Figure 19 shows a typical TEM image of TiO2 nanorods prepared from the solutions with the weight ratio of precursor/solvent/surfactant 1:5:3.183 Similar to the hydrothermal method, the
Trang 1Chemical Reviews is published by the American Chemical Society 1155 Sixteenth Street N.W., Washington, DC 20036
Properties, Modifications, and Applications
Xiaobo Chen, and Samuel S Mao
Chem Rev., 2007, 107 (7), 2891-2959• DOI: 10.1021/cr0500535 • Publication Date (Web): 23 June 2007
Downloaded from http://pubs.acs.org on March 2, 2009
More About This Article
Additional resources and features associated with this article are available within the HTML version:
• Supporting Information
• Links to the 71 articles that cite this article, as of the time of this article download
• Access to high resolution figures
• Links to articles and content related to this article
• Copyright permission to reproduce figures and/or text from this article
Trang 22 Synthetic Methods for TiO2 Nanostructures 2892
2.2 Micelle and Inverse Micelle Methods 2895
2.6 Direct Oxidation Method 2902
2.7 Chemical Vapor Deposition 2903
2.8 Physical Vapor Deposition 2904
2.14 TiO2 Opal and Photonic Materials 2907
2.15 Preparation of TiO2 Nanosheets 2908
3 Properties of TiO2Nanomaterials 2909
3.1 Structural Properties of TiO2 Nanomaterials 2909
3.2 Thermodynamic Properties of TiO2
Nanomaterials
29113.3 X-ray Diffraction Properties of TiO2
Nanomaterials
29123.4 Raman Vibration Properties of TiO2
3.5 Electronic Properties of TiO2 Nanomaterials 2913
3.6 Optical Properties of TiO2Nanomaterials 2915
3.7 Photon-Induced Electron and Hole Properties
of TiO2Nanomaterials
2918
4 Modifications of TiO2 Nanomaterials 2920
4.1 Bulk Chemical Modification: Doping 2921
4.1.1 Synthesis of Doped TiO2Nanomaterials 2921
4.1.2 Properties of Doped TiO2Nanomaterials 2921
4.2 Surface Chemical Modifications 2926
Second Generation
29305.1.3 Nonmetal-Doped TiO2 Nanomaterials:
Third Generation
2931
5.2 Photovoltaic Applications 29325.2.1 The TiO2Nanocrystalline Electrode in
DSSCs
29325.2.2 Metal/Semiconductor Junction Schottky
Diode Solar Cell
29385.2.3 Doped TiO2Nanomaterials-Based SolarCell
29385.3 Photocatalytic Water Splitting 29395.3.1 Fundamentals of Photocatalytic Water
Splitting
29395.3.2 Use of Reversible Redox Mediators 29395.3.3 Use of TiO2Nanotubes 29405.3.4 Water Splitting under Visible Light 29415.3.5 Coupled/Composite Water-Splitting
29435.4.3 Counterelectrode for an Electrochromic
Device
29445.4.4 Photoelectrochromic Devices 2945
An exponential growth of research activities has been seen
in nanoscience and nanotechnology in the past decades.13-17New physical and chemical properties emerge when the size
of the material becomes smaller and smaller, and down to
* Corresponding author E-mail: XChen3@lbl.gov.
† E-mail: SSMao@lbl.gov.
10.1021/cr0500535 CCC: $65.00 © 2007 American Chemical Society
Published on Web 06/23/2007
Trang 3the nanometer scale Properties also vary as the shapes of
the shrinking nanomaterials change Many excellent reviews
and reports on the preparation and properties of nanomaterials
have been published recently.6-44Among the unique
proper-ties of nanomaterials, the movement of electrons and holes
in semiconductor nanomaterials is primarily governed by the
well-known quantum confinement, and the transport
proper-ties related to phonons and photons are largely affected by
the size and geometry of the materials.13-16 The specific
surface area and surface-to-volume ratio increase
dramati-cally as the size of a material decreases.13,21The high surface
area brought about by small particle size is beneficial to many
TiO2-based devices, as it facilitates reaction/interaction
between the devices and the interacting media, which mainly
occurs on the surface or at the interface and strongly depends
on the surface area of the material Thus, the performance
of TiO2-based devices is largely influenced by the sizes of
the TiO2building units, apparently at the nanometer scale
As the most promising photocatalyst,7,11,12,33 TiO2
mate-rials are expected to play an important role in helping solve
many serious environmental and pollution challenges TiO2also bears tremendous hope in helping ease the energy crisisthrough effective utilization of solar energy based onphotovoltaic and water-splitting devices.9,31,32As continuedbreakthroughs have been made in the preparation, modifica-tion, and applications of TiO2nanomaterials in recent years,especially after a series of great reviews of the subject inthe 1990s.7,8,10-12,33,45 we believe that a new and compre-hensive review of TiO2nanomaterials would further promoteTiO2-based research and development efforts to tackle theenvironmental and energy challenges we are currently facing.Here, we focus on recent progress in the synthesis, properties,modifications, and applications of TiO2nanomaterials Thesyntheses of TiO2 nanomaterials, including nanoparticles,nanorods, nanowires, and nanotubes are primarily categorizedwith the preparation method The preparations of mesopo-rous/nanoporous TiO2, TiO2 aerogels, opals, and photonicmaterials are summarized separately In reviewing nanoma-terial synthesis, we present a typical procedure and repre-sentative transmission or scanning electron microscopyimages to give a direct impression of how these nanomate-rials are obtained and how they normally appear For detailedinstructions on each synthesis, the readers are referred tothe corresponding literature
The structural, thermal, electronic, and optical properties
of TiO2nanomaterials are reviewed in the second section
As the size, shape, and crystal structure of TiO2rials vary, not only does surface stability change but alsothe transitions between different phases of TiO2 underpressure or heat become size dependent The dependence ofX-ray diffraction patterns and Raman vibrational spectra onthe size of TiO2nanomaterials is also summarized, as theycould help to determine the size to some extent, althoughcorrelation of the spectra with the size of TiO2nanomaterials
nanomate-is not straightforward The review of modifications of TiO2nanomaterials is mainly limited to the research related tothe modifications of the optical properties of TiO2nanoma-terials, since many applications of TiO2nanomaterials areclosely related to their optical properties TiO2nanomaterialsnormally are transparent in the visible light region By doping
or sensitization, it is possible to improve the optical ity and activity of TiO2 nanomaterials in the visible lightregion Environmental (photocatalysis and sensing) andenergy (photovoltaics, water splitting, photo-/electrochromics,and hydrogen storage) applications are reviewed with anemphasis on clean and sustainable energy, since the increas-ing energy demand and environmental pollution create apressing need for clean and sustainable energy solutions Thefundamentals and working principles of the TiO2nanoma-terials-based devices are discussed to facilitate the under-standing and further improvement of current and practicalTiO2nanotechnology
sensitiv-2 Synthetic Methods for TiO2 Nanostructures 2.1 Sol − Gel Method
The sol-gel method is a versatile process used in makingvarious ceramic materials.46-50In a typical sol-gel process,
a colloidal suspension, or a sol, is formed from the hydrolysisand polymerization reactions of the precursors, which areusually inorganic metal salts or metal organic compoundssuch as metal alkoxides Complete polymerization and loss
of solvent leads to the transition from the liquid sol into asolid gel phase Thin films can be produced on a piece of
Dr Xiaobo Chen is a research engineer at The University of California at
Berkeley and a Lawrence Berkeley National Laboratory scientist He
obtained his Ph.D Degree in Chemistry from Case Western Reserve
University His research interests include photocatalysis, photovoltaics,
hydrogen storage, fuel cells, environmental pollution control, and the related
materials and devices development
Dr Samuel S Mao is a career staff scientist at Lawrence Berkeley National
Laboratory and an adjunct faculty at The University of California at
Berkeley He obtained his Ph.D degree in Engineering from The University
of California at Berkeley in 2000 His current research involves the
development of nanostructured materials and devices, as well as ultrafast
laser technologies Dr Mao is the team leader of a high throughput
materials processing program supported by the U.S Department of
Ener-gy
Trang 4gel method from hydrolysis of a titanium precusor.51-78This
process normally proceeds via an acid-catalyzed hydrolysis
step of titanium(IV) alkoxide followed by
condensa-tion.51,63,66,79-91 The development of Ti-O-Ti chains is
favored with low content of water, low hydrolysis rates, and
excess titanium alkoxide in the reaction mixture
Three-dimensional polymeric skeletons with close packing result
from the development of Ti-O-Ti chains The formation
of Ti(OH)4 is favored with high hydrolysis rates for a
medium amount of water The presence of a large quantity
of Ti-OH and insufficient development of three-dimensional
polymeric skeletons lead to loosely packed first-order
particles Polymeric Ti-O-Ti chains are developed in the
presence of a large excess of water Closely packed
first-order particles are yielded via a three-dimensionally
devel-oped gel skeleton.51,63,66,79-91From the study on the growth
kinetics of TiO2 nanoparticles in aqueous solution using
titanium tetraisopropoxide (TTIP) as precursor, it is found
that the rate constant for coarsening increases with
temper-ature due to the tempertemper-ature dependence of the viscosity of
the solution and the equilibrium solubility of TiO2.63
Second-ary particles are formed by epitaxial self-assembly of primSecond-ary
particles at longer times and higher temperatures, and the
number of primary particles per secondary particle increases
with time The average TiO2 nanoparticle radius increases
linearly with time, in agreement with the
Lifshitz-Slyozov-Wagner model for coarsening.63
Highly crystalline anatase TiO2nanoparticles with different
sizes and shapes could be obtained with the polycondensation
of titanium alkoxide in the presence of tetramethylammonium
hydroxide.52,62 In a typical procedure, titanium alkoxide is
added to the base at 2°C in alcoholic solvents in a
three-neck flask and is heated at 50-60°C for 13 days or at
90-100°C for 6 h A secondary treatment involving autoclave
heating at 175 and 200 °C is performed to improve the
crystallinity of the TiO2nanoparticles Representative TEM
images are shown in Figure 1 from the study of Chemseddine
et al.52
A series of thorough studies have been conducted by
Sugimoto et al using the sol-gel method on the formation
of TiO2nanoparticles of different sizes and shapes by tuning
the reaction parameters.67-71Typically, a stock solution of
a 0.50 M Ti source is prepared by mixing TTIP with
triethanolamine (TEOA) ([TTIP]/[TEOA] ) 1:2), followed
by addition of water The stock solution is diluted with a
shape controller solution and then aged at 100°C for 1 day
and at 140°C for 3 days The pH of the solution can be
tuned by adding HClO4or NaOH solution Amines are used
as the shape controllers of the TiO2nanomaterials and act
as surfactants These amines include TEOA,
diethylenetri-amine, ethylenedidiethylenetri-amine, trimethylenedidiethylenetri-amine, and
triethyl-enetetramine The morphology of the TiO nanoparticles
sodium stearate The shape control is attributed to the tuning
of the growth rate of the different crystal planes of TiO2nanoparticles by the specific adsorption of shape controllers
to these planes under different pH conditions.70
A prolonged heating time below 100°C for the as-preparedgel can be used to avoid the agglomeration of the TiO2nano-particles during the crystallization process.58,72 By heatingamorphous TiO2in air, large quantities of single-phase ana-tase TiO2nanoparticles with average particle sizes between
7 and 50 nm can be obtained, as reported by Zhang andBanfield.73-77Much effort has been exerted to achieve highlycrystallized and narrowly dispersed TiO2nanoparticles usingthe sol-gel method with other modifications, such as asemicontinuous reaction method by Znaidi et al.78and a two-stage mixed method and a continuous reaction method byKim et al.53,54
By a combination of the sol-gel method and an anodicalumina membrane (AAM) template, TiO2 nanorods havebeen successfully synthesized by dipping porous AAMsinto a boiled TiO2 sol followed by drying and heatingprocesses.92,93In a typical experiment, a TiO2sol solution isprepared by mixing TTIP dissolved in ethanol with a solutioncontaining water, acetyl acetone, and ethanol An AAM isimmersed into the sol solution for 10 min after being boiled
in ethanol; then it is dried in air and calcined at 400°C for
10 h The AAM template is removed in a 10 wt % H3PO4aqueous solution The calcination temperature can be used
to control the crystal phase of the TiO2 nanorods At lowtemperature, anatase nanorods can be obtained, while athigh temperature rutile nanorods can be obtained The poresize of the AAM template can be used to control the size ofthese TiO2nanorods, which typically range from 100 to 300
nm in diameter and several micrometers in length ently, the size distribution of the final TiO2 nanorods islargely controlled by the size distribution of the pores ofthe AAM template In order to obtain smaller and mono-sized TiO2nanorods, it is necessary to fabricate high-qualityAAM templates Figure 3 shows a typical TEM for TiO2nanorods fabricated with this method Normally, the TiO2nanorods are composed of small TiO2 nanoparticles ornanograins
Appar-By electrophoretic deposition of TiO2colloidal suspensionsinto the pores of an AAM, ordered TiO2 nanowire arrayscan be obtained.94In a typical procedure, TTIP is dissolved
in ethanol at room temperature, and glacial acetic acid mixedwith deionized water and ethanol is added under pH ) 2-3with nitric acid Platinum is used as the anode, and an AAMwith an Au substrate attached to Cu foil is used as thecathode A TiO2sol is deposited into the pores of the AMMunder a voltage of 2-5 V and annealed at 500°C for 24 h.After dissolving the AAM template in a 5 wt % NaOHsolution, isolated TiO nanowires are obtained In order to
Trang 5fabricate TiO2nanowires instead of nanorods, an AAM with
long pores is a must
TiO2 nanotubes can also be obtained using the sol-gel
method by templating with an AAM95-98and other organic
compounds.99,100For example, when an AAM is used as the
template, a thin layer of TiO sol on the wall of the pores of
the AAM is first prepared by sucking TiO2sol into the pores
of the AAM and removing it under vacuum; TiO2nanowiresare obtained after the sol is fully developed and the AAM isremoved In the procedure by Lee and co-workers,96a TTIPsolution was prepared by mixing TTIP with 2-propanol and2,4-pentanedione After the AAM was dipped into this
Figure 1 TEM images of TiO2 nanoparticles prepared by hydrolysis of Ti(OR)4 in the presence of tetramethylammonium hydroxide
Reprinted with permission from Chemseddine, A.; Moritz, T Eur J Inorg Chem 1999, 235 Copyright 1999 Wiley-VCH.
Figure 2 TEM images of uniform anatase TiO2nanoparticles Reprinted from Sugimoto, T.; Zhou, X.; Muramatsu, A J Colloid Interface
Sci 2003, 259, 53, Copyright 2003, with permission from Elsevier.
Trang 6solution, it was removed from the solution and placed under
vacuum until the entire volume of the solution was pulled
through the AAM The AAM was hydrolyzed by water vapor
over a HCl solution for 24 h, air-dried at room temperature,
and then calcined in a furnace at 673 K for 2 h and cooled
to room temperature with a temperature ramp of 2°C/h Pure
TiO2nanotubes were obtained after the AAM was dissolved
in a 6 M NaOH solution for several minutes.96Alternatively,
TiO2 nanotubes could be obtained by coating the AAM
membranes at 60°C for a certain period of time (12-48 h)
with dilute TiF4under pH ) 2.1 and removing the AAM
after TiO2nanotubes were fully developed.97Figure 4 shows
a typical SEM image of the TiO2nanotube array from the
AAM template.97
In another scheme, a ZnO nanorod array on a glass
substrate can be used as a template to fabricate TiO2
nanotubes with the sol-gel method.101Briefly, TiO sol is
deposited on a ZnO nanorod template by dip-coating with aslow withdrawing speed, then dried at 100°C for 10 min,and heated at 550 °C for 1 h in air to obtain ZnO/TiO2nanorod arrays The ZnO nanorod template is etched-up byimmersing the ZnO/TiO2nanorod arrays in a dilute hydro-chloric acid aqueous solution to obtain TiO2nanotube arrays.Figure 5 shows a typical SEM image of the TiO2nanotubearray with the ZnO nanorod array template The TiO2nanotubes inherit the uniform hexagonal cross-sectionalshape and the length of 1.5µm and inner diameter of 100-
120 nm of the ZnO nanorod template As the concentration
of the TiO2sol is constant, well-aligned TiO2nanotube arrayscan only be obtained from an optimal dip-coating cyclenumber in the range of 2-3 cycles A dense porous TiO2thick film with holes is obtained instead if the dip-coatingnumber further increases The heating rate is critical to theformation of TiO2 nanotube arrays When the heating rate
is extra rapid, e.g., above 6 °C min-1, the TiO2 coat willeasily crack and flake off from the ZnO nanorods due togreat tensile stress between the TiO2 coat and the ZnOtemplate, and a TiO2film with loose, porous nanostructure
is obtained
2.2 Micelle and Inverse Micelle Methods
Aggregates of surfactant molecules dispersed in a liquidcolloid are called micelles when the surfactant concentrationexceeds the critical micelle concentration (CMC) The CMC
is the concentration of surfactants in free solution inequilibrium with surfactants in aggregated form In micelles,the hydrophobic hydrocarbon chains of the surfactants areoriented toward the interior of the micelle, and the hydro-philic groups of the surfactants are oriented toward thesurrounding aqueous medium The concentration of the lipidpresent in solution determines the self-organization of themolecules of surfactants and lipids The lipids form a singlelayer on the liquid surface and are dispersed in solution belowthe CMC The lipids organize in spherical micelles at thefirst CMC (CMC-I), into elongated pipes at the second CMC(CMC-II), and into stacked lamellae of pipes at the lamellarpoint (LM or CMC-III) The CMC depends on the chemicalcomposition, mainly on the ratio of the head area and thetail length Reverse micelles are formed in nonaqueousmedia, and the hydrophilic headgroups are directed towardthe core of the micelles while the hydrophobic groups are
Figure 3 TEM image of anatase nanorods and a single nanorod
composed of small TiO2 nanoparticles or nanograins (inset)
Reprinted from Miao, L.; Tanemura, S.; Toh, S.; Kaneko, K.;
Tanemura, M J Cryst Growth 2004, 264, 246, Copyright 2004,
with permission from Elsevier
Figure 4 SEM image of TiO2nanotubes prepared from the AAO
template Reprinted with permission from Liu, S M.; Gan, L M.;
Liu, L H.; Zhang, W D.; Zeng, H C Chem Mater 2002, 14,
1391 Copyright 2002 American Chemical Society
Figure 5 SEM of a TiO2nanotube array; the inset shows the ZnOnanorod array template Reprinted with permission from Qiu, J J.;
Yu, W D.; Gao, X D.; Li, X M Nanotechnology 2006, 17, 4695.
Copyright 2006 IOP Publishing Ltd
Trang 7directed outward toward the nonaqueous media There is no
obvious CMC for reverse micelles, because the number of
aggregates is usually small and they are not sensitive to the
surfactant concentration Micelles are often globular and
roughly spherical in shape, but ellipsoids, cylinders, and
bilayers are also possible The shape of a micelle is a function
of the molecular geometry of its surfactant molecules and
solution conditions such as surfactant concentration,
tem-perature, pH, and ionic strength
Micelles and inverse micelles are commonly employed to
synthesize TiO2 nanomaterials.102-110 A statistical
experi-mental design method was conducted by Kim et al to
optimize experimental conditions for the preparation of TiO2
nanoparticles.103The values of H2O/surfactant, H2O/titanium
precursor, ammonia concentration, feed rate, and reaction
temperature were significant parameters in controlling TiO2
nanoparticle size and size distribution Amorphous TiO2
nanoparticles with diameters of 10-20 nm were synthesized
and converted to the anatase phase at 600°C and to the more
thermodynamically stable rutile phase at 900 °C Li et al
developed TiO2 nanoparticles with the chemical reactions
between TiCl4 solution and ammonia in a reversed
micro-emulsion system consisting of cyclohexane,
poly(oxyethyl-ene)5 nonyle phenol ether, and poly(oxyethylene)9 nonyle
phenol ether.104The produced amorphous TiO2nanoparticles
transformed into anatase when heated at temperatures from
200 to 750 °C and into rutile at temperatures higher than
750°C Agglomeration and growth also occurred at elevated
temperatures
Shuttle-like crystalline TiO2nanoparticles were synthesized
by Zhang et al with hydrolysis of titanium tetrabutoxide in
the presence of acids (hydrochloric acid, nitric acid, sulfuric
acid, and phosphoric acid) in NP-5 (Igepal
CO-520)-cyclohexane reverse micelles at room temperature.110 The
crystal structure, morphology, and particle size of the TiO2
nanoparticles were largely controlled by the reaction
condi-tions, and the key factors affecting the formation of rutile at
room temperature included the acidity, the type of acid used,
and the microenvironment of the reverse micelles
Ag-glomeration of the particles occurred with prolonged reaction
times and increasing the [H2O]/[NP-5] and [H2
O]/[Ti-(OC4H9)4] ratios When suitable acid was applied, round TiO2
nanoparticles could also be obtained Representative TEM
images of the shuttle-like and round-shaped TiO2
nanopar-ticles are shown in Figure 6 In the study carried out by Lim
et al., TiO2nanoparticles were prepared by the controlled
hydrolysis of TTIP in reverse micelles formed in CO2with
the surfactants ammonium carboxylate perfluoropolyether
(PFPECOO-NH4+) (MW 587) and poly(dimethyl amino
ethyl methacrylate-block-1H,1H,2H,2H-perfluorooctyl
meth-acrylate) (PDMAEMA-b-PFOMA).106It was found that the
crystallite size prepared in the presence of reverse micelles
increased as either the molar ratio of water to surfactant or
the precursor to surfactant ratio increased
The TiO2nanomaterials prepared with the above micelle
and reverse micelle methods normally have amorphous
structure, and calcination is usually necessary in order to
induce high crystallinity However, this process usually leads
to the growth and agglomeration of TiO2nanoparticles The
crystallinity of TiO2nanoparticles initially (synthesized by
controlled hydrolysis of titanium alkoxide in reverse micelles
in a hydrocarbon solvent) could be improved by annealing
in the presence of the micelles at temperatures considerably
lower than those required for the traditional calcination
treatment in the solid state.108This procedure could producecrystalline TiO2 nanoparticles with unchanged physicaldimensions and minimal agglomeration and allows thepreparation of highly crystalline TiO2nanoparticles, as shown
in Figure 7, from the study of Lin et al.108
2.3 Sol Method
The sol method here refers to the nonhydrolytic sol-gelprocesses and usually involves the reaction of titaniumchloride with a variety of different oxygen donor molecules,e.g., a metal alkoxide or an organic ether.111-119
Figure 6 TEM images of the shuttle-like and round-shaped (inset)
TiO2nanoparticles From: Zhang, D., Qi, L., Ma, J., Cheng, H J.
Mater Chem 2002, 12, 3677 (http://dx.doi.org/10.1039/b206996b).
s Reproduced by permission of The Royal Society of Chemistry
Figure 7 HRTEM images of a TiO2nanoparticle after annealing.Reprinted with permission from Lin, J.; Lin, Y.; Liu, P.; Meziani,
M J.; Allard, L F.; Sun, Y P J Am Chem Soc 2002, 124, 11514.
Copyright 2002 American Chemical Society
Trang 8The condensation between Ti-Cl and Ti-OR leads to the
formation of Ti-O-Ti bridges The alkoxide groups can
be provided by titanium alkoxides or can be formed in situ
by reaction of the titanium chloride with alcohols or ethers
In the method by Trentler and Colvin,119a metal alkoxide
was rapidly injected into the hot solution of titanium halide
mixed with trioctylphosphine oxide (TOPO) in heptadecane
at 300°C under dry inert gas protection, and reactions were
completed within 5 min For a series of alkyl substituents
including methyl, ethyl, isopropyl, and tert-butyl, the reaction
rate dramatically increased with greater branching of R, while
average particle sizes were relatively unaffected Variation
of X yielded a clear trend in average particle size, but without
a discernible trend in reaction rate Increased nucleophilicity
(or size) of the halide resulted in smaller anatase nanocrystals
Average sizes ranged from 9.2 nm for TiF4 to 3.8 nm for
TiI4 The amount of passivating agent (TOPO) influenced
the chemistry Reaction in pure TOPO was slower and
resulted in smaller particles, while reactions without TOPO
were much quicker and yielded mixtures of brookite, rutile,
and anatase with average particle sizes greater than 10 nm
Figure 8 shows typical TEM images of TiO2nanocrystals
developed by Trentler et al.119
In the method used by Niederberger and Stucky,111TiCl4
was slowly added to anhydrous benzyl alcohol under
vigorous stirring at room temperature and was kept at
40-150°C for 1-21 days in the reaction vessel The precipitate
was calcinated at 450°C for 5 h after thoroughly washing
The reaction between TiCl4and benzyl alcohol was found
suitable for the synthesis of highly crystalline anatase phase
TiO2 nanoparticles with nearly uniform size and shape at
very low temperatures, such as 40°C The particle size could
be selectively adjusted in the range of 4-8 nm with the
appropriate thermal conditions and a proper choice of the
relative amounts of benzyl alcohol and titanium tetrachloride
The particle growth depended strongly on temperature, and
lowering the titanium tetrachloride concentration led to a
considerable decrease of particle size.111
Surfactants have been widely used in the preparation of a
variety of nanoparticles with good size distribution and
dispersity.15,16Adding different surfactants as capping agents,
such as acetic acid and acetylacetone, into the reaction matrix
can help synthesize monodispersed TiO2nanoparticles.120,121For example, Scolan and Sanchez found that monodispersenonaggregated TiO2nanoparticles in the 1-5 nm range wereobtained through hydrolysis of titanium butoxide in the
presence of acetylacetone and p-toluenesulfonic acid at 60
°C.120The resulting nanoparticle xerosols could be dispersed
in water-alcohol or alcohol solutions at concentrationshigher than 1 M without aggregation, which is attributed tothe complexation of the surface by acetylacetonato ligandsand through an adsorbed hybrid organic-inorganic layer
made with acetylacetone, p-toluenesulfonic acid, and
wa-ter.120With the aid of surfactants, different sized and shaped TiO2nanorods can be synthesized.122-130For example, the growth
of high-aspect-ratio anatase TiO2nanorods has been reported
by Cozzoli and co-workers by controlling the hydrolysisprocess of TTIP in oleic acid (OA).122-126,130Typically, TTIPwas added into dried OA at 80-100 °C under inert gasprotection (nitrogen flow) and stirred for 5 min A 0.1-2 Maqueous base solution was then rapidly injected and kept at80-100°C for 6-12 h with stirring The bases employedincluded organic amines, such as trimethylamino-N-oxide,trimethylamine, tetramethylammonium hydroxide, tetrabut-ylammonium hydroxyde, triethylamine, and tributylamine
In this reaction, by chemical modification of the titaniumprecursor with the carboxylic acid, the hydrolysis rate oftitanium alkoxide was controlled Fast (in 4-6 h) crystal-lization in mild conditions was promoted with the use ofsuitable catalysts (tertiary amines or quaternary ammoniumhydroxides) A kinetically overdriven growth mechanism led
to the growth of TiO2nanorods instead of nanoparticles.123Typical TEM images of the TiO2 nanorods are shown inFigure 9.123
Recently, Joo et al.127and Zhang et al.129reported similarprocedures in obtaining TiO2nanorods without the use ofcatalyst Briefly, a mixture of TTIP and OA was used togenerate OA complexes of titanium at 80°C in 1-octadecene
Figure 8 TEM image of TiO2nanoparticles derived from reaction
of TiCl4and TTIP in TOPO/heptadecane at 300°C The inset shows
a HRTEM image of a single particle Reprinted with permission
from Trentler, T J.; Denler, T E.; Bertone, J F.; Agrawal, A.;
Colvin, V L J Am Chem Soc 1999, 121, 1613 Copyright 1999
American Chemical Society
Figure 9 TEM of TiO2nanorods The inset shows a HRTEM of
a TiO2nanorod Reprinted with permission from Cozzoli, P D.;
Kornowski, A.; Weller, H J Am Chem Soc 2003, 125, 14539.
Copyright 2003 American Chemical Society
Trang 9The injection of a predetermined amount of oleylamine at
260 °C led to various sized TiO2 nanorods.129 Figure 10
shows TEM images of TiO2nanorods with various lengths,
and 2.3 nm TiO2nanoparticles prepared with this method.129
In the surfactant-mediated shape evolution of TiO2
nano-crystals in nonaqueous media conducted by Jun et al.,128it
was found that the shape of TiO2 nanocrystals could be
modified by changing the surfactant concentration The
synthesis was accomplished by an alkyl halide elimination
reaction between titanium chloride and titanium
isopro-poxide Briefly, a dioctyl ether solution containing TOPO
and lauric acid was heated to 300°C followed by addition
of titanium chloride under vigorous stirring The reaction
was initiated by the rapid injection of TTIP and quenched
with cold toluene At low lauric acid concentrations,
bullet-and diamond-shaped nanocrystals were obtained; at higher
concentrations, rod-shaped nanocrystals or a mixture of
nanorods and branched nanorods was observed The
bullet-and diamond-shaped nanocrystals bullet-and nanorods were
elon-gated along the [001] directions The TiO2 nanorods were
found to simultaneously convert to small nanoparticles as a
function of the growth time, as shown in Figure 11, due to
the minimization of the overall surface energy via dissolution
and regrowth of monomers during an Ostwald ripening
2.4 Hydrothermal Method
Hydrothermal synthesis is normally conducted in steel
pressure vessels called autoclaves with or without Teflon
liners under controlled temperature and/or pressure with thereaction in aqueous solutions The temperature can beelevated above the boiling point of water, reaching thepressure of vapor saturation The temperature and the amount
of solution added to the autoclave largely determine theinternal pressure produced It is a method that is widely usedfor the production of small particles in the ceramics industry.Many groups have used the hydrothermal method to prepareTiO2nanoparticles.131-140For example, TiO2nanoparticlescan be obtained by hydrothermal treatment of peptizedprecipitates of a titanium precursor with water.134 Theprecipitates were prepared by adding a 0.5 M isopropanolsolution of titanium butoxide into deionized water ([H2O]/[Ti] ) 150), and then they were peptized at 70°C for 1 h inthe presence of tetraalkylammonium hydroxides (peptizer).After filtration and treatment at 240 °C for 2 h, theas-obtained powders were washed with deionized water andabsolute ethanol and then dried at 60°C Under the sameconcentration of peptizer, the particle size decreased withincreasing alkyl chain length The peptizers and theirconcentrations influenced the morphology of the particles.Typical TEM images of TiO2nanoparticles made with thehydrothermal method are shown in Figure 12.134
In another example, TiO2nanoparticles were prepared byhydrothermal reaction of titanium alkoxide in an acidicethanol-water solution.132Briefly, TTIP was added dropwise
to a mixed ethanol and water solution at pH 0.7 with nitricacid, and reacted at 240°C for 4 h The TiO nanoparticles
Figure 10 TEM images of TiO2nanorods with lengths of (A) 12 nm, (B) 30 nm, and (C) 16 nm (D) 2.3 nm TiO2nanoparticles Inset
in parts C and D: HR-TEM image of a single TiO2nanorod and nanoparticle Reprinted with permission from Zhang, Z.; Zhong, X.; Liu,
S.; Li, D.; Han, M Angew Chem., Int Ed 2005, 44, 3466 Copyright 2005 Wiley-VCH.
Trang 10synthesized under this acidic ethanol-water environment
were mainly primary structure in the anatase phase without
secondary structure The sizes of the particles were controlled
to the range of 7-25 nm by adjusting the concentration of
Ti precursor and the composition of the solvent system
Besides TiO2nanoparticles, TiO2nanorods have also beensynthesized with the hydrothermal method.141-146Zhang et
al obtained TiO2nanorods by treating a dilute TiCl4solution
at 333-423 K for 12 h in the presence of acid or inorganicsalts.141,143-146Figure 13 shows a typical TEM image of theTiO2 nanorods prepared with the hydrothermal method.141The morphology of the resulting nanorods can be tuned withdifferent surfactants146or by changing the solvent composi-tions.145A film of assembled TiO2nanorods deposited on aglass wafer was reported by Feng et al.142 These TiO2nanorods were prepared at 160°C for 2 h by hydrothermaltreatment of a titanium trichloride aqueous solution super-saturated with NaCl
TiO2nanowires have also been successfully obtained withthe hydrothermal method by various groups.147-151Typically,TiO2nanowires are obtained by treating TiO2white powders
in a 10-15 M NaOH aqueous solution at 150-200°C for24-72 h without stirring within an autoclave Figure 14shows the SEM images of TiO2nanowires and a TEM image
of a single nanowire prepared by Zhang and co-workers.150TiO2nanowires can also be prepared from layered titanateparticles using the hydrothermal method as reported by Wei
Figure 11 Time dependent shape evolution of TiO2 nanorods:
(a) 0.25 h; (b) 24 h; (c) 48 h Scale bar ) 50 nm Reprinted with
permission from Jun, Y W.; Casula, M F.; Sim, J H.; Kim, S
Y.; Cheon, J.; Alivisatos, A P J Am Chem Soc 2003, 125, 15981.
Copyright 2003 American Chemical Society
Figure 12 TEM images of TiO2 nanoparticles prepared by thehydrothermal method Reprinted from Yang, J.; Mei, S.; Ferreira,
J M F Mater Sci Eng C 2001, 15, 183, Copyright 2001, with
permission from Elsevier
Figure 13 TEM image of TiO2 nanorods prepared with thehydrothermal method Reprinted with permission from Zhang, Q.;
Gao, L Langmuir 2003, 19, 967 Copyright 2003 American
Chemical Society
Trang 11et al.152In their experiment, layer-structured Na2Ti3O7was
dispersed into a 0.05-0.1 M HCl solution and kept at
140-170°C for 3-7 days in an autoclave TiO2nanowires were
obtained after the product was washed with H2O and finally
dried In the formation of a TiO2 nanowire from layered
H2Ti3O7, there are three steps: (i) the exfoliation of layered
Na2Ti3O7; (ii) the nanosheets formation; and (iii) the
nanow-ires formation.152In Na2Ti3O7, [TiO6] octahedral layers are
held by the strong static interaction between the Na+cations
between the [TiO6] octahedral layers and the [TiO6] unit
When the larger H3+O cations replace the Na+cations in
the interlayer space of [TiO6] sheets, this static interaction
is weakened because the interlayer distance is enlarged As
a result, the layered compounds Na2Ti3O7 are gradually
exfoliated When Na+is exchanged by H+in the dilute HCl
solution, numerous H2Ti3O7 sheet-shaped products are
formed Since the nanosheet does not have inversion
sym-metry, an intrinsic tension exists The nanosheets split to form
nanowires in order to release the strong stress and lower the
total energy.152 A representative TEM image of TiO2
nanowires from Na2Ti3O7is shown in Figure 15.152
The hydrothermal method has been widely used to prepare
TiO2nanotubes since it was introduced by Kasuga et al in
1998.153-175Briefly, TiO2powders are put into a 2.5-20 M
NaOH aqueous solution and held at 20-110°C for 20 h in
an autoclave TiO2nanotubes are obtained after the products
are washed with a dilute HCl aqueous solution and distilled
water They proposed the following formation process of
TiO2nanotubes.154When the raw TiO2material was treated
with NaOH aqueous solution, some of the Ti-O-Ti bonds
were broken and Ti-O-Na and Ti-OH bonds were formed
New Ti-O-Ti bonds were formed after the Ti-O-Na and
Ti-OH bonds reacted with acid and water when the material
was treated with an aqueous HCl solution and distilled water
The Ti-OH bond could form a sheet Through the
dehydra-tion of Ti-OH bonds by HCl aqueous soludehydra-tion, Ti-O-Ti
bonds or Ti-O-H-O-Ti hydrogen bonds were generated
The bond distance from one Ti to the next Ti on the surface
decreased This resulted in the folding of the sheets and the
connection between the ends of the sheets, resulting in theformation of a tube structure In this mechanism, the TiO2nanotubes were formed in the stage of the acid treatmentfollowing the alkali treatment Figure 16 shows typical TEMimages of TiO2nanotubes made by Kasuga et al.153However,
Du and co-workers found that the nanotubes were formedduring the treatment of TiO2in NaOH aqueous solution.161
A 3D f 2D f 1D formation mechanism of the TiO2nanotubes was proposed by Wang and co-workers.171It statedthat the raw TiO2 was first transformed into lamellarstructures and then bent and rolled to form the nanotubes.For the formation of the TiO2nanotubes, the two-dimensionallamellar TiO2 was essential Yao and co-workers furthersuggested, based on their HRTEM study as shown in Figure
Figure 14 SEM images of TiO2nanowires with the inset showing
a TEM image of a single TiO2nanowire with a [010] selected area
electron diffraction (SAED) recorded perpendicular to the long axis
of the wire Reprinted from Zhang, Y X.; Li, G H.; Jin, Y X.;
Zhang, Y.; Zhang, J.; Zhang, L D Chem Phys Lett 2002, 365,
300, Copyright 2002, with permission from Elsevier
Figure 15 TEM images of TiO2nanowires made from the layered
Na2Ti3O7 particles, with the HRTEM image shown in the inset.Reprinted from Wei, M.; Konishi, Y.; Zhou, H.; Sugihara, H.;
Arakawa, H Chem Phys Lett 2004, 400, 231, Copyright 2004,
with permission from Elsevier
Figure 16. TEM image of TiO2 nanotubes Reprinted withpermission from Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino,
T.; Niihara, K Langmuir 1998, 14, 3160 Copyright 1998 American
Chemical Society
Trang 1217, that TiO2nanotubes were formed by rolling up the
single-layer TiO2 sheets with a rolling-up vector of [001] and
attracting other sheets to surround the tubes.172Bavykin and
co-workers suggested that the mechanism of nanotube
formation involved the wrapping of multilayered nanosheets
rather than scrolling or wrapping of single layer nanosheets
followed by crystallization of successive layers.156 In the
mechanism proposed by Wang et al., the formation of TiO2
nanotubes involved several steps.176During the reaction with
NaOH, the Ti-O-Ti bonding between the basic building
blocks of the anatase phase, the octahedra, was broken and
a zigzag structure was formed when the free octahedras
shared edges between the Ti ions with the formation of
hydroxy bridges, leading to the growth along the [100]
direction of the anatase phase Two-dimensional crystalline
sheets formed from the lateral growth of the formation of
oxo bridges between the Ti centers (Ti-O-Ti bonds) in the
[001] direction and rolled up in order to saturate these
dangling bonds from the surface and lower the total energy,
resulting in the formation of TiO2nanotubes.176
2.5 Solvothermal Method
The solvothermal method is almost identical to the
hydrothermal method except that the solvent used here is
nonaqueous However, the temperature can be elevated much
higher than that in hydrothermal method, since a variety of
organic solvents with high boiling points can be chosen The
solvothermal method normally has better control than
hy-drothermal methods of the size and shape distributions and
the crystallinity of the TiO2nanoparticles The solvothermal
method has been found to be a versatile method for the
synthesis of a variety of nanoparticles with narrow sizedistribution and dispersity.177-179The solvothermal methodhas been employed to synthesize TiO2 nanoparticles andnanorods with/without the aid of surfactants.177-185 Forexample, in a typical procedure by Kim and co-workers,184TTIP was mixed with toluene at the weight ratio of 1-3:10and kept at 250°C for 3 h The average particle size of TiO2powders tended to increase as the composition of TTIP inthe solution increased in the range of weight ratio of 1-3:
10, while the pale crystalline phase of TiO2was not produced
at 1:20 and 2:5 weight ratios.184By controlling the lyzation reaction of Ti(OC4H9)4and linoleic acid, redispers-ible TiO2nanoparticles and nanorods could be synthesized,
hydro-as found by Li et al recently.177The decomposition of NH4HCO3could provide H2O for the hydrolyzation reaction, andlinoleic acid could act as the solvent/reagent and coordinationsurfactant in the synthesis of nanoparticles Triethylaminecould act as a catalyst for the polycondensation of the Ti-O-Ti inorganic network to achieve a crystalline product andhad little influence on the products’ morphology The chainlengths of the carboxylic acids had a great influence on theformation of TiO2, and long-chain organic acids wereimportant and necessary in the formation of TiO2.177Figure
-18 shows a representative TEM image of TiO2nanoparticlesfrom their study.177
TiO2nanorods with narrow size distributions can also bedeveloped with the solvothermal method.177,183For example,
in a typical synthesis from Kim et al., TTIP was dissolved
in anhydrous toluene with OA as a surfactant and kept at
250 °C for 20 h in an autoclave without stirring.183 Longdumbbell-shaped nanorods were formed when a sufficientamount of TTIP or surfactant was added to the solution, due
to the oriented growth of particles along the [001] axis At
a fixed precursor to surfactant weight ratio of 1:3, theconcentration of rods in the nanoparticle assembly increased
as the concentration of the titanium precursor in the solutionincreased The average particle size was smaller and the sizedistribution was narrower than is the case for particlessynthesized without surfactant The crystalline phase, diam-eter, and length of these nanorods are largely influenced bythe precursor/surfactant/solvent weight ratio Anatase nano-
Figure 17 (a) HRTEM images of TiO2 nanotubes (b)
Cross-sectional view of TiO2 nanotubes Reused with permission from
B D Yao, Y F Chan, X Y Zhang, W F Zhang, Z Y Yang, N
Wang, Applied Physics Letters 82, 281 (2003) Copyright 2003,
American Institute of Physics
Figure 18 TEM micrographs of TiO2nanoparticles prepared withthe solvothermal method Reprinted with permission from Li, X
L.; Peng, Q.; Yi, J X.; Wang, X.; Li, Y D Chem.sEur J 2006,
12, 2383 Copyright 2006 Wiley-VCH.
Trang 13rods were obtained from the solution with a precursor/
surfactant weight ratio of more than 1:3 for a precursor/
solvent weight ratio of 1:10 or from the solution with a
precursor/solvent weight ratio of more than 1:5 for a
precursor/surfactant weight ratio of 1:3 The diameter and
length of these nanorods were in the ranges of 3-5 nm and
18-25 nm, respectively Figure 19 shows a typical TEM
image of TiO2 nanorods prepared from the solutions with
the weight ratio of precursor/solvent/surfactant ) 1:5:3.183
Similar to the hydrothermal method, the solvothermal
method has also been used for the preparation of TiO2
nanowires.180-182Typically, a TiO2powder suspension in an
5 M NaOH water-ethanol solution is kept in an autoclave
at 170-200°C for 24 h and then cooled to room temperature
naturally TiO2 nanowires are obtained after the obtained
sample is washed with a dilute HCl aqueous solution and
dried at 60 °C for 12 h in air.181 The solvent plays an
important role in determining the crystal morphology
Solvents with different physical and chemical properties can
influence the solubility, reactivity, and diffusion behavior
of the reactants; in particular, the polarity and coordinating
ability of the solvent can influence the morphology and the
crystallization behavior of the final products The presence
of ethanol at a high concentration not only can cause the
polarity of the solvent to change but also strongly affects
the ζ potential values of the reactant particles and the
increases solution viscosity For example, in the absence of
ethanol, short and wide flakelike structures of TiO2 were
obtained instead of nanowires When chloroform is used,
TiO2nanorods were obtained.181Figure 20 shows
representa-tive TEM images of the TiO2nanowires prepared from the
solvothermal method.181Alternatively, bamboo-shaped
Ag-doped TiO2nanowires were developed with titanium
butox-ide as precursor and AgNO3 as catalyst.180 Through the
electron diffraction (ED) pattern and HRTEM study, the Ag
phase only existed in heterojunctions between single-crystalTiO2nanowires.180
2.6 Direct Oxidation Method
TiO2 nanomaterials can be obtained by oxidation oftitanium metal using oxidants or under anodization Crystal-line TiO2nanorods have been obtained by direct oxidation
of a titanium metal plate with hydrogen peroxide.186-191Typically, TiO2nanorods on a Ti plate are obtained when acleaned Ti plate is put in 50 mL of a 30 wt % H2O2solution
at 353 K for 72 h The formation of crystalline TiO2occursthrough a dissolution precipitation mechanism By theaddition of inorganic salts of NaX (X ) F-, Cl-, and SO42-),the crystalline phase of TiO2 nanorods can be controlled.The addition of F-and SO42-helps the formation of pureanatase, while the addition of Cl- favors the formation ofrutile.189 Figure 21 shows a typical SEM image of TiO2nanorods prepared with this method.186
At high temperature, acetone can be used as a good oxygensource and for the preparation of TiO nanorods by oxidizing
Figure 19 TEM micrographs and electron diffraction patterns of
products prepared from solutions at the weight ratio of precursor/
solvent/surfactant ) 1:5:3 Reprinted from Kim, C S.; Moon, B
K.; Park, J H.; Choi, B C.; Seo, H J J Cryst Growth 2003, 257,
309, Copyright 2003, with permission from Elsevier
Figure 20 TEM images of TiO2 nanowires synthesized by the
solvothermal method From: Wen, B.; Liu, C.; Liu, Y New J.
Chem 2005, 29, 969 (http://dx.doi.org/10.1039/b502604k) s
Reproduced by permission of The Royal Society of Chemistry(RSC) on behalf of the Centre National de la Recherche Scientifique(CNRS)
Figure 21. SEM morphology of TiO2 nanorods by directlyoxidizing a Ti plate with a H2O2solution Reprinted from Wu, J
M J Cryst Growth 2004, 269, 347, Copyright 2004, with
permission from Elesevier
Trang 14a Ti plate with acetone as reported by Peng and Chen.192
The oxygen source was found to play an important role
Highly dense and well-aligned TiO2 nanorod arrays were
formed when acetone was used as the oxygen source, and
only crystal grain films or grains with random nanofibers
growing from the edges were obtained with pure oxygen or
argon mixed with oxygen The competition of the oxygen
and titanium diffusion involved in the titanium oxidation
process largely controlled the morphology of the TiO2 With
pure oxygen, the oxidation occurred at the Ti metal and the
TiO2interface, since oxygen diffusion predominated because
of the high oxygen concentration When acetone was used
as the oxygen source, Ti cations diffused to the oxide surface
and reacted with the adsorbed acetone species Figure 22
shows aligned TiO2nanorod arrays obtained by oxidizing a
titanium substrate with acetone at 850°C for 90 min.192
As extensively studied, TiO2nanotubes can be obtained
by anodic oxidation of titanium foil.193-228 In a typical
experiment, a clean Ti plate is anodized in a 0.5% HF
solution under 10-20 V for 10-30 min Platinum is used
as counterelectrode Crystallized TiO2nanotubes are obtained
after the anodized Ti plate is annealed at 500°C for 6 h in
oxygen.210The length and diameter of the TiO2nanotubes
could be controlled over a wide range (diameter, 15-120
nm; length, 20 nm to 10 µm) with the applied potential
between 1 and 25 V in optimized phosphate/HF
electro-lytes.229Figure 23 shows SEM images of TiO2nanotubes
created with this method.208
2.7 Chemical Vapor Deposition
Vapor deposition refers to any process in which materials
in a vapor state are condensed to form a solid-phase material
These processes are normally used to form coatings to alter
the mechanical, electrical, thermal, optical, corrosion
resis-tance, and wear resistance properties of various substrates
They are also used to form free-standing bodies, films, and
fibers and to infiltrate fabric to form composite materials
Recently, they have been widely explored to fabricate various
nanomaterials Vapor deposition processes usually take place
within a vacuum chamber If no chemical reaction occurs,
this process is called physical vapor deposition (PVD);
otherwise, it is called chemical vapor deposition (CVD) InCVD processes, thermal energy heats the gases in the coatingchamber and drives the deposition reaction
Thick crystalline TiO2films with grain sizes below 30 nm
as well as TiO2nanoparticles with sizes below 10 nm can
be prepared by pyrolysis of TTIP in a mixed helium/oxygenatmosphere, using liquid precursor delivery.230When depos-ited on the cold areas of the reactor at temperatures below
90°C with plasma enhanced CVD, amorphous TiO2particles can be obtained and crystallize with a relativelyhigh surface area after being annealed at high temperatures.231TiO2nanorod arrays with a diameter of about 50-100 nmand a length of 0.5-2 µm can be synthesized by metal
nano-organic CVD (MOCVD) on a WC-Co substrate using TTIP
as the precursor.232Figure 24 shows the TiO2nanorods grown on fused silicasubstrates with a template- and catalyst-free MOCVDmethod.233In a typical procedure, titanium acetylacetonate(Ti(C10H14O5)) vaporizing in the low-temperature zone of afurnace at 200-230°C is carried by a N2/O2flow into thehigh-temperature zone of 500-700°C, and TiO2nanostruc-tures are grown directly on the substrates The phase and
Figure 22 SEM images of large-scale nanorod arrays prepared
by oxidizing a titanium with acetone at 850°C for 90 min From:
Peng, X.; Chen, A J Mater Chem 2004, 14, 2542 (http://
dx.doi.org/10.1039/b404750h) s Reproduced by permission of The
Royal Society of Chemistry
Figure 23 SEM images of TiO2nanotubes prepared with anodicoxidation Reprinted with permission from Varghese, O K.; Gong,
D.; Paulose, M.; Ong, K G.; Dickey, E C.; Grimes, C A AdV.
Mater 2003, 15, 624 Copyright 2003 Wiley-VCH.
Figure 24 SEM images of TiO2 nanorods grown at 560 °C
Reprinted with permission from Wu, J J.; Yu, C C J Phys Chem.
B 2004, 108, 3377 Copyright 2004 American Chemical Society.
Trang 15morphology of the TiO2 nanostructures can be tuned with
the reaction conditions For example, at 630 and 560 °C
under a pressure of 5 Torr, single-crystalline rutile and
anatase TiO2nanorods were formed respectively, while, at
535 °C under 3.6 Torr, anatase TiO2nanowalls composed
of well-aligned nanorods were formed.233
In addition to the above CVD approaches in preparing
TiO2nanomaterials, other CVD approaches are also used,
such as electrostatic spray hydrolysis,234 diffusion flame
pyrolysis,235-239 thermal plasma pyrolysis,240-246ultrasonic
spray pyrolysis,247laser-induced pyrolysis,248,249and
ultronsic-assisted hydrolysis,250,251among others
2.8 Physical Vapor Deposition
In PVD, materials are first evaporated and then condensed
to form a solid material The primary PVD methods include
thermal deposition, ion plating, ion implantation, sputtering,
laser vaporization, and laser surface alloying TiO2nanowire
arrays have been fabricated by a simple PVD method or
thermal deposition.252-254Typically, pure Ti metal powder
is on a quartz boat in a tube furnace about 0.5 mm away
from the substrate Then the furnace chamber is pumped
down to∼300 Torr and the temperature is increased to 850
°C under an argon gas flow with a rate of 100 sccm and
held for 3 h After the reaction, a layer of TiO2nanowires
can be obtained.254 A layer of Ti nanopowders can be
deposited on the substrate before the growth of TiO2
nanowires,252,253and Au can be employed as catalyst.252A
typical SEM image of TiO2nanowires made with the PVD
method is shown in Figure 25.252
2.9 Electrodeposition
Electrodeposition is commonly employed to produce a
coating, usually metallic, on a surface by the action of
reduction at the cathode The substrate to be coated is used
as cathode and immersed into a solution which contains a
salt of the metal to be deposited The metallic ions are
attracted to the cathode and reduced to metallic form With
the use of the template of an AAM, TiO2nanowires can be
obtained by electrodeposition.255,256In a typical process, the
electrodeposition is carried out in 0.2 M TiCl solution with
pH ) 2 with a pulsed electrodeposition approach, andtitanium and/or its compound are deposited into the pores
of the AAM By heating the above deposited template at
500°C for 4 h and removing the template, pure anatase TiO2nanowires can be obtained Figure 26 shows a representativeSEM image of TiO2nanowires.256
2.10 Sonochemical Method
Ultrasound has been very useful in the synthesis of a widerange of nanostructured materials, including high-surface-area transition metals, alloys, carbides, oxides, and colloids.The chemical effects of ultrasound do not come from a directinteraction with molecular species Instead, sonochemistryarises from acoustic cavitation: the formation, growth, andimplosive collapse of bubbles in a liquid Cavitationalcollapse produces intense local heating (∼5000 K), high pres-
sures (∼1000 atm), and enormous heating and cooling rates
(>109K/s) The sonochemical method has been applied toprepare various TiO2nanomaterials by different groups.257-269
Yu et al applied the sonochemical method in preparinghighly photoactive TiO2 nanoparticle photocatalysts withanatase and brookite phases using the hydrolysis of titaniumtetraisoproproxide in pure water or in a 1:1 EtOH-H2Osolution under ultrasonic radiation.109Huang et al found thatanatase and rutile TiO2nanoparticles as well as their mixturescould be selectively synthesized with various precursorsusing ultrasound irradiation, depending on the reactiontemperature and the precursor used.259Zhu et al developedtitania whiskers and nanotubes with the assistance ofsonication as shown in Figure 27.269They found that arrays
of TiO2nanowhiskers with a diameter of 5 nm and nanotubeswith a diameter of∼5 nm and a length of 200-300 nm could
be obtained by sonicating TiO2particles in NaOH aqueoussolution followed by washing with deionized water and adilute HNO3aqueous solution
2.11 Microwave Method
A dielectric material can be processed with energy in theform of high-frequency electromagnetic waves The principal
Figure 25 SEM images of the TiO2nanowire arrays prepared by
the PVD method Reprinted from Wu, J M.; Shih, H C.; Wu, W
T Chem Phys Lett 2005, 413, 490, Copyright 2005, with
permission from Elsevier Figure 26 Cross-sectional SEM image of TiO2nanowires
elec-trodeposited in AAM pores Reprinted from Liu, S.; Huang, K
Sol Energy Mater Sol Cells 2004, 85, 125, Copyright 2004, with
permission from Elsevier
Trang 16frequencies of microwave heating are between 900 and 2450
MHz At lower microwave frequencies, conductive currents
flowing within the material due to the movement of ionic
con-stituents can transfer energy from the microwave field to the
material At higher frequencies, the energy absorption is
pri-marily due to molecules with a permanent dipole which tend
to reorientate under the influence of a microwave electric
field This reorientation loss mechanism originates from the
inability of the polarization to follow extremely rapid
rever-sals of the electric field, so the polarization phasor lags the
applied electric field This ensures that the resulting current
density has a component in phase with the field, and therefore
power is dissipated in the dielectric material The major
advantages of using microwaves for industrial processing are
rapid heat transfer, and volumetric and selective heating
Microwave radiation is applied to prepare various TiO2
nanomaterials.270-276Corradi et al found that colloidal titania
nanoparticle suspensions could be prepared within 5 min to
1 h with microwave radiation, while 1 to 32 h was needed
for the conventional synthesis method of forced hydrolysis
at 195°C.270Ma et al developed high-quality rutile TiO2
nano-rods with a microwave hydrothermal method and found that
they aggregated radially into spherical secondary
nanopartic-les.272Wu et al synthesized TiO2nanotubes by microwave
radiation via the reaction of TiO2crystals of anatase, rutile,
or mixed phase and NaOH aqueous solution under a certain
microwave power.275Normally, the TiO2nanotubes had the
central hollow, open-ended, and multiwall structure with
diameters of 8-12 nm and lengths up to 200-1000 nm.275
2.12 TiO2 Mesoporous/Nanoporous Materials
In the past decade, mesoporous/nanoporous TiO2materials
have been well studied with or without the use of organic
surfactant templates.28,80,264,265,277-312Barbe et al reported thepreparation of a mesoporous TiO2film by the hydrothermalmethod as shown Figure 28.80In a typical experiment, TTIPwas added dropwise to a 0.1 M nitric acid solution undervigorous stirring and at room temperature A white precipitateformed instantaneously Immediately after the hydrolysis, thesolution was heated to 80°C and stirred vigorously for 8 hfor peptization The solution was then filtered on a glass frit
to remove agglomerates Water was added to the filtrate toadjust the final solids concentration to∼5 wt % The solution
was put in a titanium autoclave for 12 h at 200-250°C.After sonication, the colloidal suspension was put in a rotaryevaporator and evaporated to a final TiO2concentration of
11 wt % The precipitation pH, hydrolysis rate, autoclaving
pH, and precursor chemistry were found to influence themorphology of the final TiO2nanoparticles
Alternative procedures without the use of hydrothermalprocesses have been reported by Liu et al.292and Zhang et
al.311 In the report by Liu et al., 24.0 g of titanium(IV)
n-butoxide ethanol solution (weight ratio of 1:7) was
prehydrolyzed in the presence of 0.32 mL of a 0.28 M HNO3aqueous solution (TBT/HNO3∼ 100:1) at room temperature
for 3 h 0.32 mL of deionized water was added to theprehydrolyzed solution under vigorous stirring and stirredfor an additional 2 h The sol solution in a closed vesselwas kept at room temperature without stirring to gel andage After aging for 14 days, the gel was dried at roomtemperature, ground into a fine powder, washed thoroughlywith water and ethanol, and dried to produce porous TiO2.Upon calcination at 450 °C for 4 h under air, crystallizedmesoporous TiO2material was obtained.292
Yu et al prepared three-dimensional and thermally stablemesoporous TiO2 without the use of any surfactants.265Briefly, monodispersed TiO2 nanoparticles were formedinitially by ultrasound-assisted hydrolysis of acetic acid-modified titanium isopropoxide Mesoporous spherical orglobular particles were then produced by controlled conden-
Figure 27 TEM images of TiO2nanotubes (A) and nanowhiskers
(B) prepared with the sonochemical method From: Zhu, Y.; Li,
H.; Koltypin, Y.; Hacohen, Y R.; Gedanken, A Chem Commun.
2001, 2616 (http://dx.doi.org/10.1039/b108968b) s Reproduced by
permission of The Royal Society of Chemistry
Figure 28 SEM image of the mesoporous TiO2film synthesizedfrom the acetic acid-modified precursor and autoclaved at 230°C.Reprinted with permission from Barbe, C J.; Arendse, F.; Comte,
P.; Jirousek, M.; Lenzmann, F.; Shklover, V.; Gra¨tzel, M J Am.
Ceram Soc 1997, 80, 3157 Copyright 1997 Blackwell Publishing.
Trang 17sation and agglomeration of these sol nanoparticles under
high-intensity ultrasound radiation The mesoporous TiO2had
a wormhole-like structure consisting of TiO2nanoparticles
and a lack of long-range order.265
In the template method used by the Stucky
group278-280,287,295,302,306-307,313and other groups,264,293,297,303,309
structure-directing agents were used for organizing
network-forming metal oxide species in nonaqueous solutions These
structure-directing agents were also called organic templates
The most commonly used organic templates were
amphi-philic poly(alkylene oxide) block copolymers, such as
HO-(CH2CH2O)20(CH2CH(CH3)O)70(CH2CH2O)20H (designated
EO20PO70EO20, called Pluronic P-123) and HO(CH2CH2O)106
-(CH2CH(CH3)O)70(CH2CH2O)106H (designated EO106PO70
-EO106, called Pluronic F-127) In a typical synthesis,
poly-(alkylene oxide) block copolymer was dissolved in ethanol
Then TiCl4precursor was added with vigorous stirring The
resulting sol solution was gelled in an open Petri dish at 40
°C in air for 1-7 days Mesoporous TiO2was obtained after
removing the surfactant species by calcining the as-made
sample at 400°C for 5 h in air.306Figure 29 shows typical
TEM images of the mesoporous TiO2 Besides triblock
co-polymers as structure-directing agents, diblock co-polymers were
also used such as [CnH2n-1(OCH2CH2)y OH, Brij 56 (B56, n/y
) 16/10) or Brij 58 (B58, n/y ) 16/20)] by Sanchez et al.285
Other surfactants employed to direct the formation of
mesoporous TiO2include tetradecyl phosphate (a 14-carbon
chain) by Antonelli and Ying277and commercially available
dodecyl phosphate by Putnam and co-workers,298
cetyltri-methylammonium bromide (CTAB) (a cationic
surfac-tant),281,283,296the recent Gemini surfactant,294and
dodecyl-amine (a neutral surfactant).304 Carbon nanotubes310 and
mesoporous SBA-15286have also been used as the skeleton
for mesoporous TiO2
2.13 TiO2 Aerogels
The study of TiO2 aerogels is worthy of special
men-tion.314-326 The combination of sol-gel processing with
supercritical drying offers the synthesis of TiO2aerogels with
morphological and chemical properties that are not easily
achieved by other preparation methods, i.e., with high surface
area Campbell et al prepared TiO2 aerogels by sol-gel
synthesis from titanium n-butoxide in methanol with the
subsequent removal of solvent by supercritical CO2.315For
a typical synthesis process, titanium n-butoxide was added
to 40 mL of methanol in a dry glovebox This solution was
combined with another solution containing 10 mL of
methanol, nitric acid, and deionized water The concentration
of the titanium n-butoxide was kept at 0.625 M, and the
molar ratio of water/HNO3/titanium n-butoxide was 4:0.1:
1 The gel was allowed to age for 2 h and then extracted in
a standard autoclave with supercritical CO2at a flow rate of
24.6 L/h, at 343 K under 2.07× 107Pa for 2-3 h, resulting
in complete removal of solvent After extraction, the sample
was heated in a vacuum oven at 3.4 kPa and 383 K for 3 h
to remove the residual solvent and at 3.4 kPa and 483 K for
3 h to remove any residual organics The pretreated sample
had a brown color and turned white after calcination at 773
K or above The resulting TiO2aerogel, after calcination at
773 K for 2 h, had a BET surface area of >200 m2/g,
contained mesopores in the range 2-10 nm, and was of the
pure anatase form Dagan et al found the TiO2 aerogels
obtanied by using a Ti/ethanol/H2O/nitric acid ratio of 1:20:
3:0.08 could have a porosity of 90% and surface areas of
600 m2/g, as compared to a surface area of 50 m2/g for TiO2P25.316,317Figure 30 shows a typical SEM image of a TiO2aerogel with a surface area of 447 m2/g and an interporestructure constructed by near uniform grains of ellipticalshapes with 30 nm× 50 nm axes.326
Figure 29. TEM micrographs of two-dimensional hexagonalmesoporous TiO2recorded along the (a) [110] and (b) [001] zoneaxes, respectively The inset in part a is selected-area electrondiffraction patterns obtained on the image area (c) TEM image ofcubic mesoporous TiO2accompanied by the corresponding (inset)EDX spectrum Reprinted with permission from Yang, P.; Zhao,
D.; Margolese, D I.; Chmelka, B F.; Stucky, G D Chem Mater.
1999, 11, 2813 Copyright 1999 American Chemical Society.
Trang 182.14 TiO2 Opal and Photonic Materials
The syntheses of TiO2opal and photonic materials have
been well studied by various groups.327-358Holland et al
reported the preparation of TiO2 inverse opal from the
corresponding metal alkoxides, using latex spheres as
templates.334,335Millimeter-thick layers of latex spheres were
deposited on filter paper in a Buchner funnel under vacuum
and soaked with ethanol Titanium ethoxide was added
dropwise to cover the latex spheres completely while suction
was applied Typical mass ratios of alkoxide to latex were
between 1.4 and 3 After drying the composite in a vacuum
desiccator for 3 to 24 h, the latex spheres were removed by
calcination in flowing air at 575 °C for 7 to 12 h, leaving
hard and brittle powder particles with 320- to 360-nm voids
The carbon content of the calcined samples varied from 0.4
to 1.0 wt %, indicating that most of the latex templates had
been removed from the 3D host Figure 31 shows an
illustration of the simple synthesis of TiO2inverse opal and
an SEM image of TiO2inverse opals Similar studies have
also been carried out by other researchers.327,356
Dong and Marlow prepared TiO2inversed opals with a
skeleton-like structure of TiO2rods by a template-directed
method using monodispersed polystyrene particles of size
270 nm.328-330,345Infiltration of a titania precursor (Ti(i-OPr)4
in EtOH) was followed by a drying and calcination
proce-dure The precursor concentration was varied from 30% to
100%, and the calcination temperature was tuned from 300
to 700°C A SEM picture of the TiO2inversed opal is shown
in Figure 32.329The skeleton structure consists of
rhombo-hedral windows and TiO2cylinders forming a highly regular
network The cylinders connect the centers of the former
octahedral and tetrahedral voids of the opal These voids form
a CaF2 lattice which is filled with cylindrical bonds
con-necting the Ca and F sites
Wang et al reported their study on the large-scale
fabrication of ordered TiO2nanobowl arrays.354The process
starts with a self-assembled monolayer of polystyrene (PS)
spheres, which is used as a template for atomic layer
deposition of a TiO2layer After ion-milling, toluene-etching,
and annealing of the TiO2-coated spheres, ordered arrays of
nanostructured TiO2nanobowls can be fabricated as shown
in Figure 33
Wang et al fabricated a 2D photonic crystal by coating
patterned and aligned ZnO nanorod arrays with TiO2.355PS
spheres were self-assembled to make a monolayer mask on
a sapphire substrate, which was then covered with a layer
of gold After removing the PS spheres with toluene, ZnOnanorods were grown using a vapor-liquid-solid process.Finally, a TiO2 layer was deposited on the ZnO nanorods
by introducing TiCl4and water vapors into the atomic layerdeposition chamber at 100°C Figure 34 shows SEM images
of a ZnO nanorod array and the TiO2-coated ZnO nanorodarray
Li et al reported the preparation of ordered arrays of TiO2opals using opal gel templates under uniaxial compression
at ambient temperature during the TiO2sol/gel process.337The aspect ratio was controllable by the compression degree,
R Polystyrene inverse opal was template synthesized using
silica opals as template The silica was removed with 40 wt
% aqueous hydrofluoric acid Monomer solutions consisting
of dimethylacrylamide, acrylic acid, and amide in 1:1:0.02 weight ratios were dissolved in a water/
methylenebisacryl-Figure 30. SEM image of a TiO2 aerogel Reprinted with
permission from Zhu, Z.; Tsung, L Y.; Tomkiewicz, M J Phys.
Chem 1995, 99, 15945 Copyright 1995 American Chemical
Society
Figure 31 (A) Schematic illustration of the synthesis of a TiO2
inversed opal (B) SEM image of the TiO2inversed opal Reprinted
with permission from Holland, B T.; Blanford, C.; Stein, A Science
1998, 281, 538 (http://www.sciencemag.org) Copyright 1998
AAAS
Trang 19ethanol mixture (4:7 wt/wt) with total monomer content 30
wt % Ethanol was used to facilitate diffusion of the
monomer solution into the inverse opal polystyrene After
the inverse opal was infiltrated by the monomer solution
containing 1 wt % of the initiator AIBN and a subsequent
free radical polymerization at 60°C for 3 h, a solid composite
resulted The initial inverse opal polystyrene template was
then removed with chloroform in a Soxhlet extractor for 12
h, whereupon the opal gel was formed By using different
compositions of the monomer solution, hole sizes, and
stacking structures of the starting inverse opal templates, opal
gels with correspondingly different properties can be duced Water was completely removed from the opalhydrogel by repeatedly rinsing it with a large amount ofethanol Afterward, the opal gel was put into a large amount
pro-of tetrabutyl titanate (TBT) at ambient temperature for 24
h The TBT-swollen opal gel was then immersed in a water/ethanol (1:1 wt/wt) mixture for 5 h to let the TiO2sol/gelprocess proceed Figure 35A shows the opal structure of thegel/titania composite spheres formed After calcination, TiO2opal with distinctive spherical contours could be found The
compression degree, R, was adjusted by the spacer height
when the substrates were compressed When the substrateswere slightly compressed against each other to the extent ofproducing a 20% reduction in the thickness of the composi-tion opal, the deformation of the template-synthesized titaniaspheres was not substantial (Figure 35B) When the com-pression degree was increased to the point of reaching 35%deformation in the opal gel, noticeably deformed titania opalscould be obtained (Figure 35C and D)
2.15 Preparation of TiO2 Nanosheets
The preparation of TiO2nanosheets has also been exploredrecently.359-368Typically, TiO2nanosheets were synthesized
by delaminating layered protonic titanate into colloidal singlelayers A stoichiometric mixture of Cs2CO3and TiO2wascalcined at 800°C for 20 h to produce a precursor, cesiumtitanate, Cs0.7Ti1.82500.175O4 (0: vacancy), about 70 g ofwhich was treated with 2 L of a 1 M HCl solution at roomtemperature This acid leaching was repeated three times byrenewing the acid solution every 24 h The resulting acid-exchanged product was filtered, washed with water, and air-dried The obtained protonic titanate, H0.7Ti1.82500.175O4‚H2O,was shaken vigorously with a 0.017 M tetrabutylammoniumhydroxide solution at ambient temperature for 10 days Thesolution-to-solid ratio was adjusted to 250 cm3 g-1 Thisprocedure yielded a stable colloidal suspension with an
Figure 32 SEM picture of a TiO2skeleton with a cylinder radius
of about 0.06a a is the lattice constant of the cubic unit cell.
Reprinted from Dong, W.; Marlow, F Physica E 2003, 17, 431,
Copyright 2003, with permission from Elsevier
Figure 33. (A) Experimental procedure for fabricating TiO2
nanobowl arrays (B) Low- and high- (inset) magnification SEM
image of TiO2nanobowl arrays Reprinted with permission from
Wang, X D.; Graugnard, E.; King, J S.; Wang, Z L.; Summers,
C J Nano Lett 2004, 4, 2223 Copyright 2004 American Chemical
Society
Figure 34 (A) SEM images of short and densely aligned ZnO
nanorod array on a sapphire substrate Inset: An optical image ofthe aligned ZnO nanorods over a large area (B) SEM image ofthe TiO2-coated ZnO nanorod array Reprinted with permission fromWang, X.; Neff, C.; Graugnard, E.; Ding, Y.; King, J S.; Pranger,
L A.; Tannenbaum, R.; Wang, Z L.; Summers, C J AdV Mater.
2005, 17, 2103 Copyright 2005 Wiley-VCH.
Trang 20opalescent appearance Figure 36 shows TEM and AFM
images of TiO2nanosheets with thicknesses of 1.2-1.3 nm,
which is the height of the TiO2nanosheet with a monolayer
of water molecules on both sides (0.70 + 0.25× 2) thick.366
3 Properties of TiO2 Nanomaterials
3.1 Structural Properties of TiO2 Nanomaterials
Figure 37 shows the unit cell structures of the rutile and
anatase TiO2.11 These two structures can be described in
terms of chains of TiO6octahedra, where each Ti4+ ion is
surrounded by an octahedron of six O2-ions The two crystal
structures differ in the distortion of each octahedron and by
the assembly pattern of the octahedra chains In rutile, the
octahedron shows a slight orthorhombic distortion; in anatase,the octahedron is significantly distorted so that its symmetry
is lower than orthorhombic The Ti-Ti distances in anataseare larger, whereas the Ti-O distances are shorter than those
in rutile In the rutile structure, each octahedron is in contactwith 10 neighbor octahedrons (two sharing edge oxygen pairsand eight sharing corner oxygen atoms), while, in the anatasestructure, each octahedron is in contact with eight neighbors(four sharing an edge and four sharing a corner) Thesedifferences in lattice structures cause different mass densitiesand electronic band structures between the two forms ofTiO2
Hamad et al performed a theoretical calculation on TinO2n
clusters (n ) 1-15) with a combination of simulated
Figure 35 SEM of the TiO2 opals (A) A gel/titania composite opal fabricated without compressing the opal gel template during thesol/gel process (Inset) Image of the sample after calcination at 450°C for 3 h (B-D) (Main panel) Oblate titania opal materials aftercalcination at 450°C for 3 h, subject to compression degree R of (B) 20%, (C) 35%, and (D) 50% The images were taken for the fractured
surfaces containing the direction of applied compression (Inset) Image of the same sample, but with the fracture surface perpendicular to
the direction of applied compression From: Ji, L.; Rong, J.; Yang, Z Chem Commun 2003, 1080 (http://dx.doi.org/10.1039/b300825h)
s Reproduced by permission of The Royal Society of Chemistry
Trang 21annealing, Monte Carlo basin hopping simulation, and
genetic algorithms methods.369They found that the calculated
global minima consisted of compact structures, with titanium
atoms reaching high coordination rapidly as n increased For
n g 11, the particles had at least a central octahedron
surrounded by a shell of surface tetrahedra, trigonal
bipyra-mids, and square base pyramids
Swamy et al found the metastability of anatase as a
function of pressure was size dependent, with smaller
crystallites preserving the structure to higher pressures.370
Three size regimes were recognized for the pressure-induced
phase transition of anatase at room temperature: an
anatase-amorphous transition regime at the smallest crystallite sizes,
an anatase-baddeleyite transition regime at intermediatecrystallite sizes, and an anatase-R-PbO2transition regimecomprising large nanocrystals to macroscopic single crystals.Barnard et al performed a series of theoretical studies onthe phase stability of TiO2nanoparticles in different environ-ments by a thermodynamic model.371-375 They found thatsurface passivation had an important impact on nanocrystalmorphology and phase stability The results showed thatsurface hydrogenation induced significant changes in theshape of rutile nanocrystals, but not in anatase, and that thesize at which the phase transition might be expected increaseddramatically when the undercoordinated surface titaniumatoms were H-terminated For spherical particles, the cross-over point was about 2.6 nm For a clean and faceted surface,
at low temperatures (a phase transition pointed at an averagediameter of approximately 9.3-9.4 nm for anatase nano-crystals), the transition size decreased slightly to 8.9 nm whenthe surface bridging oxygens were H-terminated, and the sizeincreased significantly to 23.1 nm when both the bridgingoxygens and the undercoordinated titanium atoms of thesurface trilayer were H-terminated Below the cross point,the anatase phase was more stable than the rutile phase.371
In their study on TiO2 nanoparticles in vacuum or waterenvironments, they found that the phase transition size inwater (15.1 nm) was larger than that under vacuum (9.6nm).373 In their predictions on the transition enthalpy ofnanocrystalline anatase and rutile, they found that thermo-chemical results could differ for various faceted or spherical
Figure 36 (A) TEM of Ti1-δO2δ-nanosheets (B and C) AFM image and height scan of the TiO2nanosheets deposited on a Si wafer.(D) Structural model for a hydrated TiO2 nanosheet Closed, open, and shaded circles represent Ti atom, O atom, and H2O molecules,respectively All the water sites are assumed to be half occupied Reprinted with permission from Sasaki, T.; Ebina, Y.; Kitami, Y.; Watanabe,
M.; Oikawa, T J Phys Chem B 2001, 105, 6116 Copyright 2001 American Chemical Society.
Figure 37 Lattice structure of rutile and anatase TiO2 Reprinted
with permission from Linsebigler, A L.; Lu, G.; Yates, J T., Jr
Chem ReV 1995, 95, 735 Copyright 1995 American Chemical
Society
Trang 22nanoparticles as a function of shape, size, and degree of
surface passivation.372 Their study on anatase and rutile
titanium dioxide polymorphs passivated with complete
monolayers of adsorbates by varying the hydrogen to oxygen
ratio with respect to a neutral, water-terminated surface
showed that termination with water consistently resulted in
the lowest values of surface free energy when hydrated or
with a higher fraction of H on the surface on both anatase
and rutile surfaces, but conversely, the surfaces generally
had a higher surface free energy when they had an equal
ratio of H and O in the adsorbates or were O-terminated.375
They demonstrated that, under different pH conditions from
acid to basic, the phase transition size of a TiO2nanoparticle
varied from 6.9 to 22.7 nm, accompanied with shape changes
of the TiO2nanoparticles as shown in Figure 38.374
Enyashin and Seifert conducted a theoretical study on the
structural stability of TiO2layer modifications (anatase and
lepidocrocite) using the density-functional-based tight
bind-ing method (DFTB).376They found that anatase nanotubes
were the most stable modifications in a comparison of
single-walled nanotubes, nanostrips, and nanorolls Their stability
increased as their radii grew The energies for all TiO2
nanostructures relative to the infinite monolayer followed a
1/R2curve
Chen et al found that severe distortions existed in Ti site
environments in the structures of 1.9 nm TiO2nanoparticles
compared to those octahedral Ti sites in bulk anatase Ti using
K-edge XANES.377The distorted Ti sites were likely to adopt
a pentacoordinate square pyramidal geometry due to the
truncation of the lattice The distortions in the TiO2lattice
were mainly located on the surface of the nanoparticles and
were responsible for binding with other small molecules
Qian et al found that the density of the surface states on
TiO2nanoparticles was likely dependent upon the details of
the preparation methods.378The TiO2nanoparticles prepared
from basic sol were found to have more surface states than
those prepared from acidic sol based on a surface
photo-voltage spectroscopy study
3.2 Thermodynamic Properties of TiO2
Nanomaterials
Rutile is the stable phase at high temperatures, but anatase
and brookite are common in fine grained (nanoscale) natural
and synthetic samples On heating concomitant with ing, the following transformations are all seen: anatase tobrookite to rutile, brookite to anatase to rutile, anatase torutile, and brookite to rutile These transformation sequencesimply very closely balanced energetics as a function ofparticle size The surface enthalpies of the three polymorphsare sufficiently different that crossover in thermodynamicstability can occur under conditions that preclude coarsening,with anatase and/or brookite stable at small particle size.73,74However, abnormal behaviors and inconsistent results areoccasionally observed
coarsen-Hwu et al found the crystal structure of TiO2nanoparticlesdepended largely on the preparation method.379 For smallTiO2nanoparticles (<50 nm), anatase seemed more stableand transformed to rutile at >973 K Banfield et al foundthat the prepared TiO2 nanoparticles had anatase and/orbrookite structures, which transformed to rutile after reaching
a certain particle size.73,380Once rutile was formed, it grewmuch faster than anatase They found that rutile became morestable than anatase for particle size > 14 nm
Ye et al observed a slow brookite to anatase phasetransition below 1053 K along with grain growth, rapidbrookite to anatase and anatase to rutile transformationsbetween 1053 K and 1123 K, and rapid grain growth of rutileabove 1123 K as the dominant phase.381They concluded thatbrookite could not transform directly to rutile but had totransform to anatase first However, direct transformation
of brookite nanocrystals to rutile was observed above 973
K by Kominami et al.382
In a later study, Zhang and Banfield found that thetransformation sequence and thermodynamic phase stabilitydepended on the initial particle sizes of anatase and brookite
in their study on the phase transformation behavior ofnanocrystalline aggregates during their growth for isothermaland isochronal reactions.74They concluded that, for equallysized nanoparticles, anatase was thermodynamically stablefor sizes < 11 nm, brookite was stable for sizes between 11and 35 nm, and rutile was stable for sizes > 35 nm.Ranade et al investigated the energetics of the TiO2polymorphs (rutile, anatase, and brookite) by high-temper-ature oxide melt drop solution calorimetry, and they foundthe energetic stability crossed over between the three phases
as shown in Figure 39.383The dark solid line represents thephases of lowest enthalpy as a function of surface area Rutilewas energetically stable for surface area < 592 m2/mol (7
m2/g or >200 nm), brookite was energetically stable from
Figure 38 Morphology predicted for anatase (top), with (a)
hydrogenated surfaces, (b) hydrogen-rich surface adsorbates, (c)
hydrated surfaces, (d) hydrogen-poor adsorbates, and (e) oxygenated
surfaces, and for rutile (bottom), with (f) hydrogenated surfaces,
(g) hydrogen-rich surface adsorbates, (h) hydrated surfaces, (i)
hydrogen-poor adsorbates, and (j) oxygenated surfaces Reprinted
with permission from Barnard, A S.; Curtiss, L A Nano Lett.
2005, 5, 1261 Copyright 2005 American Chemical Society.
Figure 39 Enthalpy of nanocrystalline TiO2 Reprinted withpermission from Ranade, M R.; Navrotsky, A.; Zhang, H Z.; Ban-field, J F.; Elder, S H.; Zaban, A.; Borse, P H.; Kulkarni, S K.;
Doran, G S.; Whitfield, H J Proc Natl Acad Sci U.S.A 2002,
99, 6476 Copyright 2002 National Academy of Sciences, U.S.A.
Trang 23592 to 3174 m2/mol (7-40 m2/g or 200-40 nm), and anatase
was energetically stable for greater surface areas or smaller
sizes (<40 nm) The anatase and rutile energetics cross at
1452 m2/mol (18 m2/g or 66 nm) Assuming spherical
particles, the calculated average diameters of rutile and
brookite for a 7 m2/g surface area were 201 and 206 nm,
and those of brookite and anatase for a 40 m2/g surface area
are 36 and 39 nm These differences in particle size at the
same surface area existed because of the differences in
density If the phase transformation took place without further
coarsening, the particle size should be smaller after the
transformation Phase stability in a thermodynamic sense is
governed by the Gibbs free energy (∆G ) ∆H - T∆S) rather
than the enthalpy Rutile and anatase have the same entropy
Thus, the T ∆S will not significantly perturb the sequence of
stability seen from the enthalpies For nanocrystalline TiO2,
if the initially formed brookite had surface area > 40 m2/g,
it was metastable with respect to both anatase and rutile,
and the sequence brookite to anatase to rutile during
coarsening was energetically downhill If anatase formed
initially, it could coarsen and transform first to brookite (at
40 m2/g) and then to rutile The energetic driving force for
the latter reaction (brookite to rutile) was very small,
explaining the natural persistence of coarse brookite In
contrast, the absence of coarse-grained anatase was consistent
with the much larger driving force for its transformation to
rutile.383
Li et al found that only anatase to rutile phase
transforma-tion occurred in the temperature range of 973-1073 K.384
Both anatase and rutile particle sizes increased with the
increase of temperature, but the growth rate was different,
as shown in Figure 40 Rutile had a much higher growth
rate than anatase The growth rate of anatase leveled off at
800 °C Rutile particles, after nucleation, grew rapidly,
whereas anatase particle size remained practically unchanged
With the decrease of initial particle size, the onset transition
temperature was decreased An increased lattice compression
of anatase with increasing temperature was observed Larger
distortions existed in samples with smaller particle size The
values for the activation energies obtained were 299, 236,
and 180 kJ/mol for 23, 17, and 12 nm TiO2nanoparticles,
respectively The decreased thermal stability in finer
nano-particles was primarily due to the reduced activation energy
as the size-related surface enthalpy and stress energy
increased
3.3 X-ray Diffraction Properties of TiO2
Nanomaterials
XRD is essential in the determination of the crystal
structure and the crystallinity, and in the estimate of the
crystal grain size according to the Scherrer equation
where K is a dimensionless constant, 2 θ is the diffraction
angle,λ is the wavelength of the X-ray radiation, and β is
the full width at half-maximum (fwhm) of the diffraction
peak.385 Crystallite size is determined by measuring the
broadening of a particular peak in a diffraction pattern
associated with a particular planar reflection from within the
crystal unit cell It is inversely related to the fwhm of an
individual peaksthe narrower the peak, the larger the
crystallite size The periodicity of the individual crystallite
domains reinforces the diffraction of the X-ray beam,resulting in a tall narrow peak If the crystals are randomlyarranged or have low degrees of periodicity, the result is abroader peak This is normally the case for nanomaterialassemblies Thus, it is apparent that the fwhm of thediffraction peak is related to the size of the nanomaterials.Figure 41 shows the XRD patterns for TiO2nanoparticles
of different sizes111 and for TiO2 nanorods of differentlengths.129As the nanoparticle size increased, the diffractionpeaks became narrower In the anatase nanoparticle andnanorods developed by Zhang et al., the diameters of theTiO2nanoparticles and nanorods were both around 2.3 nm.The nanorods were elongated along the [001] direction with
preferred anisotropic growth along the c-axis of the anatase
lattice, which was indicated by the strong peak intensity andnarrow width of the (004) reflection and relatively lowerintensity and broader width for the other reflections With
an increase in length of the nanorods, the (004) diffractionpeak became much stronger and sharper, whereas other peaksremained similar in shape and intensity.129Similar resultshave been observed by other groups.123,127,177,183
3.4 Raman Vibration Properties of TiO2Nanomaterials
As the size of TiO2nanomaterials decreases, the featuredRaman scattering peaks become broader.255,318,370,386-395Thesize effect on the Raman scattering in nanocrystallineTiO is interpreted as originating from phonon confine-
Figure 40 (A) Changes in particle sizes of anatase and rutile
phases as a function of the annealing temperatures (B) Arrenhius
plot of ln(AR/A0) vs 1/T for activation energy calculations as a
function of the size of the TiO2nanoparticles ARand A0 are theintegrated diffraction peak intensity from rutile (110), and the totalintegrated anatase (101) and rutile (110) peak intensity, respectively.Reused with permission from W Li, C Ni, H Lin, C P Huang,
and S Ismat Shah, Journal of Applied Physics, 96, 6663 (2004).
Copyright 2004, American Institute of Physics
Trang 24ment,255,318,370,386,387,395 nonstoichiometry,391,392 or internal
stress/surface tension effects.390Among these theories, the
most convincing is the three-dimensional confinement of
phonons in nanocrystals.255,318,370,386,387,394,395 The phonon
confinement model is also referred to as the spatial
correla-tion model or q vector relaxacorrela-tion model It links the q vector
The anatase TiO2has six Raman-active fundamentals inthe vibrational spectrum: three Egmodes centered around
144, 197, and 639 cm-1(designated here Eg(1), Eg(2), and Eg(3),respectively), two B1gmodes at 399 and 519 cm-1 (desig-nated B1g(1) and B1g(2d)), and an A1gmode at 513 cm-1.370
As the particle size decreases, the Raman peaks showincreased broadening and systematic frequency shifts (Figure42).370The most intense Eg(1)mode shows the maximum blueshift and significant broadening with decreasing crystallitesize A small blue shift is seen for the Eg(2)mode, while the
B1g(1)mode and the B1g(2)+A1gmodes show very small blueshifts and red shifts (the latter peak represents a combinedeffect of two individual modes), respectively Whereas thefrequency shifts for the A1g and B1g modes are not pro-nounced, increased broadening with decreasing crystallitesize is clearly seen for these modes The Eg(3)mode showssignificant broadening and a red shift with decreasingcrystallite size
Choi et al found a volume contraction effect in anataseTiO2 nanoparticles due to increasing radial pressure asparticle size decreases, and they suggested that the effects
of decreasing particle size on the force constants andvibrational amplitudes of the nearest neighbor bonds con-tributed to both broadening and shifts of the Raman bandswith decreasing particle diameter.388
3.5 Electronic Properties of TiO2 Nanomaterials
The DOS of TiO2is composed of Ti eg, Ti t2g(dyz, dzx,and dxy), O pσ(in the Ti3O cluster plane), and O pπ(out ofthe Ti3O cluster plane), as shown in Figure 43A.396The uppervalence bands can be decomposed into three main regions:theσ bonding in the lower energy region mainly due to O
pσbonding; theπ bonding in the middle energy region; and
O pπ states in the higher energy region due to O pπnonbonding states at the top of the valence bands where thehybridization with d states is almost negligible The contri-bution of theπ bonding is much weaker than that of the σ
bonding The conduction bands are decomposed into Ti eg(>5 eV) and t2gbands (<5 eV) The dxystates are dominantlylocated at the bottom of the conduction bands (the verticaldashed line in Figure 43A) The rest of the t2g bands areantibonding with p states The main peak of the t2gbands isidentified to be mostly dyzand dzx states
In the molecular-orbital bonding diagram in Figure 43B,
a noticeable feature can be found in the nonbonding statesnear the band gap: the nonbonding O pp orbital at the top
of the valence bands and the nonbonding dxystates at thebottom of the conduction bands A similar feature can beseen in rutile; however, it is less significant than in anatase.397
In rutile, each octahedron shares corners with eight neighborsand shares edges with two other neighbors, forming a linearchain In anatase, each octahedron shares corners with four
Figure 41 (A) Powder XRD patterns of TiO2samples of different
diameters: (a) 5 nm; (b) 7 nm; (c) 13 nm Reprinted with permission
from Niederberger, M.; Bartl, M H.; Stucky, G D Chem Mater.
2002, 14, 4364 Copyright 2002 American Chemical Society (B)
Powder XRD patterns of TiO2samples of diameter 2.3 nm: (a)
spherical particles; (b) 16-nm nanorods; (c) 30-nm nanorods
Reprinted with permission from Zhang, Z.; Zhong, X.; Liu, S.; Li,
D.; Han, M Angew Chem., Int Ed 2005, 44, 3466 Copyright
2005 Wiley-VCH
Trang 25neighbors and shares edges with four other neighbors,
forming a zigzag chain with a screw axis Thus, anatase is
less dense than rutile Also, anatase has a large metal-metal
distance of 5.35 Å As a consequence, the Ti dxyorbitals at
the bottom of the conduction band are quite isolated, while
the t2gorbitals at the bottom of the conduction band in rutile
provide the metal-metal interaction with a smaller distance
of 2.96 Å
The electronic structure of TiO2 has been studied with
various experimental techniques, i.e., with X-ray
photoelec-tron and X-ray absorption and emission
spectroscop-ies.379,398-405 Figure 44 shows a schematic energy level
diagram of the lowest unoccupied MOs of a [TiO6]8-cluster
with O h , D 2h (rutile), and D 2d(anatase) symmetry and the Ti
K-edge XANES and O K-edge ELNES spectra for rutile and
anatase.398The anatase structure is a tetragonally distorted
octahedral structure in which every titanium cation is
surrounded by six oxygen atoms in an elongated octahedral
geometry (D 2d) The further splitting of the 3d levels of Ti3+due to the asymmetric crystals is shown for rutile and anatasestructures The fine electronic structure of TiO2 can bedirectly probed by Ti K-edge X-ray-absorption near-edgestructure (XANES), and the right panel of Figure 44Bcontains O K-edge experimental electron-energy-loss near-edge structure (ELNES) spectra.398
Hwu et al found that the crystal field splitting ofnanocrystal TiO2was approximately 2.1 eV, slightly smallerthan that of bulk TiO2, as shown in Figure 45A.379Luca et
al found that 1s f np transitions broadened as particle size(increased or decreased) in the postedge region in the X-rayabsorption spectroscopy for TiO2 nanoparticles.403 Also, aclear trend in the X-ray absorption spectroscopy for differentsized TiO2nanoparticles was observed, as shown in Figure45B from the study by Choi et al.401
Figure 42 (A) Ambient pressure Raman spectra of anatase with an average crystallite size of 4 ( 1 nm (A), 8 ( 2 nm (B), 20 ( 8 nm
(C), and 34 ( 5 nm (D) The spectrum marked “E” is from a bulk anatase (B) The Raman line width (fwhm) of the Eg(1)mode versuscrystallite size Reprinted with permission from Swamy, V.; Kuznetsov, A.; Dubrovinsky, L S.; Caruso, R A.; Shchukin, D G.; Muddle,
B C Phys ReV B 2005, 71, 184302/1 (http://link.aps.org/abstract/PRB/v71/p184302) Copyright 2005 by the American Physical Society.
Trang 26It is well-known that for nanoparticles the band gap energy
increases and the energy band becomes more discrete with
decreasing size.84,406,407 As the size of a semiconductor
band gap blue shift (<0.1-0.2 eV) caused by quantum sizeeffects for spherical particles sizes down to 2 nm.58,60Suchsmall effects are mainly due to the relatively high effectivemass of carriers in TiO2 and an exciton radius in theapproximate range 0.75-1.90 nm.84 On the other hand,Serpone et al suggested that the blue shifts in the effectiveband gap of TiO2with particle sizes of 21, 133, and 267 Åmay in fact not be a quantum confinement effect.410Mon-ticone et al did an excellent study on the quatum size effects
in anatase nanoparticles and found no quantum size effect
in anatase TiO2nanoparticles for sizes 2R g 1.5 nm, but
they did find unusual variation of the oscillator strength ofthe first allowed direct transition with particle size.411
3.6 Optical Properties of TiO2 Nanomaterials
The main mechanism of light absorption in pure conductors is direct interband electron transitions Thisabsorption is especially small in indirect semiconductors, e.g.,TiO2, where the direct electron transitions between the bandcenters are prohibited by the crystal symmetry Braginskyand Shklover have shown the enhancement of light absorp-tion in small TiO2 crystallites due to indirect electrontransitions with momentum nonconservation at the inter-face.412This effect increases at a rough interface when theshare of the interface atoms is larger The indirect transitionsare allowed due to a large dipole matrix element and a largedensity of states for the electron in the valence band.Considerable enhancement of the absorption is expected insmall TiO2 nanocrystals, as well as in porous and micro-crystalline semiconductors, when the share of the interfaceatoms is sufficiently large A rapid increase in the absorption
semi-takes place at low (h ν < Eg+ Wc, where Wcis the width ofthe conduction band) photon energies Electron transitions
to any point in the conduction band become possible when
h ν ) Eg+ Wc Further enhancement of the absorption occursdue to an increase of the electron density of states in onlythe valence band The interface absorption becomes the mainmechanism of light absorption for the crystallites that aresmaller than 20 nm.412
Sato and Sakai et al showed through calculation andmeasurement that the band gap of TiO2nanosheets was largerthan the band gap of bulk TiO2, due to lower dimensionality,i.e., a 3D to 2D transition, as shown in Figure 46.360,413Fromthe measurement, it was found that the lower edge of theconduction band for the TiO2nanosheet was approximately0.1 V higher, while the upper edge of the valence band was0.5 V lower than that of anatase TiO2.360The absorption ofthe TiO2nanosheet colloid blue shifted (>1.4 eV) relative
to that of bulk TiO2 crystals (3.0-3.2 eV), due to a quantization effect, accompanied with a strong photolumi-nescence of well-developed fine structures extending intothe visible light regime.362,363The band gap energy shift,∆E,
size-Figure 43 (A) Total and projected densities of states (DOSs) of
the anatase TiO2structure The DOS is decomposed into Ti eg, Ti
t2g(dyz, dzx, and dxy), O pσ(in the Ti3O cluster plane), and O pπ
(out of the Ti3O cluster plane) components The top of the valence
band (the vertical solid line) is taken as the zero of energy The
vertical dashed line indicates the conduction-band minimum as a
guide to the eye (B) Molecular-orbital bonding structure for anatase
TiO2: (a) atomic levels; (b) crystal-field split levels; (c) final
interaction states The thin-solid and dashed lines represent large
and small contributions, respectively Reprinted with permission
from Asahi, R.; Taga, Y.; Mannstadt, W.; Freeman, A J Phys.
ReV B 2000, 61, 7459 (http://link.aps.org/abstract/PRB/v61/p7459).
Copyright 2000 by the American Physical Society
Trang 27by exciton confinement in anisotropic two-dimensional
crystallites is formulated as follows:
where h is Plank’s constant, µ xz and µ y are the reduced
effective masses of the excitons, and L , L , and L are the
crystallite dimensions in the parallel and perpendiculardirections with respect to the sheet, respectively Since thefirst term can be ignored, the blue shift is predominantlygoverned by the sheet thickness The onset of a 270 nm peak
in the photoluminescence of TiO2nanosheets was assigned
to resonant luminescence The series of peaks extending into
a longer wavelength region were attributed to interband levelsgenerated by the intrinsic Ti site vacancies The contrasting
Figure 44 (A) Schematic energy level diagram of the lowest unoccupied MOs of a [TiO6]8-cluster with O h , D 2h (rutile), and D 2d(anatase)symmetry (B) Ti K-edge XANES and O K-edge ELNES spectra for rutile (a) and anatase (b) Reprinted with permission from Wu, Z Y.;
Ouvrared, G.; Gressier, P.; Natoli, C R Phys ReV B 1997, 55, 10382 (http://link.aps.org/abstract/PRB/v55/p10382) Copyright 1997 by
the American Physical Society
Trang 28Within the effective mass model, the energy spectrum of
2D TiO2nanosheets can be described by eq 10, where the
“plus” and “minus” signs correspond to the conduction and
valence bands, respectively, EG is the energy gap, p is
Planck’s constant, and meand mh are the effective masses
of the electrons and holes, respectively
The electronic band structure of a TiO2 nanotube can be
obtained from this relation by zone-folding and is given by
a series of quasi-1D sub-bands with different indices n
(Figure 48b):
This transition from the 2D to the quasi-1D energy spectrum
has a dramatic effect on the energy density of states In the
2D case, the density of states, G2D) mc.h/πp2, has a constant
value for energies outside the energy gap (see Figure 48c)
In the quasi-1D case, however, the density of states of each
sub-band
diverges at the band edge E n(0), leading to van Hove
singularities The resulting density of state is formed by a
series of sharp peaks with long overlapping tails (Figure 48c)
The energy gap between the valence and conductance bands
in the quasi-1D case is larger than that in the parental 2D
material, and the difference increases with decreasing
diameter of the nanotube The change in the energy gaps
between a nanosheet and a nanotube is
In TiO2, the effective masses of electrons me can vary
between 5m0 and 30m0, and the mass of holes mhis more
than 3m0 With me ) 9m0 and mh ) 3m0, the difference
between energy gaps of nanotubes with diameters 2.5 and 5
nm is 8 meV The energy difference between the two first
peaks in the density of states G1D(E) (Figure 48) is less than
24 meV for d ) 2.5 nm and 6 meV for d ) 5 nm, which are
too small to be resolved in room-temperature experiments
due to the thermal fluctuations of kT ) 26 meV.158
In the theoretical study conducted by Enyashin and Seifert
recently, the band structures for anatase nanotubes,
nano-strips, and nanorolls were similar to the DOS of the
corresponding bulk phase.376The valence band of both bulkTiO2and their nanostructures was composed of 3d Ti-2p
O states, and the lower part of the conduction band wasformed by 3d Ti states The differences between thesenanostructures were insignificant All anatase systems weresemiconductors with a wide direct band gap (∼4.2 eV), while
the lepidocrocite nanotubes were semiconductors with anindirect band gap (∼4.5 eV) Independent from the specific
topology of the titania nanostructures, the band gap proached the band gap of the corresponding nanocrystals withradii of about 25 Å.376
ap-In addition to the above investigation on the bulk electronicstructures for various TiO2 nanomaterials, Mora-Sero´ andBisquert investigated the Fermi level of surface states in TiO2nanoparticles by the nonequilibrium steady-state statistics ofelectrons.414They found that the electrons trapped in surfacestates did not generally equilibrate to the free electrons’ Fermi
level, EFn, and a distinct Fermi level for surface states, EFs,could be defined consistent with Fermi-Dirac statistics,determining the surface states’ occupancy far from equilib-rium The difference between the free electrons’ Fermi level
Lin, H M Nanostruct Mater 1997, 9, 355, Copyright 1997, with
permission from Elsevier (B) Ti L 2.3 absorption of TiO2 crystals with different sizes Reprinted with permission from Choi,
nano-H C.; Ahn, nano-H J.; Jung, Y M.; Lee, M K.; Shin, nano-H J.; Kim, S
B.; Sung, Y E Appl Spectrosc 2004, 58, 598 Copyright 2004
Society for Applied Spectroscopy
Trang 29and the surface Fermi level (∆EFn - EFs) was found to
depend on the rate constants for charge transfer and
detrap-ping and could reach several hundred
millielectron-volts.414
3.7 Photon-Induced Electron and Hole Properties
of TiO2 Nanomaterials
After TiO2nanoparticles absorb, impinging photons with
energies equal to or higher than its band gap (>3.0 eV),
electrons are excited from the valence band into the
unoc-cupied conduction band, leading to excited electrons in the
conduction band and positive holes in the valence band
These charge carriers can recombine, nonradiatively or
radiatively (dissipating the input energy as heat), or get
trapped and react with electron donors or acceptors adsorbed
on the surface of the photocatalyst The competition between
these processes determines the overall efficiency for various
applications of TiO2 nanoparticles These fundamentalprocesses can be expressed as follows:415
Figure 46 (A) Total and partial densities of states for (a) stacked TiO2 sheets, (b) a single-layered TiO2, (c) rutile, and (d) anatase
Reprinted with permission from Sato, H.; Ono, K.; Sasaki, T.; Yamagishi, A J Phys Chem B 2003, 107, 9824 Copyright 2003 American
Chemical Society (B) Schematic illustration of electronic band structure: (a) TiO2 nanosheets; (b) anatase Reprinted with permission
from Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T J Am Chem Soc 2004, 126, 5851 Copyright 2004 American Chemical Society (C)
UV-visible spectra of (a) TiO2 sheets and (b) a film of nanosheets on a SiO2 glass substrate The data for the colloidal suspension is
denoted by a dashed trace Reprinted with permission from Sasaki, T.; Watanabe, M J Phys Chem B 1997, 101, 10159 Copyright 1997
American Chemical Society
2-4O2(g) + vacancy (12)
Trang 30OH radicals and O vacancies, respectively The reverse of
reaction 12 generates O adatom intermediates upon exposing
defective surfaces to O2-(g).415Electrons and holes generated
in TiO2nanoparticles are localized at different defect sites
on the surface and in the bulk Electron paramagnetic
resonance (EPR) results showed that electrons were trapped
as two Ti(III) centers, while the holes were trapped as
oxygen-centered radicals covalently linked to surface
tita-nium atoms.416-419Howe and Gra¨tzel found that irradiation
at 4.2 K in vacuo produced electrons trapped at Ti4+sites
within the bulk and holes trapped at lattice oxide ions
immediately below the surface, which decayed rapidly in
the dark at 4.2 K In the presence of O2, trapped electrons
were removed and the trapped holes were stable to 77 K
Warming to room-temperature caused loss of trapped holes
and formation of O2-at the surface.416,417Hurum et al found
that, upon band gap illumination, holes appeared at the
surface and preferentially recombined with electrons in
surface trapping sites for mixed-phase TiO2, such as Degussa
P25, and recombination reactions were dominated by surface
reactions that followed charge migration.419
Colombo and Bowman studied the charge carrier dynamics
of TiO2nanoparticles with femtosecond time-resolved diffuse
reflectance spectroscopy and found a dramatic increase in
the population of trapped charge carriers within the first few
picoseconds.420,421Skinner et al found that the trapping time
for photogenerated electrons on 2 nm TiO2nanoparticles in
acetonitrile by ultrafast transient absorption was about 180
fs.422Serpone et al found that localization (trapping) of the
electron as a Ti3+species occurred with a time scale of about
30 ps and about 90% or more of the photogenerated electron/
hole pairs recombined within 10 ns.409They suggested that
photoredox chemistry occurring at the particle surface
emanated from trapped electrons and trapped holes rather
than from free valence band holes and conduction band
electrons Bahnemann et al found that, in 2.4 nm TiO2
nanoparticles, electrons were instantaneously trapped within
the duration of the laser flash (20 ns) Deeply trapped holes
were rather long-lived and unreactive, and shallowly trapped
holes were in a thermally activated equilibrium with free
holes which exhibited a very high oxidation potential.423
Szczepankiewicz and Hoffmann et al found that O2was
an efficient scavenger of conduction band electrons at the
gas/solid interface and the buildup of trapped carriers
eventually resulted in extended surface reconstruction
in-volving Ti-OH functionalities.415 They found that
photo-generated free conduction band electrons were coupled with
acoustic phonons in the lattice and their lifetimes were
lengthened when dehydrated.424 The photoexcited charge
carriers in TiO2nanoparticles produced Stark effect intensity
and wavelength shifts for surface TiO-H stretching
vibra-tions Although deep electron-trapping states affected certain
types of TiO-H stretch, shallow electron-trapping statesproduced a homogeneous electric field and were suggestednot to be associated with localized structures, but ratherdelocalized across the TiO2surface.424
Berger et al studied UV light-induced electron-hole pairexcitations in anatase TiO2nanoparticles by electron para-magnetic resonance (EPR) and IR spectroscopy.425 Thelocalized states such as holes trapped at oxygen anions (O-)
Figure 47 (A) (a) Absorption spectrum and (b) luminescence
excitation spectrum (wavelength of emission light is 400 nm) ofcolloidal TiO2nanotubes of different mean diameters: (1) 2.5 nm;(2) 3.1 nm; (3) 3.5 nm; (4) 5 nm The curves are shifted verticallyfor clarity (B) Photoluminescence spectra of colloidal TiO2
nanotubes of different mean diameters: (1) 2.5 nm; (2) 3.1 nm;(3) 3.5 nm; (4) 5 nm Room temperature, excitation wavelength
237 nm, slits width 5 nm The range of wavelengths, 455-490
nm, in the spectra is omitted due to the high signal of the secondharmonic from scattered excitation light The curves are shiftedvertically for clarity Vertical lines (5) show the positions of thepeaks in the PL spectrum of the nanosheets Reprinted withpermission from Bavykin, D V.; Gordeev, S N.; Moskalenko, A
V.; Lapkin, A A.; Walsh, F C J Phys Chem B 2005, 109, 8565.
Copyright 2005 American Chemical Society
Trang 31and electrons trapped at coordinatively unsaturated cations
(Ti3+ formation) were accessible to EPR spectroscopy
Delocalized and EPR silent electrons in the conduction band
may be traced by their IR absorption, which results from
their electronic excitation within the conduction band in the
infrared region (Figure 49) They found that, during
continu-ous UV irradiation, photogenerated electrons were either
trapped at localized sites, giving paramagnetic Ti3+centers,
or remained in the conduction band as EPR silent specieswhich may be observed by their IR absorption and that theEPR-detected holes produced by photoexcitation were O-species, produced from lattice O2- ions It was also foundthat, under high-vacuum conditions, the majority of photo-excited electrons remained in the conduction band At 298
K, all stable hole and electron states were lost
4 Modifications of TiO2 Nanomaterials
Many applications of TiO2 nanomaterials are closelyrelated to its optical properties However, the highly efficientuse of TiO2nanomaterials is sometimes prevented by its wideband gap The band gap of bulk TiO2lies in the UV regime(3.0 eV for the rutile phase and 3.2 eV for the anatase phase),which is only a small fraction of the sun’s energy (<10%),
as shown in Figure 50.11Thus, one of the goals for improvement of the performance
of TiO2nanomaterials is to increase their optical activity byshifting the onset of the response from the UV to the visibleregion.21,426-428There are several ways to achieve this goal.First, doping TiO2 nanomaterials with other elements cannarrow the electronic properties and, thus, alter the opticalproperties of TiO2nanomaterials Second, sensitizing TiO2with other colorful inorganic or organic compounds canimprove its optical activity in the visible light region Third,coupling collective oscillations of the electrons in theconduction band of metal nanoparticle surfaces to those inthe conduction band of TiO2nanomaterials in metal-TiO2nanocomposites can improve the performance In addition,the modification of the TiO2nanomaterials surface with othersemiconductors can alter the charge-transfer propertiesbetween TiO and the surrounding environment, thus im-
Figure 48 Schematic presentation of the transformation of the electron band structure of the nanosheet semiconductor accompanying the
formation of nanotubes: (a) band diagram of a 2-dimensional nanosheet; (b) band diagram of quasi-1-D nanotubes; (c) energy density ofstates for nanosheets (G2D) and nanotubes (G1D) EG 1Dand EG 2Dare the band gaps of the 1D and 2D structures, respectively k x and k yare
the wave vectors Reprinted with permission from Bavykin, D V.; Gordeev, S N.; Moskalenko, A V.; Lapkin, A A.; Walsh, F C J Phys.
Chem B 2005, 109, 8565 Copyright 2005 American Chemical Society.
Figure 49 Scheme of UV-induced charge separation in TiO2
Electrons from the valence band can either be trapped (a) by defect
states, which are located close to the conduction band (shallow
traps), or (b) in the conduction band, where they produce absorption
in the IR region Electron paramagnetic resonance spectroscopy
detects both electrons in shallow traps, Ti3+, and hole centers, O-
Reprinted with permission from Berger, T.; Sterrer, M.; Diwald,
O.; Knoezinger, E.; Panayotov, D.; Thompson, T L.; Yates, J T.,
Jr J Phys Chem B 2005, 109, 6061 Copyright 2005 American
Chemical Society
Trang 32proving the performance of TiO2 nanomaterials-based
de-vices
4.1 Bulk Chemical Modification: Doping
The optical response of any material is largely determined
by its underlying electronic structure The electronic
proper-ties of a material are closely related to its chemical
composition (chemical nature of the bonds between the atoms
or ions), its atomic arrangement, and its physical dimension
(confinement of carriers) for nanometer-sized materials The
chemical composition of TiO2 can be altered by doping
Specifically, the metal (titanium) or the nonmetal (oxygen)
component can be replaced in order to alter the material’s
optical properties It is desirable to maintain the integrity of
the crystal structure of the photocatalytic host material and
to produce favorable changes in electronic structure It
appears easier to substitute the Ti4+cation in TiO2with other
transition metals, and it is more difficult to replace the O
2-anion with other 2-anions due to differences in charge states
and ionic radii The small size of the nanoparticle is beneficial
for the modification of the chemical composition of TiO2
due to the higher tolerance of the structural distortion than
that of bulk materials induced by the inherent lattice strain
in nanomaterials.426,429
4.1.1.1 Metal-Doped TiO2 Nanomaterials Different
metals have been doped into TiO2 nanomaterials.313,430-465
The preparation methods of non-metal-doped TiO2
nanoma-terials can be divided into three types: wet chemistry,
high-temperature treatment, and ion implantation on TiO2
nano-materials Wet chemistry methods usually involve hydrolysis
of a titanium precursor in a mixture of water and other
reagents, followed by heating Choi et al performed a
systematic study of TiO2nanoparticles doped with 21 metal
ions by the sol-gel method and found the presence of metal
ion dopants significantly influenced the photoreactivity,
charge carrier recombination rates, and interfacial
electron-transfer rates.434Li et al developed La3+-doped TiO2by the
sol-gel process and found that the lanthanum doping could
inhibit the phase transformation of TiO2, enhance the thermal
stability of the TiO2, reduce the crystallite size, and increase
the Ti3+content on the surface.442Nagaveni et al prepared
W, V, Ce, Zr, Fe, and Cu ion-doped anatase TiO2
nanopar-ticles by a solution combustion method and found that the
solid solution formation was limited to a narrow range of
concentrations of the dopant ions.448Wang et al prepared
largely dependent on both the nature and the concentration
of the alkaline, with the best crystallinity obtained for doped TiO2and the lowest for K-doped TiO2.430Cao et al.prepared Sn4+-doped TiO2nanoparticle films by the plasma-enhanced CVD method and found that, after doping by Sn,more surface defects were present on the surface.433Gracia
Li-et al synthesized M (Cr, V, Fe, Co)-doped TiO2by ion beaminduced CVD and found that TiO2 crystallized into theanatase or rutile structures depending on the type and amount
of cations present with partial segregation of the cations inthe form of M2Onafter annealing.438Wang et al synthesizedFe(III)-doped TiO2nanoparticles using oxidative pyrolysis
of liquid-feed organometallic precursors in a frequency (RF) thermal plasma and found that the formation
radiation-of rutile was strongly promoted with iron doping compared
to the anatase phase being prevalent in the undoped TiO2.246
4.1.1.2 Nonmetal-Doped TiO2 Nanomaterials Various
nonmetal elements, such as B, C, N, F, S, Cl, and Br, havebeen successfully doped into TiO2nanomaterials C-dopedTiO2nanomateirals have been obtained by heating titaniumcarbide472-474or by annealing TiO2 under CO gas flow athigh temperatures (500-800°C)475or by direct burning of
a titanium metal sheet in a natural gas flame.476N-doped TiO2 nanomaterials have been synthesized byhydrolysis of TTIP in a water/amine mixture and the post-treatment of the TiO2sol with amines426,428,477-482or directlyfrom a Ti-bipyridine complex483or by ball milling of TiO2
in a NH3water solution.484N-doped TiO2nanomaterials werealso obtained by heating TiO2under NH3flux at 500-600
°C485,486 or by calcination of the hydrolysis product ofTi(SO4)2with ammonia as precipitator487or by decomposition
of gas-phase TiCl4 with an atmosphere microwave plasmatorch488 or by sputtering/ion-implanting techniques withnitrogen489,490or N2+gas flux.491
S-doped TiO2nanomaterials were synthesized by mixingTTIP with ethanol containing thiourea492-494or by heatingsulfide powder495,496or by using sputtering or ion-implantingtechniques with S+ion flux.497-499Different doping methodscan induce the different valence states of the dopants Forexample, the incorporated S from thiourea had S4+ or S6+state,492-494while direct heating of TiS2or sputtering with
S+induced the S2-anion.496-499F-doped TiO2nanomaterials were synthesized by mixingTTIP with ethanol containing H2O-NH4F,500-502 or byheating TiO2 under hydrogen fluoride503,504 or by spraypyrolysis from an aqueous solution of H2TiF6505,506or usingion-implanting techniques with F+ion flux.507Cl-and Br-co-doped nanomaterials were synthesized by adding TiCl4
to ethanol containing HBr.508
4.1.2.1 Electronic Properties of Doped TiO2
Nanoma-terials 4.1.2.1.1 Metal-Doped TiO Nanomaterials
Ac-Reprinted with permission from Linsebigler, A L.; Lu, G.; Yates,
J T., Jr Chem ReV 1995, 95, 735 Copyright 1995 American
Chemical Society
Trang 33cording to Soratin and Schwarz’s study, the electronic states
of TiO2can be decomposed into three parts: theσ bonding
of the O pσand Ti egstates in the lower energy region; the
π bonding of the O p πand Ti egstates in the middle energy
region; and the O pπ states in the higher energy region
(Figure 51A).397,509The bottom of the lower conduction band
(CB) consisting of the Ti dxyorbital contributes to the
metal-metal interactions due to theσ bonding of the Ti t2g-Ti t2g
states At the top of the lower CB, the rest of the Ti2gstates
are antibonding with the O pπstates The upper CB consists
of theσ antibonding orbitals between the O p σ and Ti eg
states
The electronic structures, i.e., the densities of states
(DOSs), of V-, Cr-, Mn-, Fe-, Co-, and Ni-doped TiO2were
analyzed by ab initio band calculations based on the density
functional theory with the full-potential linearized augmented
plane wave (FLAPW) method by Umebayashi et al (Figure
51B).509They found that when TiO2was doped with V, Cr,
Mn, Fe, or Co, an electron occupied level formed and the
electrons were localized around each dopant As the atomic
number of the dopant increased, the localized level shifted
to lower energy The energy of the localized level due to
Co doping was low enough that it lay at the top of the valence
band while the other metals produced midgap states The
electrons from the Ni dopant were somewhat delocalized,
thus significantly contributing to the formation of the valence
band with the O p and Ti 3d electrons The states due to the
3d dopants shifted to a lower energy as the atomic number
of the dopant increased For Ti1-xVxO2: two localized levels
occurred at 1.5 eV above the VB (a) and between the lower
and upper CBs (b) Level a was occupied by one electron
consisting of the V t2g and O pπ states and was localized
around V Level b consisted of the V eg and O pσ states
forming the σ antibonding orbital For Cr- and Mn-doped
TiO2, state c was localized at 1.0 eV (0.5 eV for Mn) above
the VB due to Cr (Mn) t2gand O pπ, the former of which
was occupied by 2 (3) electrons Theσ antibonding orbital
formed by the Cr (Mn) egand O pσstates occurred withinthe lower CB For Fe- and Co-doped TiO2, the localized level(e) was situated 0.2 eV above the VB (or at the top of the
VB for Co) due to theπ antibonding of the Fe egand O pπ
states This level was occupied by four (or five for Co)electrons The Fe (Co) egstate was split into d z2(f) and d x2-y2
(g) orbitals in the band gap For Ni-doped TiO2, the π
antibonding of the Ni t2g and O pπ states was somewhatdelocalized and appeared within the VB (h) due to the Ni eg
states from the d z2and d x2-y2orbitials situated in the bandgap The electron densities around the dopant were large inthe VB and small in the CB compared to the case of pureTiO2 The metal-O interaction strengthened, and the metal-metal interaction became weak as a result of the 3d metaldoping
Li et al found that 1.5 at % Nd3+-doped TiO2nanoparticlesreduced the band gap by as much as 0.55 eV and that theband gap narrowing was primarily attributed to the substi-tutional Nd3+ions, which introduced electron states into theband gap of TiO2 to form the new lowest unoccupiedmolecular orbital (LUMO).444Wang and Doren found that
Nd 4f electrons changed the electronic structure of Nd-dopedTiO2into the half-metallic or the insulating ground state510and that V 3d states were located at the bottom of theconduction band of the TiO2host in V-doped TiO2, whichwas shown to be a half-metal or an insulator from theirtheoretical studies.511
4.1.2.1.2 Nonmetal-Doped TiO 2 Nanomaterials Recent
theoretical and experimental studies have shown that thedesired band gap narrowing of TiO2can also be achieved
by using nonmetal dopants (refs 385, 428, 444, 489, 481,
482, 484, 503, 504, and 512-547) Asahi and co-workerscalculated the electronic band structures of anatase TiO2withdifferent substitutional dopants, including C, N, F, P, or S,using the FLAPW method in the framework of the localdensity approximation (LDA) as shown in Figure 52.489Inthis study, C dopant introduced deep states in the gap.489
Figure 51 (A) Bonding diagram of TiO2 (B) DOS of the metal-doped TiO2(Ti1-xAxO2: A ) V, Cr, Mn, Fe, Co, or Ni) Gray solid lines:total DOS Black solid lines: dopant’s DOS The states are labeled (a) to (j) Reprinted from Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai,
K J Phys Chem Solids 2002, 63, 1909, Copyright 2002, with permission from Elsevier.
Trang 34Nakano et al found three deep levels located at
ap-proximately 0.86, 1.30, and 2.34 eV below the conduction
band in C-doped TiO2, which were attributed to the intrinsic
nature of TiO2for the first one and the two levels newly
introduced by the C doping.522In particular, the pronounced
2.34 eV band contributed to band gap narrowing by mixing
with the O 2p valence band.522 Lee et al., in their
first-principles density-functional LDA pseudopotential
calcula-and rutile polymorphs, N 2p localized states were just above the top of the O 2p valence band.512,513 In anatase, thesedopant states caused a red shift of the absorption band edgetoward the visible region, while, in rutile, an overall blueshift was found by the N-induced contraction of the O 2pband.512Experimental evidence supported the statement thatnitrogen-doped TiO2formed nitrogen-induced midgap levelsslightly above the oxygen 2p valence band.486Lee et al., intheir first-principles density-functional LDA pseudopotentialcalculations of electronic properties of N-doped TiO2, foundthat the bands originating from N 2p states appeared in theband gap of TiO2; however, the mixing of N with O 2p stateswas too weak to produce a significant band gap narrowing.517Wang and Doren found that N doping introduced some states
at the valence band edge and thus made the original bandgap of TiO2smaller, and that a vacancy could induce somestates in the band gap region, which acted as shallowdonors.510Nakano et al found that, in N-doped TiO2, deeplevels located at approximately 1.18 and 2.48 eV below theconduction band were attributed to the O vacancy state as
an efficient generation-recombination center and to the Ndoping which contributed to band gap narrowing by mixingwith the O 2p valence band, respectively.523Okato et al.found that, at high doping levels, N was difficult to substitutefor O to contribute to the band gap narrowing, instead givingrise to the undesirable deep-level defects.524
S dopant induced a similar band gap narrowing asnitrogen,489and the mixing of the sulfur 3p states with thevalence band was found to contribute to the increased width
of the valence band, leading to the narrowing of the bandgap.495,497When S existed as S4+, replacing Ti4+, sulfur 3sstates induced states just above the O 2p valence states, and
S 3p states contributed to the conduction band of TiO2 asshown in Figure 53A.494
When F replaced the O in the TiO2 lattice, F 2p stateswere localized below the O 2p valence states without anymixing with the valence or conduction band as shown inFigure 53B, and additional states appeared just below theconduction edge, due to the electron occupied level composed
of the t2gstate of the Ti 3d orbital.507The electronic changeinduced by F dopant was considered to be similar to the Ovacancy, thus reducing the effective band gap and improvingvisible light photoresponse.507Li et al found that F dopingproduced several beneficial effects including the creation ofsurface oxygen vacancies, the enhancement of surfaceacidity, and the increase of Ti3+ ions, and doped N atomsformed a localized energy state above the valence band ofTiO2, whereas doped F atoms themselves had no influence
on the band structure in N-F-co-doped TiO2.519
4.1.2.2 Optical Properties of Doped TiO 2
Nanomate-rials 4.1.2.2.1 Optical Properties of Metal-Doped TiO 2
Nanomaterials A red shift in the band gap transition or a
Figure 52 (A) Total DOSs of doped TiO2and (B) the projected
DOSs into the doped anion sites, calculated by FLAPW, for the
dopants F, N, C, S, and P located at a substitutional site for an O
atom in the anatase TiO2 crystal (eight TiO2 units per cell) Ni
-doped stands for N doping at an interstitial site, and Ni+s-doped
stands for doping at both substitutional and interstitial sites
Reprinted with permission from Asahi, R.; Morikawa, T.; Ohwaki,
T.; Aoki, K.; Taga, Y Science 2001, 293, 269
(http://www-.sciencemag.org) Copyright 2001 AAAS
Trang 35visible light absorption was observed in metal-doped TiO2
(refs 433-435, 438, 444, 445, 448, 449, 460-463, 465, 466,
470, 509, 548, and 549) For V-, Mn-, or Fe-doped TiO2,
the absorption spectra shifted to a lower energy region with
an increase in the dopant concentration.434,445,460This red shift
was attributed to the charge-transfer transition between the
d electrons of the dopant and the CB (or VB) of TiO2
Metal-ion doped TiO2prepared by ion implantation with various
transition-metal ions such as V, Cr, Mn, Fe, and Ni was
found to have a large shift in the absorption band toward
the visible light region, with the order of the effectiveness
in the red shift being V > Cr > Mn > Fe > Ni.466-471Anpo
et al found that the absorption band of Cr-ion-implanted
TiO2shifted smoothly toward the visible light region, with
the extent of the red shift depending on the amount of metal
ions implanted as shown in Figure 54A.470Impregnated or
chemically Cr-ion-doped TiO2 showed no shift in the
absorption edge of TiO2; however, a new absorption band
appeared at around 420 nm as a shoulder peak due to the
formation of an impurity energy level within the band gap,
with its intensity increasing with the number of Cr ions
(Figure 54B).470
In the study by Umebayashi et al., visible light absorption
of V-doped TiO2was due to the transition between the VB
and the V t2g level.509 The holes in the VB produced an
anodic photocurrent The photoexcitation processes under
visible light of V-, Cr-, and Mn-doped TiO2are illustrated
in Figure 55 Photoexcitation for V-, Cr-, Mn-, and Fe-doped
TiO2 occurred via the t2g level of the dopant The visible
light absorption for Mn- and Fe-doped TiO2was due to the
optical transitions from the impurity band tail into the CB
The Mn (Fe) t level was close to the VB and easily
overlapped in highly impure media The visible lightabsorption for the Cr-doped TiO2can be attributed to a donortransition from the Cr t2glevel into the CB and the acceptortransition from the VB to the Cr t2glevel
Stucky et al found that up to 8 mol % Eu3+ions could bedoped into mesoporous anatase TiO2, and excitation of theTiO2 electrons within their band gap led to nonradiativeenergy transfer to the Eu3+ ions with a bright red lumines-cence.287The mesoporous TiO2acted as a sensitizer
4.1.2.2.2 Optical Properties of Nonmetal-Doped TiO 2
Nanomaterials Nonmetal doped TiO2normally has a colorfrom white to yellow or even light gray, and the onset ofthe absorption spectra red shifted to longer wavelengths (refs
385, 426, 478, 483, 489, 494, 495, 497, 498, 505, 506, 512,
516, 518, 519, 521, and 529) In N-doped TiO2rials, the band gap absorption onset shifted 600 nm from
nanomate-380 nm for the undoped TiO2, extending the absorption up
to 600 nm, as shown in Figure 56.426The optical absorption
of N-doped TiO2 in the visible light region was primarilylocated between 400 and 500 nm, while that of oxygen-deficient TiO2was mainly above 500 nm from their density-functional theory study.520N-F-co-doped TiO2prepared byspray pyrolysis absorbs light up to 550 nm in the visiblelight spectrum.518The S-doped TiO2 also displayed strongabsorption in the region from 400 to 600 nm.494The red shifts
in the absorption spectra of doped TiO2 are generallyattributed to the narrowing of the band gap in the electronicstructure after doping.489 C-doped TiO2 showed long-tailabsorption spectra in the visible light region.472,543Cl-, Br-,and Cl-Br-doped TiO2 had increased optical responsecompared to the case of pure TiO2in the visible region.508Livraghi et al recently found that N-doped TiO2containedsingle atom nitrogen impurity centers localized in the bandgap of the oxide which were responsible for visible light
Figure 53 (A) Total DOS of S-doped TiO2 Reprinted with
permission from Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai,
K.; Mitsui, T.; Matsumura, M Appl Catal A 2004, 265, 115,
Copyright 2004, with permission from Elsevier (B) Total DOSs
of F-doped TiO2calculated by FLAPW Egindicates the (effective)
band gap energy The impurity states are labeled (I) and (II)
Reprinted from Yamaki, T.; Umebayashi, T.; Sumita, T.;
Yama-moto, S.; Maekawa, M.; Kawasuso, A.; Itoh, H Nucl Instrum.
Methods Phys Res., Sect B 2003, 206, 254, Copyright 2003, with
permission from Elsevier
Figure 54 (A) The UV-vis absorption spectra of TiO2(a) and
Cr ion-implanted TiO2 photocatalysts (b-d) The amount ofimplanted Cr ions (µmol/g) was (a) 0, (b) 0.22, (c) 0.66, or (d) 1.3.
(B) The UV-vis absorption spectra of TiO2(a) and Cr ion-dopedTiO2(b′-d′) photocatalysts prepared by an impregnation method.The amount of doped Cr ions (wt%) was (a) 0, (b′) 0.01, (c′) 0.1,(d′) 0.5, or (e′) 1 Reprinted from Anpo, M.; Takeuchi, M J Catal.
2003, 216, 505, Copyright 2003, with permission from Elsevier.