.
B 104, 4387 (2000).
198. T. Kasuga, M. Hiramatsu, M. Hirano, A. Hoson, and K. Oyamada,
J. Mater. Res. 12, 607 (1997).
199. S.-H. Hahn, D.-J. Kim, S.-H. Oh, E.-J. Kim, and S.-W. Kim, in Proc.
5th Korea-Russia International Symp. on Science and Technology,
p. 337. IEEE, Piscataway, 2001.
200. F. P. Getton, V. A. Self, J. M. Ferguson, J. G. Leadley, P. A. Sermon,
and M. Montes, Ceram. Trans. 81, 355 (1998).
201. R. van Grieken, J. Aguado, M. J. Lpez-Muoz, and J. Marugn, J.
Photocem. Photobiol. A: Chem. 148, 315 (2002).
202. T. Tanaka, K. Teramura, T. Yamamoto, S. Takenaka, S. Yoshida, and
T. Funabiki, J. Photocem. Photobiol. A: Chem. 148, 277 (2002).
203. B. Pal, M. Sharon, and G. Nogami, Mater. Chem. Phys. 59, 254 (1999).
204. J. Yang, D. Li, X. Wang, X. Yang, and L. Lu, J. Solid State Chem. 165,
193 (2002).
205. M. E. Rincon, A. Jimenez, A. Orihuela, and G. Martinez, Sol. Energy
Mater. Sol. Cells 70, 163 (2001).
206. K. Kato and K.-I. Niihara, Thin Solid Films 298, 76 (1997).
207. A. P. Hong, D. W. Bahnemann, and M. R. Hoffmann, J. Phys. Chem.
91, 6245 (1987).
208. K. B. Dhanalakshmi, S. Latha, S. Anandan, and P. Maruthamuthu,
Intern. J. Hydrogen Energy 26, 669 (2001).
209. S. Wenfeng and A. Yoshida, Sol. Energy Mater. Sol. Cells 69, 189
(2001).
210. H. N. Ghosh and S. Adhikari, Langmuir 17, 4129 (2001).
211. Y. Lei, L. D. Zhang, G. W. Meng, G. H. Li, X. Y. Zhang, C. H. Liang,
W. Chen, and S. X. Wang, Appl. Phys. Lett. 78, 1125 (2001).
212. T. Akita, K. Tanaka, K. Okuma, T. Koyanagi, and M. Haruta, J. Electr.
Microsc. 50, 473 (2001).
213. G. L. Li, G. H. Wang, and J. M. Hong, J. Mater. Res. 14, 3346 (1999).
214. T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, and K. Niihara, Langmuir 14, 3160 (1998).
215. H. Imai, Y. Takei, K. Shimizu, M. Matsuda, and H. Hirashima,
J. Mater. Chem. 9, 2971 (1999).
216. T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, and K. Niihara, Adv.
Mater. 11, 1307 (1999).
217. G. H. Du, Q. Chen, R. C. Che, Z. Y. Yuan, and L. M. Peng, Appl.
Phys. Lett. 79, 3702 (2001).
218. A. Chemseddine and T. Moritz, Eur. J. Inorg. Chem. 235 (1999).
219. J. C. Yu, J. Yu, W. Ho, and L. Zhang, Chem. Comm. 1942 (2001).

Nanocrystalline TiO2 for Photocatalysis


220. H. Parala, A. Devi, R. Bhakta, and R. A. Fischer, J. Mater. Chem. 12,
1625 (2002).
221. B. R. Weinberger and R. B. Garber, Appl. Phys. Lett. 66, 2409 (1995).
222. M. M. Gmez, J. Lu, E. Olsson, A. Hagfeldt, and C. G. Granqvist,
Sol. Energy Mater. Sol. Cells 64, 385 (2000).
223. M. Suzuki, T. Ito, and Y. Taga, Appl. Phys. Lett. 78, 3968 (2001).
224. P. Hoyer, Adv. Mater. 8, 857 (1996).
225. P. Hoyer, Langmuir 12, 1411 (1996).
226. B. B. Lakshimi, C. J. Patrissi, and C. R. Martin, Chem. Mater. 9, 2544
(1997).
227. Y. J. Song and Z. L. Wang, Adv. Mater. 11, 469 (1999).
228. M. Adachi, Y. Murata, M. Harada, and S. Yoshikawa, Chem. Lett.
942 (2000).
229. M. Zhang, Y. Bando, and K. Wada J. Mater. Sci. Lett. 20, 167 (2001).
230. A. Michailowski, D. AlMawlawi, G. Cheng, and M. Moskovits,
Chem. Phys. Lett. 349, 1 (2001).
231. Y. C. Zhu and C. X. Ding, Nanostruct. Mater. 11, 427 (1999).
232. J. Augustynski, J. Electrochim. Acta 38, 43 (1993).
233. J. K. Burdett, Inorgan. Chem. 24, 2244 (1985).
234. J. K. Burdett, T. Hughbands, J. M. Gordon, J. W. Richardson, and J.
V. Smith, J. Am. Chem. Soc. 109, 3639 (1987).
235. A. Fahmi, C. Minot, B. Silvi, and M. Causa, Phys. Rev. B 47, 11717
(1993).
236. X.-F. Yu, N.-Z. Wu, Y.-C. Xie, and Y.-Q. Tang, J. Mater. Sci. Lett. 20,
319 (2001).
237. J. Nair, P. Nair, F. Mizukami, Y. Oosawa, and T. Okubo, Mater. Res.
Bull. 34, 1275 (1999).
238. D. G. Rickerby, Philos. Mag. B 76, 573 (1997).
239. B. Huber, H. Gnaser, and C. Ziegler, Anal. Bioanal. Chem. (2003).
240. P. A. Cox, F. W. H. Dean, and A. A. Williams, Vacuum 33, 839 (1983).
241. C. R. Henry, Surf. Sci. Rep. 31, 231 (1998).
242. H.-J. Freund, Faraday Discuss. 114, 1 (1999).
243. U. Diebold, in The Chemical Physics of Solid Surfaces, Oxide Surfaces, Vol. 9. (D. P. Woodruff, Ed.), Elsevier, Amsterdam, 2001.
244. G. Lu, A. Linsebigler, and J. T. Yates, J. Phys. Chem. 98, 11733 (1994).
245. S. Fischer, A. W. Munz, K.-D. Schierbaum, and W. Gpel, Surf. Sci.
337, 17 (1995).
246. D. A. Bonnell, Prog. Surf. Sci. 57, 187 (1998).
247. U. Diebold, J. Lehman, T. Mahmoud, M. Kuhn, G. Leonardelli,
W. Hebenstreit, M. Schmid, and P. Varga, Surf. Sci. 411, 137 (1998).
248. M. Li, W. Hebenstreit, U. Diebold, M. A. Henderson, and D. R.
Jennison, Faraday Discuss. 114, 245 (1999).
249. R. A. Bennett, P. Stone, and M. Bowker, Faraday Discuss. 114, 267
(1999).
250. Y. Iwasawa, H. Onishi, K. Fukui, S. Suzuki, and T. Sasaki, Faraday
Discuss. 114, 259 (1999).
251. T. Fujino, M. Katayama, K. Inudzuka, T. Okuno, and K. Oura, Appl.
Phys. Lett. 79, 2716 (2001).
252. I. M. Brookes, C. A. Muryn, and G. Thornton, Phys. Rev. Lett. 87,
266103-1 (2001).
253. R. Schaub, P. Thostrup, N. Lopez, E. Lgsgaard, I. Stensgarrd, J. K.
Nrskov, and F. Besenbacher, Phys. Rev. Lett. 87, 266104-1 (2001).
254. G. S. Herman, M. R. Sievers, and Y. Gao, Phys. Rev. Lett. 84, 3354
(2000).
255. R. Hengerer, B. Bolliger, M. Erbudak, and M. Grtzel, Surf. Sci. 460,
162 (2000).
256. Y. Liang, S. Gan, and S. A. Chambers, Phys. Rev. B 63, 235402-1
(2001).
257. M. Lazzeri and A. Selloni, Phys. Rev. Lett. 87, 266105-1 (2001).
258. A. Vittadini, A. Selloni, F. P. Rotzinger, and M. Grtzel, Phys. Rev.
Lett. 81, 2954 (1998).
259. R. E. Tanner, Y. Liang, and E. I. Altman, Surf. Sci. 506, 251 (2002).
260. A. El-Azab, S. Gan, and Y. Liang, Surf. Sci. 506, 93 (2002).
261. X. Peng, J. Wickham, and A. P. Alivisatos, J. Am. Chem. Soc. 120,
5343 (1998).
262. K. S. Hamad, R. Roth, J. Rockenberger, T. van Buuren, and A. P.
Alivisatos, Phys. Rev. Lett. 83, 3474 (1999).

533

Nanocrystalline TiO2 for Photocatalysis


263. N. A. Hill and K. B. Whaley, J. Chem. Phys. 100, 2831 (1994).
264. Z. Y. Wu, J. Zhang, K. Ibrahim, D. C. Xian, G. Li, Y. Tao, T. D. Hu,
S. Bellucci, A. Marcelli, Q. H. Zhang, L. Gao, and Z. Z. Chen, Appl.
Phys. Lett. 80, 2973 (2002).
265. L. X. Chen, T. Rajh, Z. Y. Wang, and M. Thurnauer, J. Phys. Chem.
B 101, 10688 (1997).
266. A. P. Xagas, E. Androulaki, A. Hiskia, and P. Falaras, Thin Solid
Films 357, 173 (1999).
267. A. Provata, P. Falaras, and A. Xagas, Chem. Phys. Lett. 297, 484
(1998).
268. J. Y. Ying, L. F. Chi, H. Fuchs, and H. Gleiter, Nanostruct. Mater. 3,
273 (1993).
269. E. Comini, G. Sberveglieri, M. Ferroni, V. Guidi, C. Frigeri, and D.
Boscarino, J. Mater. Res. 16, 1559 (2001).
270. R. I. Bickley, T. Gonzalez-Carreno, J. S. Lees, L. Palmisano, and R.
J. D. Tilley, J. Sol. State Chem. 92, 178 (1991).
271. P. E. de Jongh and D. Vanmaekelbergh, Phys. Rev. Lett. 77, 3427
(1996).
272. A. Zaban, A. Meier, and B. A. Gregg, J. Phys. Chem. B 101, 7985
(1997).
273. A. Solbrand, H. Lindstrm, H. Rensmo, A. Hagfeldt, and
S. Lindquist, J. Phys. Chem. B 101, 2514 (1997).
274. J. Nelson, Phys. Rev. B 59, 15 374 (1999).
275. R. Knenkamp, R. Henninger, and P. Hoyer, J. Phys. Chem. 97, 7328
(1993).
276. R. Knenkamp and R. Henninger, Appl. Phys. A 58, 87 (1994).
277. R. Knenkamp, Phys. Rev. B 61, 11057 (2000).
278. F. Cao, G. Oskam, G. J. Meyer, and P. C. Searson, J. Phys. Chem. 100,
17 021 (1996).
279. P. M. Sommeling, H. C. Rieffe, J. M. Kroon, J. A. M. van Roosmalen,
A. Schonecker, and W. C. Sinke, in Proceedings of the 14th EC Photovoltaic Solar Energy Conference (H. Ossenbrink, P. Helm, and
H. Ehmann, Eds.), pp. 18161819. H. S. Stephens and Associates,
Bedford, 1997.
280. A. Solbrand, H. Lindstrom, H. Rensmo, A. Hagfeldt, S.-E. Lindquist,
and S. Sodergren, J. Phys. Chem. B 101, 2514 (1997).
281. K. Schwarzburg and F. Willig, Appl. Phys. Lett. 58, 2520 (1991).
282. B. Levy, W. Liu, and S. E. Gilbert, J. Phys. Chem. B 101, 1810 (1997).
283. L. Dloczik, O. Ileperuma, I. Lauermann, L. M. Peter, E. A. Ponomarev, G. Redmond, N. J. Shaw, and I. Uhlendorf, J. Phys. Chem. B
101, 10 281 (1997).
284. A. Hagfeldt, S.-E. Lindquist, and M. Grtzel, Sol. Energy Mater. Sol.
Cells 32, 245 (1994).
285. S. A. Haque, Y. Tachibana, D. R. Klug, and J. R. Durrant, J. Phys.
Chem. B 102, 1745 (1998).
286. C. J. Barbe, F. Arendse, P. Compte, M. Jirousek, F. Lenzmann,
V. Shklover, and M. Grtzel, J. Am. Ceram. Soc. 80, 3157 (1997).
287. F. Cao, G. Oskam, P. C. Searson, J. M. Stipkala, T. A. Heimer,
F. Farzad, and G. J. Meyer, J. Phys. Chem. 99, 11 974 (1995).
288. A. Goossens, B. van der Zanden, and J. Schooman, Chem. Phys. Lett.
331, 1 (2000).
289. M. Takahashi, K. Tsukigi, T. Uchino, and T. Yoko, Thin Solid Films
388, 231 (2001).
290. Th. Dittrich, E. A. Lebedev, and J. Weidmann, Phys. Stat. Sol. (A)
165, R5 (1998).
291. J. Weidmann, Th. Dittrich, E. Konstantinova, I. Lauermann,
I. Uhlendorf, and F. Koch, Sol. Energy Mater. Sol. Cells 56, 153 (1999).
292. T. Dittrich, J. Weidmann, F. Koch, I. Uhlendorf, and I. Lauermann,
Appl. Phys. Lett. 75, 3980 (1999).
293. V. Duzhko, V. Yu. Timoshenko, F. Koch, and Th. Dittrich, Phys. Rev.
B 64, 075204 (2001).
294. V. Kytin and Th. Dittrich, Phys. Stat. Sol. (A) 185, 461 (2001).
295. V. Kytin, V. Duzhko, V. Yu. Timoshenko, J. Rappich, and Th. Dittrich,
Phys. Stat. Sol. (A) 185, R1 (2001).
296. V. Kytin, Th. Dittrich, F. Koch, and E. Lebedev, Appl. Phys. Lett. 79,
108 (2001).
297. P. Knauth and H. L. Tuller, J. Appl. Phys. 85, 897 (1999).

298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
308.
309.
310.
311.
312.
313.
314.
315.
316.
317.
318.
319.
320.
321.
322.
323.
324.
325.
326.
327.
328.
329.
330.
331.
332.
333.
334.
335.
336.

J. Maier, Prog. Solid State Chem. 23, 171 (1995).


J. A. S. Ikeda and Y.-M. Chiang, J. Am. Ceram. Soc. 76, 2437 (1993).
J. F. Baumard and E. Tani, J. Chem. Phys. 67, 857 (1977).
D. M. Smyth, Prog. Solid State Chem. 15, 145 (1984).
K. Hoshino, N. L. Peterson, and C. L. Wiley, J. Phys. Chem. Solids
46, 1397 (1985).
F. Millot, M.-G. Blanchin, R. Tetot, J.-F. Marucco, B. Poumellec, C.
Picard, and B. Touzelin, Prog. Solid State Chem. 17, 263 (1987).
A. Bernasik, M. Rekas, M. Sloma, and W. Weppner, Solid State Ionics
72, 12 (1994).
D. Mardare, M. Tasca, M. Delibas, and G. I. Rusu, Appl. Surf. Sci.
156, 200 (2000).
D. Mardare, C. Baban, R. Gavrila, M. Modreanu, and G. I. Rusu,
Surf. Sci. 507510, 468 (2002).
M. Gmez, J. Rodrguez, S. Tingry, A. Hagfeldt, S.-E. Lindquist, and
C. G. Granqvist, Sol. Energy Mat. Sol. Cells 59, 277 (1999).
H. Tang, K. Prassad, R. Sanjins, P. E. Schmid, and F. Lvy, J. Appl.
Phys. 75, 2042 (1994).
R. Sanjins, H. Tang, H. Berger, F. Gozzo, G. Margaritondo, and F.
Lvy, J. Appl. Phys. 75, 2945 (1994).
H. Wittmer, S. Holten, H. Kliem, and H. D. Breuer, Phys. Stat. Sol.
(A) 181, 461 (2000).
V. I. Makarov, S. A. Kochubei, and I. Khmelinskii, Chem. Phys. Lett.
355, 504 (2002).
K. Wilke and H. D. Breuer, Z. Phys. Chem. 213, 135 (1999).
M. Anpo, Sol. Energy Mater. Sol. Cells 38, 221 (1995).
M. A. Fox, Sol. Energy Mater. Sol. Cells 38, 381 (1995).
V. Augugliaro, V. Loddo, L. Palmisano, and M. Schiavello, Sol. Energy
Mater. Sol. Cells 38, 411 (1995).
D. Bahnemann, D. Bockelmann, and R. Goslich, Sol. Energy Mater.
24, 564 (1991).
D. M. Blake, J. Webb, C. Turchi, and K. Magrini, Sol. Energy Mater.
24, 584 (1991).
J. Yu, X. Zhao, and Q. Zhao, J. Mater. Sci. Lett. 19, 1015 (2000).
S. Ruan, F. Wu, T. Zhang, W. Gao, B. Xu, and M. Zhao, Mater. Chem.
Phys. 69, 7 (2001).
K. Baba and R. Hatada, Surf. Coat. Technol. 136, 241 (2001).
T. Sumita, T. Yamaki, S. Yamamoto, and A. Miyashita, Jpn. J. Appl.
Phys., Part 1 40, 4007 (2001).
L. Gao and Q. Zhang, Scr. Mater. 44, 1195 (2001).
A. J. Maira, K. L. Yeung, J. Soria, J. M. Coronado, C. Belver, C. Y.
Lee, and V. Augugliaro, Appl. Catal. B: Environ. 29, 327 (2001).
K. Y. Jung, S. B. Park, and S.-K. Ihm, Appl. Catal. A: Gen. 224, 229
(2002).
M. S. Jeon, T. K. Lee, D. H. Kim, H. Joo, and H. T. Kim, Sol. Energy
Mater. Sol. Cells 57, 217 (1999).
G.-J. Chee, Y. Nomura, K. Ikebukuro, and I. Karube, Sens. Actuators
B 80, 15 (2001).
Y. Xu, L. Jiang, X. Lu, J. Wang, and D. Zhou, High Technol. Letters
(China) 6, 63 (2000).
J. Sheng, L. Shivalingappa, J. Karasawa, and T. Fukami, J. Mater. Sci.
34, 6201 (1999).
B. Kim, D. Byun, J. K. Lee, and D. Park, Jpn. J. Appl. Phys., Part 1
41, 222 (2002).
H. Tada and M. Tanaka, Langmuir 13, 360 (1997).
C.-G. Wu, L.-F. Tzeng, Y.-T. Kuo, and C. H. Shu, Appl. Catal. A: Gen.
226, 199 (2002).
A. Rachel, B. Lavedrine, M. Subrahmanyam, and P. Boule, Catal.
Comm. 3, 165 (2002).
R. L. Pozzo, J. L. Giombi, M. A. Baltans, and A. E. Cassano, Appl.
Catal. B: Environ. 38, 61 (2002).
Y. Paz, Z. Luo, L. Rabenberg, and A. Heller, J. Mater. Res. 10, 2842
(1995).
M. Nakamura, L. Sirghi, T. Aoki, and Y. Hatanaka, Surf. Sci. 507
510, 778 (2002).
T. Watanabe, A. Nakajima, R. Wang, M. Minabe, S. Koizumi,
A. Fujishima, and K. Hashimoto, Thin Solid Films 351, 260 (1999).

534
337. M. Miyauchi, N. Kieda, S. Hishita, T. Mitsuhashi, A. Nakajima,
T. Watanabe, and K. Hashimoto, Surf. Sci. 511, 401 (2002).
338. V. Romeas, P. Pichat, C. Guillard, T. Chopin, and C. Lehaut, J. Phys.
IV, Proc. 9, PR3/247 (1999).
339. K. Miyashita, S. Kuroda, T. Ubukata, T. Ozawa, and H. Kubota,
J. Mater. Sci. 36, 3877 (2001).
340. K. Miyashita, S. Kuroda, T. Sumita, and T. Ubukata, J. Mater. Sci.
Lett. 20, 2137 (2001).
341. J. C. Yu, J. Yu, W. Ho, and J. Zhao, J. Photochem. Photobiol. A:
Chem. 148, 331 (2002).
342. J. C. Yu, J. Yu, and J. Zhao, Appl. Catal. B: Environ. 36, 31 (2002).
343. Y. Ohko, S. Saitoh, T. Tatsuma, and A. Fujishima, J. Electrochem.
Soc. 148, B24 (2001).
344. D. W. Bahnemann, M. Hilgendorff, and R. Memming, J. Phys. Chem.
B 101, 4265 (1997).
345. D. W. Goodman, Surf. Rev. Lett. 2, 9 (1995).
346. C. T. Campbell, Surf. Sci. Rep. 27, 1 (1997).
347. M. Bumer and H.-J. Freund, Prog. Surf. Sci. 61, 127 (1999).
348. H.-J. Freund, Surf. Sci. 500, 271 (2002).
349. S. Sato and J. M. White, Chem. Phys. Lett. 72, 83 (1980).
350. D. Hufschmidt, D. Bahnemann, J. J. Testa, C. A. Emilio, and M. I.
Litter, J. Photochem. Photobiol. A: Chem. 148, 223 (2002).
351. U. Siemon, D. Bahnemann, J. J. Testa, D. Rodrguez, M. I. Litter,
N. Bruno, J. Photochem. Photobiol. A: Chem. 148, 247 (2002).
352. F. B. Li and X. Z. Li, Appl. Catal. A: Gen. 228, 15 (2002).
353. M. Dat, Y. Ichihashi, T. Yamashita, A. Chiorino, F. Boccuzzi, and
M. Haruta, Catal. Today 72, 89 (2002).
354. V. Vamathevan, R. Amal, D. Beydoun, G. Low, S. McEvoy, J. Photochem. Photobiol. A: Chem. 148, 233 (2002).
355. D. Dvoranov, V. Brezov, M. Mazra, and M. A. Malati, Appl.
Catal. B: Environ. 37, 91 (2002).
356. H. Yamashita, M. Harada, J. Misaka, M. Takeuchi, K. Ikeue, and M.
Anpo, J. Photochem. Photobiol. A: Chem. 148, 257 (2002).
357. A. Di Paola, E. Garca-Lpez, S. Ikeda, G. Marc, B. Ohtani, and L.
Palmisano, Catal. Today 75, 87 (2002).
358. A.-W. Xu, Y. Gao, and H.-Q. Liu, J. Catal. 207, 151 (2002).
359. R. Vogel, K. Pohl, and H. Weller, Chem. Phys. Lett. 174, 241 (1999).
360. S. Kohtani, A. Kudo, and T. Ssakata, Chem. Phys. Lett. 206, 166
(1993).
361. R. Vogel, P. Hoyer, and H. Weller, J. Phys. Chem. 98, 3183 (1994).
362. D. Liu and P. V. Kamat, J. Phys. Chem. 97, 10769 (1993).
363. K. Vinodgopal and P. V. Kamat, Environ. Sci. Technol. 29, 841 (1995).
364. I. Bedja and P. V. Kamat, J. Phys. Chem. 99, 9182 (1995).
365. R. Q. Long and R. T. Yang, J. Catal. 207, 158 (2002).
366. G. Coln, M. C. Hidalgo, and J. A. Navio, Appl. Catal. A: Gen. 231,
185 (2002).
367. M. K. Nazeeruddin, A. Kay, I. Rodicio, B. R. Humphry, E. Mueller,
P. Liska, N. Vlachopoulos, and M. Grtzel, J. Am. Chem. Soc. 115,
6382 (1993).
368. F. Willig, R. Eichberger, N. S. Sundaresan, and B. A. Parkinson, J.
Am. Chem. Soc. 112, 2702 (1990).
369. P. V. Kamat, S. Das, K. G. Thomas, and M. V. George, Chem. Phys.
Lett. 178, 75 (1991).
370. R. Eichberger and F. Willig, Chem. Phys. 141, 159 (1990).
371. B. Burfeindt, T. Hannappel, W. Storck, and F. Willig, J. Phys. Chem.
100, 16463 (1996).
372. N. J. Cherepy, G. P. Smestad, M. Grtzel, and J. Z. Zhang, J. Phys.
Chem. B 101, 9342 (1997).
373. T. Hannappel, B. Burfeindt, W. Storck, and F. Willig, J. Phys. Chem.
B 101, 6799 (1997).
374. J. Randy, R. J. Ellingson, J. B. Asbury, S. Ferrere, H. N. Ghosh, J. R.
Sprague, T. Lian, and A. J. Nozik, J. Phys. Chem. B 102, 6455 (1998).
375. A. Furube, T. Asahi, H. Masuhara, H. Yamashita, and M. Anpo, J.
Phys. Chem. B 103, 3120 (1999).
376. H. Al-Ekabi, B. Butters, D. Delany, J. Ireland, N. Lewis, T. Powell,
and J. Story, Trace Met. Environ. 3, 321 (1993).
377. V. Iliev and D. Tomova, Catal. Comm. 3, 287 (2002).

Nanocrystalline TiO2 for Photocatalysis


378. Q. Dai and J. Rabani, J. Photochem. Photobiol. A: Chem. 148, 17
(2002).
379. C. G. Garcia, C. J. Kleverlaan, C. A. Bignozzi, and N. Y. M. Iha, J.
Photochem. Photobiol. A: Chem. 147, 143 (2002).
380. J. E. Moser, P. Bonhoete, and M. Grtzel, Coord. Chem. Rev. 171,
245 (1998).
381. G. Centi, P. Ciambelli, S. Perathoner, and P. Russo, Catal. Today 75,
3 (2002).
382. R. Abe, K. Sayama, K. Domen, and H. Arakawa, Chem. Phys. Lett.
344, 339 (2001).
383. M. Ashokkumar, Int. J. Hydrog. Energy 23, 427 (1998).
384. M. Anpo, H. Yamashita, Y. Ichihashi, Y. Fujii, and M. Honda, J. Phys.
Chem. B 101, 2632 (1997).
385. M. Anpo, Nuovo Cimento 19D, 1641 (1997).
386. P. G. Smirniotis, D. A. Pea, and B. S. Uphade, Angew. Chem. 113,
2537 (2001).
387. S. Daito, F. Tochikubo, and T. Watanabe, Jpn. J. Appl. Phys., Part 1
40, 2475 (2001).
388. S. B. Kim and S. C. Hong, Appl. Catal. B: Environ. 35, 305 (2002).
389. P. B. Amama, K. Itoh, and M. Murabayashi, Appl. Catal. B: Environ.
37, 321 (2002).
390. A. Bouzaza and A. Laplanche, J. Photochem. Photobiol. A: Chem.
150, 207 (2002).
391. B. Dabrowski, A. Zaleska, M. Janczarek, J. Hupka, and J. D. Miller,
J. Photochem. Photobiol. A: Chem. 151, 201 (2002).
392. J. Araa, O. Gonzlez Daz, M. Miranda Saracho, J. M. Doa
Rodrguez, J. A. Herrera Melin, and J. Prez Pea, Appl. Catal. B:
Environ. 36, 113 (2002).
393. D. W. Bahnemann, S. N. Kholuiskaya, R. Dillert, A. I. Kulak, and A.
I. Kokorin, Appl. Catal. B: Environ. 36, 161 (2002).
394. A. Rachel, M. Sarakha, M. Subrahmanyam, and P. Boule, Appl. Catal.
B: Environ. 37, 293 (2002).
395. A. Rachel, M. Subrahmanyam, and P. Boule, Appl. Catal. B: Environ.
37, 301 (2002).
396. J. A. Byrne, A. Davidson, P. S. M. Dunlop, and B. R. Eggins, J. Photochem. Photobiol. A: Chem. 148, 365 (2002).
397. M. Bekbolet, A. S. Suphandag, and C. S. Uyguner, J. Photochem.
Photobiol. A: Chem. 148, 121 (2002).
398. Y. Cho and W. Choi, J. Photochem. Photobiol. A: Chem. 148, 129
(2002).
399. S. Parra, J. Olivero, and C. Pulgarin, Appl. Catal. B: Environ. 36, 75
(2002).
400. M. Muneer and D. Bahnemann, Appl. Catal. B: Environ. 36, 95
(2002).
401. S. Parra, S. Malato, and C. Pulgarin, Appl. Catal. B: Environ. 36, 131
(2002).
402. E. Vulliet, C. Emmelin, J.-M. Chovelon, C. Guillard, and J.-M. Herrmann, Appl. Catal. B: Environ. 38, 127 (2002).
403. H. Einaga, S. Futamura, and T. Ibusuki, Appl. Catal. B: Environ. 38,
215 (2002).
404. M. C. Blount and J. L. Falconer, Appl. Catal. B: Environ. 39, 39 (2002).
405. K.-I. Shimizu, T. Kaneko, T. Fujishima, T. Kodama, H. Yoshida, and
Y. Kitayama, Appl. Catal. A: Gen. 225, 185 (2002).
406. S. Vijaikumar, N. Somasundaram, C. Srinivasan, Appl. Catal. A: Gen.
225, 129 (2002).
407. M. Kang, Appl. Catal. B: Environ. 37, 187 (2002).
408. B. Ohtani, K. Iwai, H. Kominami, T. Matsuura, Y. Kera, and
S. Nishimoto, Chem. Phys. Lett. 242, 315 (1995).
409. S. Horikoshi, N. Watanabe, H. Onishi, H. Hidaka, and N. Serpone,
Appl. Catal. B: Environ. 37, 117 (2002).
410. N. San, A. Hatipoglu, G. Kotrk, and Z. nar, J. Photochem. Photobiol. A: Chem. 146, 189 (2002).
411. V. Iliev, J. Photochem. Photobiol. A: Chem. 151, 195 (2002).
412. T. Nakashima, Y. Ohko, D. A. Tryk, A. Fujishima, J. Photochem.
Photobiol. A: Chem. 151, 207 (2002).
413. G. Mele, G. Ciccarella, G. Vasapollo, E. Garca-Lpez, L. Palmisano,
and M. Schiavello, Appl. Catal. B: Environ. 38, 309 (2002).

Nanocrystalline TiO2 for Photocatalysis


414. V. Durgakumari, M. Subrahmanyam, K. V. Subba Rao, A. Ratnamala, M. Noorjahan, and K. Tanaka, Appl. Catal. A: Gen. 248, 155
(2002).
415. R. W. Matthews and S. R. McEvoy, Sol. Energy 49, 507 (1992).
416. L. Zhang, T. Kanki, N. Sano, and A. Toyoda, Sol. Energy 70, 331
(2001).
417. J. Grzechulska and A. W. Morawski, Appl. Catal. B: Environ. 36, 45
(2002).
418. J. Li, C. Chen, J. Zhao, H. Zhu, and J. Orthman, Appl. Catal. B:
Environ. 37, 331 (2002).
419. Z. Sun, Y. Chen, Q. Ke, Y. Yang, and J. Yuan, J. Photochem. Photobiol. A: Chem. 149, 169 (2002).
420. M. Skmen and A. zkan, J. Photochem. Photobiol. A: Chem. 147,
77 (2002).
421. S. Sakthivel, M. V. Shankar, M. Palanichamy, B. Arabindoo, and V.
Murugesan, J. Photochem. Photobiol. A: Chem. 148, 153 (2002).
422. S. Al-Qaradawi and S. R. Salman, J. Photochem. Photobiol. A: Chem.
148, 161 (2002).
423. T. Sauer, G. Cesconeto Neto, H. J. Jos, R. F. P. M. Moreira, J. Photochem. Photobiol. A: Chem. 149, 147 (2002).
424. H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard,
and J.-M. Herrmann, Appl. Catal. B: Environ. 39, 75 (2002).
425. H. Zhan and H. Tian, Dyes Pigments 37, 231 (1998).
426. H. Zhan, K. Chen, and H. Tian, Dyes Pigments 37, 241 (1998).
427. H. Zhan, H. Tian, K. Chen, and W. Zhu, Toxicol. Environ. Chem. 69,
531 (1999).
428. S. Malato, J. Blanco, A. Vidal, and C. Richter, Appl. Catal. B: Environ.
37, 1 (2002).
429. T. An, G. Li, Y. Xiong, X. Zhu, H. Xing, and G. Liu, Mater. Phys.
Mech. 4, 101 (2001).
430. L. R. Skubal, N. K. Meshkov, T. Rajh, and M. Thurnauer, J. Photochem. Photobiol. A: Chem. 148, 393 (2002).
431. L. R. Skubal and N. K. Meshkov, J. Photochem. Photobiol. A: Chem.
148, 211 (2002).
432. Y. Ming, C. R. Chenthamarakshan, and K. Rajeshwar, J. Photochem.
Photobiol. A: Chem. 147, 199 (2002).
433. S. G. Schrank, H. J. Jos, and R. F. P. M. Moreira, J. Photochem.
Photobiol. A: Chem. 147, 71 (2002).
434. J. Aguado, R. van Grieken, M. J. Lpez-Munoz, and J. Marugn,
Catal. Today 75, 95 (2002).

535
435. X. Z. Li and Y. G. Zhao, Water Sci. Technol. 39, 249 (1999).
436. J. E. Pacheco, M. R. Prairie, and L. Yellowhorse, Trans. ASME, J.
Sol. Energy Engin. 115, 123 (1993).
437. S. M. Rodriguez, C. Richter, J. B. Galvez, and M. Vincent, Sol. Energy
56, 401 (1996).
438. D. C. Schmelling, K. A. Gray, and P. V. Kamat, Water Res. 31, 1439
(1997).
439. J.-S. Hur and Y. Koh, Biotechnol. Lett. 24, 23 (2002).
440. R. L. Ziolli and W. F. Jardim, J. Photochem. Photobiol. A: Chem. 147,
205 (2002).
441. M. Perez, F. Torrades, J. Peral, C. Lizama, C. Bravo, S. Casas, J. Freer,
and H. D. Mansilla, Appl. Catal. B: Environ. 33, 89 (2001).
442. M. Hamerski, J. Grzechulska, and A. W. Morawski, Sol. Energy 66,
395 (1999).
443. M. Bekbolet, M. Lindner, D. Weichgrebe, and D. W. Bahnemann,
Sol. Energy 56, 455 (1996).
444. C. Wei, W. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L.
Smith, K. Rajeshwar, Environ. Sci. Technol. 28, 934 (1994).
445. M. Bekblet, Water Sci. Technol. 1112, 95 (1997).
446. P. S. M. Dunlop, J. A. Byrne, N. Manga, and B. R. Eggins, J. Photochem. Photobiol. A: Chem. 148, 355 (2002).
447. J. Wist, J. Sanabria, C. Dierolf, W. Torres, and C. Pulgarin, J. Photochem. Photobiol. A: Chem. 147, 241 (2002).
448. I. Liu, L. A. Lawton, B. Cornish, and P. K. J. Robertson, J. Photochem.
Photobiol. A: Chem. 148, 349 (2002).
449. R. Armon, N. Laot, N. Narkis, and I. Neeman, J. Adv. Oxid. Technol.
3, 145 (1998).
450. M. Z. Atashbar, in Proceedings 1st IEEE Conference on Nanotechnology. p. 544. IEEE-NANO 2001, IEEE, Piscataway, NJ,
2001.
451. L. W. Miller, M. I. Tejedor, B. P. Nelson, and M. A. Anderson, J. Phys.
Chem. B 103, 8490 (1999).
452. H. Honda, A. Ishizahi, R. Soma, K. Hashimoto, and A. Fujishima, J.
Illum. Eng. Soc. (USA) 27, 42 (1998).
453. J. Mizuguchi, J. Electrochem. Soc. 148, J55 (2001).
454. T. Okamoto and I. Yamaguchi, J. Microsc. 202, 100 (2001).
455. Z. V. Saponjic, T. Rajh, J. M. Nedeljkovic, and M. C. Thurnauer,
Mater. Sci. Forum 352, 91 (2000).
456. M. Romero, J. Blanco, B. Sanchez, A. Vidal, S. Malato, A. I. Carbona,
and E. Garcia, Sol. Energy 66, 169 (1999).

Você também pode gostar