Uploaded by Кузьмин Евгений

3-s2.0-B9780128015780000059-main

advertisement
CHAPTER 5
Silica and Titania Nanodispersions
J. Nestor, J. Esquena
Institute of Advanced Chemistry of Catalonia, Department of Chemical and Biomolecular
Nanotechnology, Spanish National Research Council (IQAC-CSIC) and CIBER on Bioengineering,
Biomaterials and Nanomedicine (CIBER-BBN), Barcelona, Spain
5.1 Introduction
Tremendous effort devoted to nanomaterials has resulted in a rich database for their
synthesis, properties, modification, and application. Continuous breakthroughs in the
synthesis and modification of nanomaterials have brought new properties and new
applications with improved performance. The synthesis and processing of fine ceramic
particles at the nanometer scale have recently received considerable attention owing to
their interesting optical, electronic, and catalytic properties. Significant progress has been
made in the preparation of a large number of inorganic colloids consisting of particles
with different chemical compositions, sizes, and morphologies. However, a number of
questions still have to be answered satisfactorily, including chemical and physical
mechanisms of their formation and growth, as well as the control of particle morphology.
This chapter provides insight into silica (SiO2) and titania (TiO2) nanoparticles (NPs).
Silicon dioxide, universally known as silica, can be natural or synthetic, crystalline or
amorphous, and it has been used in glass manufacturing since ancient times. Silica is highly
ubiquitous and abundant, being the major constituent of sand. It is most commonly found in
nature as quartz and is present in various living organisms such as diatoms.1 However, silica
is also a highly complex material, existing in many different mineral forms in nature and
also being synthetically produced. Notable examples of synthetic silica include fused quartz,
fumed silica NPs, and silica hydrogels, xerogels, and aerogels.2 Applications range from
semiconductor devices to ceramics and metallurgy, to name just a few.
This chapter deals mostly with synthetic amorphous silica in colloidal dispersions. All
forms of silica contain the SieO bond, which is the most stable of all SieX covalent
bonds. The SieO siloxane bond is just 0.162 nm long, which is considerably shorter than
the sum of covalent radii of silicon and oxygen atoms (0.191 nm).3 The short length of the
siloxane bond largely accounts for its partial ionic character, and it is responsible for its
Nanocolloids. http://dx.doi.org/10.1016/B978-0-12-801578-0.00005-9
Copyright © 2016 Elsevier Inc. All rights reserved.
159
160 Chapter 5
high chemical stability. Silica NPs can be prepared by many different techniques. This
chapter focuses on methods used to prepare colloidal silica, which refers to stable
dispersions of discrete amorphous silica particles. Special attention is devoted to solegel
chemistry and processing.
Another inorganic oxide of great interest is titanium dioxide (TiO2), also known as titania.
It is being commercially manufactured at the scale of millions of tons because of its
chemical stability, low cost, availability, and its remarkably high refractive index (w2.4).
TiO2 is widely used as a white pigment in paints, as an ultraviolet (UV) absorber in
sunscreen creams, and as an abrasive in toothpastes. Since 1972, when Fujishima and
Honda4 discovered the photocatalytic splitting of water on TiO2 electrodes, TiO2
constitutes the archetypal photocatalyst in heterogeneous photocatalysis because of its
relatively high efficiency and chemical stability.5 The discovery of this phenomenon
generated a tremendous amount of research related to TiO2 photocatalysis, including solar
fuels and environmental remediation.
TiO2 has three common crystalline polymorphic phases: rutile (tetragonal), anatase
(tetragonal), and brookite (orthorhombic). Rutile and anatase are the two most common
crystal structures. Although all of them are found in nature, rutile is the most abundant
since it is the only thermodynamically stable mesomorphic phase.6 Rutile is usually
commercially used as a white pigment, whereas anatase is used because of its
photocatalytic performance. Anatase is considered the most important phase of TiO2 in
terms of trade, but it rarely exists in nature in pure form. This phase is only stable at
temperatures lower than 500 C and becomes rutile at temperatures higher than 600 C.
Anatase is used as a photocatalyst for many purposes, especially in water and air
treatment, removing pollutants, decomposing microorganisms such as bacteria and viruses,
and removing organic contaminants, such as phenol and its derivatives, as well as textile
paint formulations. TiO2 NPs possess interesting optical, dielectric, and photocatalytic
properties. Therefore many efforts have sought to produce TiO2 NPs with controlled size,
morphology, and porosity for thin films, ceramics, composites, and catalysts.
5.2 Synthesis Methods Based on SoleGel Processing in Solution
Solegel is a rather old term that has been used since the 1930s to describe the synthesis
and processing of glass and ceramics.7 Initially, solegel referred to the formation of solid
suspensions in a liquid phase (denoted as “sols”) and the formation of a covalently
cross-linked network with a very high viscosity (denoted as “gels” or, preferably,
“hydrogels”), forming a soft material that can be casted into molds. Nowadays, the term
solegel refers to synthesis based on two consecutive reaction steps: (1) formation of
solutions of solvated metal precursors, generally by hydrolyzation reactions; and (2)
polycondensation or polyesterification of the hydrolyzed species. Solegel processing
Silica and Titania Nanodispersions 161
allows advanced materials to be fabricated in a wide variety of forms: ultrafine or
spherical powders, thin-film coatings, fibers, porous or dense materials, and extremely
porous aerogel materials.2 An overview of various solegel processes is illustrated
in Fig. 5.1.
Stabilized under proper conditions, particle dispersions (sols) do not become a gel or
settle even after several years of storage, and they may contain up to about 50% silica,
with particle sizes up to 300 nm, although particles larger than about 70 nm slowly
settle. Pioneering work in silica chemistry by Iler1 led to the industrial development of
colloidal silica dispersions, commercially known as Du Pont’s Ludox silica. Stöber et al.8
extended Iler’s findings to show that using ammonia as a catalyst for the
tetraethyl orthosilicate (TEOS) hydrolysis reaction, one could control precisely the
size of microparticles, yielding so-called Stöber spherical silica microparticles,8
Figure 5.1
Schematic representation of a solegel processes for synthesis of nanomaterials. Adapted with
permission from Brinker, C. J.; Scherer, G. W. SoleGel Science: The Physics and Chemistry of
SoleGel Processing; Academic Press: London, 1990.
162 Chapter 5
which are nonporous particles with a narrow size distribution (practically monodisperse)
and a smooth surface.
The most important advantage of solegel processing is that it allows extremely pure oxide
materials to be formed, which is a factor that is not only an advantage, but a necessity for
applications in electronics and high-energy optics. Other potential advantages of solegel
particles over particles obtained by other methods are controlled size and morphology,
molecular scale homogeneity,9e11 and enhanced reactivity because of lower processing
temperatures. For successful commercial applications, these advantages must outweigh
inherent disadvantages such as higher costs, lengthy processing times, and low yields.
Mackenzie12e14 summarizes a number of potential advantages and disadvantages and the
relative economics of solegel methods. Hench and colleagues15,16 compare quantitatively
the merits of solegel-derived silica materials for optics over alternative high-temperature
processing methods.
The most widely used precursors in solegel processing are alkoxides (M(OR)n), because
they are stable at standard conditions they are easily hydrolyzed with water, and they do
not generate toxic by-products. This family of precursors has been widely used to obtain a
large variety of inorganic oxide NPs, including TiO2,17,18 Al2O3,19 Fe2O3,20 and
YBa2Cu3O7 superconductor.21
5.2.1 Synthesis of Silica NPs by Hydrolysis of Alkoxides
Solegel processing
In the case of silica, the most-used alkoxide precursor is TEOS, since it is stable at room
temperature in normal laboratory conditions. The solegel reaction of alkoxides consists of
two steps: hydrolysis and polycondensation. Hydrolysis follows the following reaction2,22:
SiðORÞ4 þ n H2 O/SiðORÞ4 nðOHÞn þ 4 n ROH
[5.1]
The fully hydrolyzed product is silicic acid (Si(OH)4), which is rather unstable and
polycondensates, producing species with high molecular mass. The condensation of the
hydrolyzed species is described by the reactions
2 SiðORÞ3 OH/ðORÞ3 Si O SiðORÞ3 þ H2 O
ðORÞ3 Si OR þ HO SiðORÞ3 /ðORÞ3 Si O SiðORÞ3 þ ROH
[5.2]
[5.3]
These three reactions normally take place simultaneously, leading to the formation of
siloxane (SieOeSi) bridges and finally producing highly pure silica. The overall
reaction is
SiðORÞ4 þ 4 H2 O/SiO2 þ 4 ROH
[5.4]
Silica and Titania Nanodispersions 163
In this reaction a solvent with intermediate hydrophilicity, such as an alcohol, is required
since water and alkoxysilanes are mutually immiscible, as shown in the TEOS/ethanol/
water ternary phase diagram in Fig. 5.2. However, gels can be prepared in silicon
alkoxideewater mixtures without added a solvent, since the alcohol released as a
by-product of the hydrolysis is sufficient to homogenize the initially phase-separated
system.
The formation of silica hydrogels was described long ago by Carman23 and studied in
detail by Iler.1 According to this model, silica primary particles are formed by hydrolysis
of a precursor and successive polymerization. Nucleation of primary particles is produced,
and particle nuclei evolve, forming a stable suspension (sol) at a pH between 7 and 10 and
in the absence of salt. Otherwise, at pH <7 and/or in the presence of electrolytes, the
primary particles aggregate, producing a three-dimensional (3D) network (gel). The gel
structure can be controlled depending on the size of the primary particles when the
solegel transition occurs.
Figure 5.2
Ternary phase diagram of tetraethyl orthosilicate/distilled alcohol (95% ethyl alcohol, 5% water)/
water at 25 C. For pure ethanol, the miscibility line is shifted slightly to the right. Reproduced with
permission from Brinker, C. J.; Scherer, G. W. SoleGel Science: The Physics and Chemistry of
SoleGel Processing; Academic Press: London, 1990.
164 Chapter 5
Nuclear magnetic resonance (NMR) results provides further information about the
mechanism, demonstrating that condensation takes place in such a way that maximizes the
number of SieOeSi bonds and minimizes the number of terminal hydroxyl groups
through internal condensation. Thus siloxane rings are quickly formed, to which
monomers are added, creating 3D particles. These particles condense to the most compact
state, leaving silanol OH groups on the outside. According to Iler,1 the 3D species serve as
nuclei. Further growth occurs by an Ostwald ripening mechanism whereby particles
increase in size and decrease in number, as highly soluble small particles dissolve and
reprecipitate on larger, less soluble nuclei. The growth stops when the difference in
solubility between the smallest and largest particles becomes only a few parts per million.
As a consequence, this mechanism makes it possible to obtain monodisperse silica
particles.
Summarizing the Iler mechanism, silica gels are formed as a result of three steps that can
occur simultaneously:
1. polymerization of monomers to form NPs
2. growth of NPs
3. Linking of NPs into chains, thus forming three-dimensional networks that extend
throughout all the liquid medium and thicken into a gel
Solegel processes are modulated by different variables such as pH and the dielectric
constant of the medium. It has been shown that at a pH between 7 and 10, and in the
absence of electrolytes, silica particle growth follows a “monomer cluster” type, leading to
stable colloidal dispersions without particle aggregation. However, at a pH <7 (or a higher
pH in the presence of salts) the growth can be defined as “clusterecluster,” leading to the
aggregation of particles and forming a hydrogel tridimensional network, instead of stable
discrete particles.
Catalysis and control of solegel reactions
Hydrolysis and condensation can be catalyzed in an acidic or basic medium. Furthermore,
strongly nucleophile anions such as fluoride ions (F) can accelerate these reactions.
Hydrolysis reactions
Hydrolysis described in Eq. [5.1] corresponds to a nucleophilic substitution of alkoxy
groups (OR) by hydroxyl group (OHe) following an SN2-type mechanism. Although
hydrolysis can occur without the addition of an external catalyst, this reaction is more
rapid and efficient when a catalyst is used. Various authors24,25 described hydrolysis of
alkoxy silanes as a function of pH (Fig. 5.3).
Hydrochloric acid and ammonia are generally used instead of other catalysts such as
acetic acid, potassium hydroxide, amines, potassium fluoride, and hydrofluoric acid.2
Silica and Titania Nanodispersions 165
Figure 5.3
pH rate profile for hydrolysis in an aqueous solution, showing the hydrolysis kinetic constant k.
Reproduced with permission from Pohl, E. R.; Osterholtz, F. D. In Molecular Characterisation of
Composites Interfaces; Kruna, G., Ishida, H., Eds.; Plenum: New York, 1985.
Furthermore, it has been shown that the rate and extent of the hydrolysis reaction is
influenced mostly by the strength and concentration of the acid or base catalyst.26
Acid-catalyzed hydrolysis mechanism Aelion et al.26 found that all strong acids behave
similarly, whereas weaker acids require longer reaction times to achieve the same extent of
alkoxide hydrolysis. From a plot of the logarithm of the hydrolysis rate constant versus
acid concentration, a slope of 1 was obtained. They concluded that the reaction was of the
first order in acid concentration.
Under acidic conditions, the alkoxide group is protonated in a rapid first step. Electron
density is depleted from the silicon atom, making it more electrophilic and thus more
susceptible to attack from water. This results in the formation of a penta-coordinate
transition state with a significant SN2-type character. The transition state decays by
displacement of an alcohol and inversion of the silicon tetrahedron, as seen in Fig. 5.4.
Base-catalyzed hydrolysis mechanism Under basic conditions, the hydrolysis reaction was
found to be of the first order in [OH] concentration. However, if TEOS concentration is
increased, the reaction deviated from a simple first-order to a more complex second-order
reaction. If weaker bases are used, such as ammonium hydroxide and pyridine, measurable
speeds of reaction were obtained only if large concentrations were present. Therefore,
compared with acidic conditions, hydrolysis kinetics at a basic pH are more strongly
affected by the nature of the solvent.
In any case, base-catalyzed hydrolysis of silicon alkoxides take places more slowly than
acid-catalyzed hydrolysis at an equivalent catalyst concentration.26 Basic alkoxide oxygens
166 Chapter 5
Figure 5.4
Acid-catalyzed hydrolysis. Reproduced with permission from Brinker, C. J.; Scherer, G. W. SoleGel
Science: The Physics and Chemistry of SoleGel Processing; Academic Press: London, 1990.
Copyright 1990 Elsevier Inc.
tend to repel the nucleophile OH. Once an initial hydrolysis has already occurred,
however, the following condensation reactions progress steadily, with each subsequent
alkoxide group eliminated from the monomer more easily than the previous one.27
Therefore, partially hydrolyzed alkoxides are more prone to attack. In addition, hydrolysis
of the forming polymer is more sterically hindered than the hydrolysis of a monomer.
Although hydrolysis in alkaline environments is slow, it still tends to be complete and
irreversible.28
Thus, under basic conditions, hydroxyl anions attack silicon atoms. Again, a SN2-type
mechanism has been proposed in which the eOH displaces eOR, with inversion of the
silicon tetrahedron. This is seen in Fig. 5.5.
Influence of the water-to-silicon molar ratio (r) on hydrolysis and condensation of silicon alkoxides
Hydrolysis is usually performed with r values ranging from less than 1 to over 50,
depending on the desired polysilicate product. From Eq. [5.2] it is inferred that
increasing r is expected to promote the hydrolysis reaction. Aelion et al.26 described that
acid-catalyzed hydrolysis of TEOS is of the first order in water concentration. However,
they also observed an apparent zero-order dependence of the water concentration under
Figure 5.5
Base-catalyzed hydrolysis. Reproduced with permission from Brinker, C. J.; Scherer, G. W. SoleGel
Science: The Physics and Chemistry of SoleGel Processing; Academic Press: London, 1990.
Silica and Titania Nanodispersions 167
basic conditions. This is probably a result of the production of monomers by siloxane
bond hydrolysis and redistribution reactions (ie, the reverse reactions of those described
in Figs. 5.4 and 5.5. Nonetheless, the most obvious effect of the increased value of r is
the acceleration of the hydrolysis reaction. In addition, higher values of r cause the
rather complete hydrolysis of monomers, before significant condensation occurs. In any
case, the overall process is complex, since the different extents of monomer hydrolysis
influence the relative rates of the alcohol- or water-producing condensation reactions. In
general, with under-stoichiometric additions of water (r 2), the alcohol-producing
condensation mechanism (Eq. [5.3]) is favored, whereas the water-forming condensation
reaction (Eq. [5.2]) is favored when r > 2.29
Although increased values of r generally promote hydrolysis, when r is increased while
maintaining a constant solvent-to-silicate ratio, the silicate concentration is reduced. This
in turn reduces the hydrolysis and condensation rates, resulting in longer gel times.
Therefore, two opposing factors define gelification kinetics. These were described by
Klein,30 who studied gel times for acid-catalyzed TEOS systems as a function of r
(H2O-to-Si molar ratio) and the initial ethyl alcoholetoeTEOS ratio. He found that
gelification time is minimal at H2O/TEOS ¼ 4, in the case of ethanol/TEOS ¼ 1.
Finally, since water is the by-product of the condensation reaction (Eq. [5.2]), large values
of r promote siloxane bond hydrolysis (the reverse of Eq. [5.1]).
Condensation reactions for the formation of silica
As mentioned before, polymerization to form siloxane bonds occurs by either an
alcohol-producing or a water-producing condensation reaction. Engelhardt et al.31 showed
that a typical sequence of condensation products is monomer, dimer, linear trimer, cyclic
trimer, cyclic tetramer, and higher-order rings (Fig. 5.3). This sequence of condensation
requires both depolymerization (ring opening) and the availability of monomers that are in
solution equilibrium with the oligomeric species and/or are generated by
depolymerization.
Influence of pH The rate of these ring-opening polymerizations and monomer-addition reactions is dependent on pH. In polymerizations below a pH of 2, condensation rates are
proportional to the [Hþ] concentration. Because the solubility of silica is quite low below
a pH of 2 (see Fig. 5.8), formation and aggregation of primary silica particles occur simultaneously, and ripening (growth of a network) contributes little to growth after particles
exceed 2 nm in diameter. Thus, 3D gel networks comprising small primary particles are
developed.
It is generally agreed that between pH 2 and pH 6, condensation rates are proportional
to [OH] concentrations. Condensation preferentially occurs in highly condensed
species, reacting with less highly condensed silica species, which are somewhat neutral.
168 Chapter 5
This suggests that the rate of dimerization is low. Once dimers form, however, they react
preferentially with monomers to form trimers, which in turn react with other monomers to
form tetrameters. Cyclization occurs because of the proximity of the chain ends and the
substantial depletion of the monomer population. Further growth occurs by adding
lower-molecular-weight species to more highly condensed species and by aggregation of
the condensed species to form chains and networks. The solubility of silica in this pH range
is again low, and particle growth stops when the particles reach 2e4 nm in diameter.28
Polymerization of silica species occurs rapidly at pH values above 2. However, at higher
pH values (mainly above pH 7), condensed species are strongly ionized and therefore
mutually repulsive. Thus growth occurs primarily through the addition of monomers to the
more highly condensed particles rather than by particle aggregation. However, because of
the greater solubility of silica species at basic pH, particles increase in size and decrease
in number as slightly more soluble small particles dissolve and re-precipitate on larger,
less soluble particles. Growth stops when the difference in solubility between the smallest
and largest particles becomes indistinguishable. This process is denoted as Ostwald
ripening of solid particles. As a consequence, particle size is strongly dependent on
temperature: higher temperatures produce larger particles. In addition, in the pH 2e6
range, the growth rate also depends on the particle size distribution.
Nature and concentration of catalyst As well as hydrolysis, condensation can proceed without
a catalyst. However, the use of catalysts for the condensation of organosiloxanes is very
helpful. Furthermore, the same catalysts are used: generally those compounds that exhibit
acidic or basic characteristics.
It has been shown that condensation reactions are acid and base specific.24 In addition,
Iler1 showed that under more basic conditions, gel times increase. Condensation reactions
continue to proceed; however, gelation does not occur. Again, catalysts that dictate a
specific pH can and do drive the type of silica particle produced, as noted in the previous
discussion of pH.
Acid-catalyzed mechanism It is generally believed that the acid-catalyzed condensation mechanism involves a protonated silanol species. Protonation of silanol makes the silicon more
electrophilic and thus susceptible to nucleophilic attack. The most basic silanol species
(silanols contained in monomers or weakly branched oligomers) are the most likely to be
protonated. Therefore condensation reactions may occur preferentially between neutral
species and protonated silanols situated on monomers, end groups of chains, and so on.
Base-catalyzed mechanism The most widely accepted mechanism for the base-catalyzed
condensation reaction involves the attack of a nucleophilic deprotonated silanol on a
neutral silicic acid (Fig. 5.6).
Silica and Titania Nanodispersions 169
Figure 5.6
Nucleophilic attack to form a siloxane bond.
Furthermore, it is generally believed that the base-catalyzed condensation mechanism
involves penta- or hexa-coordinated silicon intermediates or transition states that are
similar to those of an SN2-type mechanism.
Preparation of monodisperse silica by precipitation in solution
As described earlier, the classic method for preparing silica microparticles from alkoxides
is well known as the Stöber method.8 Tetraethoxysilane or tetraethyl TEOS is a silane
monomer produced from tetrachlorosilane, which in turn is derived from
metallurgical-grade silicon. The silicon itself is obtained by the reduction of naturally
occurring silica at temperatures over 1900 C in the presence of carbon.
Using the Stöber method, different sizes of silica NPs can be prepared, ranging from 50
to 1 mm, with a very narrow size distribution. Typical Stöber particles are dense and
nonporous, have a spherical shape, and possess very smooth surfaces. The size of
these particles depends on the alkoxide chain type and the alcohol mixture used as a
solvent.8 Particles prepared in methanol solutions are the smallest, whereas the
particle size increases with an increase in the alcohol chain length. The particle size
distribution also becomes broader when longer-chain alcohols are used as solvents.
Many studies have been devoted to unraveling the mechanisms of Stöber silica
formation.32e37
An example of a typical colloidal silica with a quite uniform distribution made by the
Stöber process is shown in Fig. 5.7.
Formation of monodisperse silica by the Stöber-Fink-Bohn method can be explained by
two mechanisms: homogeneous nucleation and aggregative growth.33e35 Homogeneous
nucleation is believed to take place via the LaMer model (monomer addition),39,40 in
which nucleation is a fast process, followed by a particle growth process without further
nucleation.41e43 By contrast, the aggregative growth (aggregation) model considers that
nucleation and aggregation occur simultaneously throughout the reactions, and the nuclei
(primary particles) aggregate together to form dimers, trimers, and larger particles
(secondary particles). These models lead to the formation of either spherical particles
(LaMer model) or gel network (aggregation model), depending on the reaction
conditions.33,44e46
170 Chapter 5
Figure 5.7
Left: Scanning electron micrograph of Stöber spherical silica powders. Right: Size distribution of
typical Stöber silica particles. Reprinted with permission from Khadikar, C. The Effect of Adsorbed
Poly(Vin 1 Alcohol) on the Properties of Model Silica Suspensions (Ph.D. dissertation); University of
Florida: Gainesville, FL, 1988.
The stoichiometric amount of water for hydrolysis is 4 mol for 1 mol of silicon alkoxide,
or only 2 mol water if condensation is completed. In the preparation of particles, however,
the typical ratio of water to TEOS is higher than 20:1, and the pH is very high, since both
factors promote condensation. This leads to the formation of compact structures (Stöber
dense particles) rather than the extended polymeric networks generally found in hydrogels.
Formation and stability of sols
As predicted by Iler1 and confirmed by Harris52 in most sols that consist of discrete
spherical particles of amorphous silica, the interior of the particles is made of anhydrous
SiO2 with a density of 2.2 g/cm3. The silicon atoms located at the surface hold OH groups
that are not lost when the silica is dried to remove free water. Calculations of the silanol
number of the silica surface by purely geometric considerations and the density of
amorphous silica indicated that there should be 7.8 silicon atoms/nm2 around the surface.
Silica sols do not conform to the DerjaguineLandaueVerweyeOverbeek theory because
they are apparently stabilized by a layer of adsorbed water that prevents coagulation even
at the isoelectric point. This stabilization by repulsive hydration forces is possible because
of the unusually small Hamaker constant of silica. To destabilize an aqueous silica sol, the
degree of hydration must be reduced. Allen and Matijevı́c47,48 showed that the addition of
salt produces surface ion exchange, and stability is reduced near the isoelectric point.
Several researchers33,44e46 made many attempts to determine the size of the primary
particles using various techniques. Through small-angle X-ray scattering, Green et al.46
Silica and Titania Nanodispersions 171
reported that the size of primary particles was 10.3 nm (in methanol) or 20.7 nm (in
ethanol). Later, Rahman et al.49 produced homogeneous and stable silica NPs with mean
particle size of 7.1 1.9 nm (in ethanol). In all cases these particles were nonporous. Iler1
attributes the dense structure to the solubility of the oxide in the aqueous medium and
speculates that this might not be the case in a nonaqueous (eg, alcoholic) solution. This
hypothesis was confirmed later by Ramsay and Booth.50 Although the cores of the primary
particles in aqueous sols are nonporous, the presence of hydroxyl groups reduces their
density. The stability of sols is explained by the fact that the repulsive barrier increases
with the size of the particle; for this reason the nuclei may be unstable against aggregation
until they reach a certain size.
Mechanisms of particle growth
Harris and coworkers46 investigated the growth of particles produced by the Stöber
method, assuming that the first hydrolysis of an alkoxide group is the rate-limiting step
in the formation of nuclei. The remaining groups hydrolyze rapidly, and small nuclei are
created from fully hydrolyzed species. In the early stages of the reaction these small
nuclei aggregate, forming colloidally stable seed particles, and in the later stages of the
reaction particle growth occurs mainly by the addition of a monomer to the surface, not
by aggregation of the small nuclei particles, as claimed by Bogush et al.33e35 Using the
results of cryoetransmission electron microscopy experiments, Bailey and Mecartney51
postulated that hydrolyzed monomer polymerizes to form microgel clusters as a result of
polysilicic acid cross-linking; these clusters collapse upon reaching a certain size and
cross-linking density. Collapsed particles made more dense by condensation are
colloidally stable with respect to each other. The denser seed particles grow by adding a
hydrolyzed monomer or polymer to their surfaces. The rate of growth of the polymers
must be slow enough so that after a sufficient number of seeds has been formed, the
polymers attach to a particle’s surface before they grow to a large enough size to
collapse and form a seed particle themselves. Boukari et al.52 proposed that the initial
particles that also have a polymeric structure are better described as mass fractals, in
which the nucleating backbones or seeds are used to build the compact and stable
particles observed later during growth. Before equilibrium is reached, particles of
various sizes and fractality are distributed. In addition, under conditions in which
condensation is rapid compared with hydrolysis, polysiloxane chains or rings are
produced, whereas in the reverse case more extensively cross-linked polymeric clusters
are formed. Therefore polymers formed in base-catalyzed reactions generate highly
branched clusters.53
The effect of electrolytes on the size of silica NPs was described by Bogush et al.33e35
and in their studies, they reported that when the electrolyte (NaCl) concentration was
increased from 0 to 104 M, the particle size increased from 340 to 710 nm.
172 Chapter 5
Surface structure of silica
The surface of hydrated silica is covered with hydroxyl groups (silanols) that are attached
in various ways to silicon atoms. Using an infrared spectroscopy method, Yaroslavsky and
colleagues,54e56 proved for the first time the existence of hydroxyl groups on a silica
surface (porous glass). This fact was soon confirmed by Kurbatov and Neuymin.57 Now,
numerous spectral and chemical data unambiguously confirm the presence of OH groups
on such SiO2 surfaces. The hydroxyl groups at the surface are involved in
temperature-dependent dehydroxylation-hydroxylation, producing siloxane bridges. The
reaction is written as:
^SieOH%SieOeSi^ þ H2 O
[5.5]
Dehydroxylation usually starts at temperatures above 500 K, forming siloxane bridges on
the surface that are less polar than hydroxyl groups. On the other hand, it must be
emphasized that a hydrated silica surface, upon being exposed to a sufficiently high
pressure of water vapor, is able to take up water by means of physical adsorption and
capillary condensation.
The surface properties of amorphous silica, which is considered to be an oxide adsorbent,
in many cases depend on the presence of silanol groups. At a sufficient concentration
these groups make such a surface hydrophilic. The OH groups act as the centers of
molecular adsorption during their specific interaction with adsorbates that are capable of
forming a hydrogen bond with the OH groups. The removal of the hydroxyl groups from
the surface of silica leads to a gradual increase in hydrophobic properties.1,58
It is commonly accepted that several types of hydroxyl groups with different reactivity
coexist on the surface of silica. Surface OH groups are subdivided as (1) isolated free
(single silanols; SiOH); (2) geminal free (geminal silanols or silanediols); ]Si(OH)2; (3)
vicinal, or bridged, OH groups bound through the hydrogen bond (H-bonded single
silanols, H-bonded geminals, and their H-bonded combinations) (Fig. 5.8). On the SiO2
surface there also exist surface siloxane groups or SieOeSi bridges with oxygen atoms on
the surface. Moreover, internal silanol groups are often present inside the silica skeleton,
located in very fine ultramicropores (<1 nm), which may also contain hydration water
molecules.
Chemical modification of the silica surface The chemical modification of the silica surface
with organo-functional groups is an important step toward the preparation of silicae
polymer nanocomposites. More precisely, surface modifications have been reported to
enhance the affinity between the organic and inorganic phases and at the same time
improve the dispersion of silica NPs within the polymer matrix.60 Modification of the silica
surface with silane coupling agents is one of the most effective techniques available. Silane
coupling agents (Si(OR)3R0 ) have the ability to bond inorganic materials such as silica NPs
Silica and Titania Nanodispersions 173
Figure 5.8
Types of silanol groups and siloxane bridges on the surface of amorphous silica, and internal OH
groups (see text). Qn terminology is used in nuclear magnetic resonance, where n indicates the
number of bridging bonds (OeSi) tied to the central silicon atom: Q4, surface siloxanes; Q3,
single silanols; Q2, geminal silanols (silanediols). Reproduced with permission from Zhuravlev, L. T. The
Surface Chemistry of Amorphous Silica. Zhuravlev Model. Colloids Surf. A 2000, 173 (1e3), 1e38.
Copyright 2000, Elsevier Inc.
to organic resins. In general, the Si(OR)3 portion of the silane-coupling agents reacts with
the inorganic reinforcement, while the organofunctional group (R0 ) reacts with the resin.
Table 5.1 shows some common silane-coupling agents used for modification of a silica
surface.
In general, the chemical modification of a silica surface using silane-coupling agents
can be conducted via an aqueous or nonaqueous system that is also known as
post modification. The nonaqueous system is usually used for grafting
3-aminopropyltriethoxysilane (APTES) molecules onto the silica surface. The main reason
for using a nonaqueous system is to prevent hydrolysis. Silanes such as APTES, which
carries amine groups (base), can undergo uncontrollable hydrolysis and polycondensation
reactions in an aqueous system. Therefore the use of an organic solvent provides better
control of reaction parameters and is preferred for coupling reaction using APTES. For a
nonaqueous system, the silane molecules are attached to the silica surface via a direct
condensation reaction, and the reaction is usually conducted at reflux conditions. On the
other hand, an aqueous system is favorable for large-scale production. In this system
the silanes undergo hydrolysis and condensation before being deposited on the surface.
174 Chapter 5
Table 5.1: Commonly used silane-coupling agents.
Name (Acronym)
Formula
Vinyltriethoxysilane (VTS)
Methacryloxypropyltriethoxysilane (MPTS)
3-Glycidyloxypropyltrimethoxysilane (GPTS)
3-Aminopropyltriethoxysilane (APTES)
3-Mercaptopropyltriethoxysilane (McPTS)
Chloropropyltriethoxysilane (CPTS)
The alkoxy molecules are hydrolyzed in contact with water. This is followed by
self-condensation reactions between the hydrolyzed silanes. Then the silane molecules are
deposited on the silica surface through the formation of siloxane bonds between the silanol
groups and hydrolyzed silanes with the release of water molecules.61 A slight increase in
the particle size (z25%) after surface modification was observed.62
Silica and Titania Nanodispersions 175
5.2.2 Synthesis of TiO2 NPs by Hydrolysis of Alkoxides
TiO2 powders possess interesting optical, dielectric, and catalytic properties, which lead to
many industrial applications such as pigments, fillers, catalyst supports, and photocatalysts.
Since commercial production started in the early 20th century, TiO2 has been widely used
as a pigment,63 in sunscreen formulations,64,65 paints,66 ointments, and toothpaste,67
among others.
Synthesis of TiO2 NPs by precipitation methods, using reactants such as titanium
tetrachloride and titanium tetrabutoxide, is common. Titanium tetrachloride and titanium
tetrabutoxide are dissolved in a solvent (usually water), and precipitation is induced by
adding a basic solution. Various factors affect this precipitation method:
1. Use of an inhibitor, accelerator, and surfactant: Absorption of the inhibitor on the
crystal surface prevents the growth of NPs, and it can be used to produce fine crystals.
Adding inhibitors in other methods (eg, microemulsion) has the same effect. A surfactant covers the crystal surface and, by changing the crystal’s interfacial energy, prevents
the growth particles; thus it can be used to produce fine NPs.157
2. Reaction temperature: An increase in temperature, and the consequent increase in
diffusion coefficient, reduces the time needed for formation of stable nuclei and thus
decreases the induction time for nucleation. In most cases high temperatures also
increase solubility and decrease supersaturation; consequently, low supersaturation
growth favors homogeneous particle growth, creating larger particles with lower
polydispersity.
The final properties of TiO2 nanopowders depend on size, morphology, and crystalline
phase. To prepare a TiO2 nanostructured material with controlled properties, several
processes have been developed in the past decade and can be classified as liquid processes
(solegel,68e70 solvothermal,71 hydrothermal72,73), solid-state processing routes
(mechanical alloying/milling,74 mechanochemical,75,76 radiofrequency thermal plasma77),
and other routes such as laser ablation.78
Nowadays, the method most used to industrially produce nanopowders of TiO2 is based
on plasma spray synthesis. This technique allows high productivity but leads to the
production of NPs containing impurities, which reduce the quality grade of the product.
Moreover, control of crystallite size distribution and shape of the NPs is not
straightforward and depends strongly on the spray feedstock.79 On the contrary,
production of NPs by a chemical solegel method is the best method to ensure the high
purity of the final product. Furthermore, the solegel process gives the ability to go all
the way from the molecular precursor to the product, making it possible to synthesize
tailor-made materials with controlled particle size distribution.
176 Chapter 5
While silica is the most prevalent type of solegel network, there is increasing interest in
systems created from titanium and other transition metal alkoxides. These systems are
significantly more difficult to study and control because transition metals are more
electrophilic than silicon and consequently exhibit greater rates of hydrolysis and
condensation. The tendency of titanium alkoxides to rapidly hydrolyze makes their
solegel transformation difficult to control. Titanium alkoxides are quite reactivedmuch
more than silicon alkoxidesdand therefore must be handled with care, in a dry
environment; they are often stabilized via chemical modification,80e82 which is performed
mainly with acetic acid, acetylacetone, or hydrogen peroxide. In any case, hydrolysis of
titanium alkoxides occurs much faster than hydrolysis of silica counterparts, and
consequently, titanium alkoxides have to be protected from contact with water. Therefore,
storage in a nitrogen atmosphere is required.
In solegel processes TiO2 is usually prepared by acid-catalyzed reactions of hydrolysis
and polycondensation of titanium alkoxide ((TiOR)n) to form oxopolymers, which
are transformed into an oxide network. These reactions are similar to those used for
the synthesis of silica (see Section 5.3.1.1) and can be schematically presented as
follows.
Hydrolysis follows a nucleophilic substitution mechanism, where R is an alkyl group:
TiðORÞn þ H2 O/TiðORÞn1 ðOHÞ þ ROH
[5.6]
Condensation dehydration (alcoxolation) is a reaction by which a bridging oxo group is
formed through the elimination of an alcohol molecule. The mechanism is basically the
same as for hydrolysis:
TiðORÞn þ TiðORÞn1 /Ti2 OðORÞ2n2 þ ROH
[5.7]
Dealcoholation (oxolation) follows the same mechanism as alcoxolation, but the R group
of the leaving species is a proton:
2TiðORÞn1 ðOHÞ/Ti2 OðORÞ2n2 þ H2 O
[5.8]
Thus, the overall reaction is
TiðORÞ4 þ 2 H2 O/TiO2 þ 4 ROH
[5.9]
where the relative rates of hydrolysis and polycondensation strongly affect the structure
and properties of metal oxides.
Factors influencing particle formation
The performance of titania in technological applications could be optimized with specific
microstructural and macrostructural control over the morphology of the material.
Nucleation and growth of titania need to be controlled at the earliest stages of production
Silica and Titania Nanodispersions 177
and stopped at the appropriate stage to obtain titania particles with the desired crystal
structure and a well-defined size, shape, and surface structure.
One of the main advantages of the solegel technique over other methods, such as
hydrothermal synthesis and chemical vapor deposition, is that it produces materials with a
large surface area.83 Experimental results have shown that powders created by an
uncontrolled solegel method generally lack the properties of uniform size, shape, and
unagglomerated state. There are several parameters for controlling the solegel process to
prepare a TiO2 nanopowder with desired properties. It has been demonstrated that the
precursor’s concentration of titanium alkoxide greatly affects the crystallization behavior
and characteristics of the final powder.84
Factors that are crucial in the formation of metal oxides include the reactivity of metal
alkoxides, the water-to-alkoxide ratio, the pH of the reaction medium, the nature of the
solvent and additives, and the reaction temperature. By varying these process parameters,
materials with different surface chemistries and microstructures can be obtained. Typically,
in the solegel method, the solegel-derived precipitates are amorphous. Therefore further
heat treatment is required to induce crystallization. To induce transition from an
amorphous to an anatase phase, an annealing temperature higher than 300 C is generally
required; this results in the dramatic increase in of particle size. The transformation of
anatase to rutile occurs at temperatures from 400 to 1100 C.85e87 This transformation
results in a drastic loss of surface area and porosity.88
The alkoxy groups most often used in the synthesis of TiO2 range from two (ethoxide) to
four (butoxide) carbon atoms long. In the same way as silicon alkoxide, the reactivity
toward hydrolysis decreases as the alkoxy chain length increases, which provides a means
of controlling the rate of particle formation.82,89 Even so, hydrolysis in the presence of
excess water is rapid and completed within seconds. To moderate the high reactivity,
alkoxides are usually diluted in alcohol before mixing with water.90 In addition, the size,
stability, and morphology of the sol produced from alkoxides is strongly affected by the
water-to-titanium molar ratio (r ¼ [H2O]/[Ti]).91,92 The formation of colloidal TiO2 at a
high r ratio is of great interest because small particles are formed under this condition.
One can distinguish two regimes: (1) At low r values (r 10) one obtains spherical,
relatively monodisperse particles with diameters of 0.5e1 mm.93e95 (2) At higher r values
the particles formed are colloidally unstable and precipitate in the form of large
aggregates. These aggregates can be chemically peptized to final sizes that are usually
smaller than 100 nm in diameter.96e98 Because of the rapid precipitation in these
conditions, which makes kinetic studies extremely difficult, the process remains poorly
understood.
The titania solegel process follows a different pathway than the silicon-based solegel
process. It is generally assumed that individual monosilicates serve as building blocks
178 Chapter 5
during the polymerization process.99 In the case of titania, however, titanium clusters
(polyoxoalkoxides) have been suggested to serve as the primary building blocks for
particles and gels. In particular, Ti11O13(OC3H7)18 has been proposed as a building block
in titanium isopropoxide (TTIP)ebased solegel processes based on results from 17O NMR
spectroscopy.100,101 A stable solution of these clusters can be obtained by hydrolysis of
TTIP at low hydrolysis ratios (h < 1). However, if h > 1.5, hydrolysis of TTIP does not
result in a stable solution of polyalkoxides but rather in the precipitation of TiO2 particles.
The precipitation takes place after an induction period in which slow particle growth is
observed, followed by rapid precipitation. The presence and size evolution of NPs
(1.5e6 nm) during the induction period has been studied by dynamic light scattering.102
Simonsen et al.103 showed that the titanium clusters formed using titanium tetraethoxide
were smaller than the clusters formed using TTIP and titanium tetrabutoxide under similar
conditions. Furthermore, the small 3-nm particles synthesized at pH 3 grow significantly
during the drying process as a result of the destabilization of the colloidal solution, leading
to the formation of a gel confirmed by high-resolution transmission electron microscopy.
Stabilization of titania NPs
Hydroxypropyl cellulose can be used to improve the monodispersity of TiO2 NPs.
According to Nagpal et al.,104 TiO2 particles less than 100 nm in diameter can be prepared
when tetraethylorthotitanate (TEOT) reacts in the presence of hydroxypropylcellulose
(HPC) dispersant for r (H2O/TEOT) ¼ 60/120. In addition, they assumed that the amount
of HPC adsorbed onto the surface of TiO2 particles increased with an increasing water
concentration, leading to greater steric stabilization. To prevent aggregation, Jean and
Ring105 also used HPC during the precipitation of TiO2 particles and synthesized
monodispersed powders. In general, TiO2 fine particles were synthesized by the controlled
hydrolysis and condensation of TEOT in a dilute alcohol solution106,107 using a batch
precipitation technique on a small scale, in which hydrolysis was carried out using a
beaker and a magnetic mixer in a nitrogen atmosphere. However, TiO2 particles created
through a batch process were not enough to obtain nanophase particles with a narrow
particle size distribution. Accordingly, the synthesis of TiO2 NPs using a semibatch
process was introduced in this work. The semibatch method, in which system TEOT
reactants are fed into a reactor at a constant rate, is better than a batch process at
controlling the size, shape, and size distribution because of its short nucleation and slow
hydrolysis rate.
Peptization
One of the problems encountered when making titania is the high reactivity in water of the
available precursors, such as alkoxides or chlorides. The reaction invariably leads to a
mixture of different polymeric species, which assume a variety of structures and sizes, and
finally lead to an amorphous solid. Although certain chemical modifications have been
Silica and Titania Nanodispersions 179
successful in controlling the reactivity of titanium alkoxides and have permitted the
isolation of intermediate polyoxoalkoxides,108e110 none of these species resemble the
structure of the bulk dioxide, and any further hydrolysis again leads to gels and amorphous
precipitates. In both cases a peptization process is required to convert these polymers into
crystalline particles.
The size of the particles at higher pH values is smaller, and they tend to agglomerate.
According to several studies, the variety of TiO2 surface charge is dependent on pH. TiO2
in sols is electrically charged because of the absorption of Hþ or OH in an aqueous
suspension. According to Bahnemann et al.,111 the surface charges of TiO2 can be
determined by chemisorptions:
of Hþ ; TiO2 þ nHþ 4TiO2 Hn þ for pH < 3:5
of OH ; TiO2 þ nOH 4TiO2 OHn for pH > 3:5
[5.10]
[5.11]
In acidic and alkaline media the strong repulsive charge among particles reduces
the probability of coalescence, and more stable sols can be formed. Studies by
Su et al.112 indicated that the isoelectric point of TiO2 powders varies in the pH range
of 5e6.8.
The hydrolysis of titanium isopropoxide or titanium ethoxide in excess water, followed by
peptization with nitric acid at refluxing temperatures, yields a sol comprising TiO2
NPs.97,98 Green et al.46 found that charged particles have a low sticking probability
because of the electrostatic repulsive forces between particles that counteract the attractive
van der Waals forces, producing smaller aggregates. Fig. 5.9 shows the effect of pH on the
surface charge of TiO2 NPs.
Figure 5.9
Effect of pH on the surface charge of an oxide. Adapted with permission from Bischoff, B. L.; Anderson,
M. A. Peptization Process in the SoleGel Preparation of Porous Anatase (TiO2). Chem. Mater. 1995,
7 (10), 1772e1778. Copyright 1995 American Chemical Society.
180 Chapter 5
A simple method to obtain TiO2 NPs by hydrolysis and peptization was described by
Bartlett et al.113 It consists of weighing TTIP precursor in an inert atmosphere, then
adding 18.5% hydrochloric acid to the precursor, giving rise to a transparent precipitate.
The precipitate is introduced in dilute hydrochloric acid solution under agitation. After
some time (which depends on the pH) a stable TiO2 dispersion is obtained.
5.3 Preparation of Silica and TiO2 NP Hydrolyzing Alkoxides
in Nanostructured Liquids
5.3.1 Synthesis in Microemulsions
Traditional solegel technology has the disadvantage that many precursors are not initially
soluble in water, and thus organic solvents (eg, alcohols) are required as reaction media.
One solution to this problem is the use of water-in-oil (W/O) microemulsions,114,115 which
can solubilize large amounts of both polar and nonpolar components. The use of
microemulsions, both oil-in-water (O/W) and W/O, has received considerable attention
because of its technological applications in the preparation of NPs with a controlled size,
achieving size distributions narrower than those achieved by conventional methods of
precipitation in solution. Microemulsions were first used for producing nano-sized
particles by Boutonnet et al.116 in 1982; they reported the preparation of platinum,
palladium, rhodium, and iridium NPs by reducing metal salts in W/O microemulsions.
These NPs were in the range of 3e5 nm, with very narrow pore size distributions. The
synthesis procedure, which is still the most widely used, consists of the preparation of two
identical microemulsions, each containing a separate reactant. The reaction is initiated
after mixing the two systems. The microemulsion droplets are subject to Brownian motion,
and they collide continuously to form aggregates, which last for very short periods of
time. The internal content of the microemulsion droplets can be exchanged rapidly as a
result of continuous coalescence and droplet disruption; consequently, diffusion of the
reactants is generally a fast process, and solid particles can be detected in a time scale of
minutes.
This microemulsion method is versatile, since hydrophilic reactants can be used in W/O
microemulsions and lipophilic ones in O/W microemulsions. Consequently, a wide variety
of different nano-sized materials can be prepared by this technique. Some examples
include gold and silver particles,117 carbonates,118 superconductors,119 semiconductors,120
magnetic materials,121 copper,122 and silica.123,124 Many scientific reports that focus on the
use of microemulsions as nanoreactors for the preparation of solid particles can be found
in the literature, and reviews are abundant.20,125e131
Osseo-Assure and Arriaga132,133 demonstrated that the particle size of silica NPs obtained
in microemulsions was extremely narrow; this had not been achieved before using the
Silica and Titania Nanodispersions 181
conventional technique of silica precipitation in a homogeneous alcohol media.8 Friberg
and Sjöblom134 studied, using NMR, the distribution of the intermediate species SinO2nr
(OH)2rx(OR)x during the formation of silica particles. They concluded that TEOS is
located within the lipophilic domains but the reactions proceed rapidly when the
molecules reach the interface of the water pools.
Particle size can be significantly increased by the addition of co-surfactants. This effect
has been observed when adding benzyl alcohol to O/W microemulsions stabilized with
sodium bis(2-ethylhexyl) sulfosuccinate (AOT) surfactant132 and adding alcohols to O/W
microemulsions prepared with block copolymer surfactants or conventional ethoxylated
surfactants.124 All these results indicate that the rates of intermicellar exchange are
increased by the addition of co-surfactants. It is well known that alcohols disrupt
molecular organization on films, increasing the film’s flexibility.135e138 This aspect has
been studied theoretically using Monte Carlo simulations, which have highlighted the
importance of film flexibility and molecular exchange between droplets.128
Control of particle size in microemulsion-based synthesis
The control of NP size is a complex and intriguing field. Primitive models were based on
the assumption that microemulsion droplets were nanoreactors that determined the size of
particles, depending on the H2O-to-surfactant ratio and the reactant concentrations in the
aqueous domain of the microemulsion droplets. However, these simple considerations
were rapidly outdated by the fact the particles (commonly around 30 nm) are significantly
bigger than the size of the microemulsion droplets that were used as “templates.” For
example, Yamauchi et al.123, who studied the preparation of silica particles using the
anionic surfactant AOT, showed that the size of the NPs (ranging from 10 to 80 nm,
depending on composition) was much larger than the size of the corresponding
microemulsion droplets (5e10 nm).
In the majority of cases there is not a direct correlation between the size of the particles
and the size of the microemulsion droplets,127 despite the fact that this dependence has
been observed in microemulsion systems based in AOT surfactant.127 In most works
reported in the literature, there is an increase in size during the reactions as a result of
reactant diffusion by particle exchange. Consequently, the microemulsion droplets cannot
be considered as “real templates,” since microemulsion droplets are not molds that restrict
particle size. For example, very large silica particles can be obtained by adding alcohols as
co-surfactants to microemulsions.124 Few particles are initially nucleated, but they can
grow rapidly thanks to fast diffusion. Fig. 5.10 shows an example of silica particles with
an approximate diameter of 150 nm.
The dynamics of NP formation and growth in microemulsion systems are complex, but
recent advances have provided clear understanding.20,127,139e141 Monte Carlo simulations
182 Chapter 5
Figure 5.10
Scanning electron microscopy image of silica particles obtained in a system comprising 10 wt%
NH3/C12e14E4.5/n-butanol/isooctane/tetraethyl orthosilicate (TEOS) (15 wt% aqueous solution,
C12e14E4.5:isooctane ¼ 8:3, 20 wt% butanol, 10 wt% TEOS). Reproduced with permission from Esquena,
J.; Tadros, T. F.; Kostarelos, K.; Solans, C. Preparation of Narrow Size Distribution Silica Particles Using Microemulsions. Langmuir 1997, 13 (24), 6400e6406. Copyright 1997 American Chemical Society.
have demonstrated the importance of film flexibility and the interdroplet exchange rate, in
addition to the H2O-to-surfactant ratio and the concentration of reactants.139,141 The main
conclusions, assuming that the chemical reactions are much faster than the material
exchange, were summarized by López-Quintela20:
1. An increase in reactant concentrations increases both particle size and polydispersity.
2. An increase in the concentration of one of the reactants, in excess with respect to the
second reactant, tends to decrease the particle size.
3. An increase in film flexibility increases the particle size. It can be achieved by adding
alcohols to the microemulsions,124 approaching the microemulsion phase boundaries,
changing the chain length of oil and co-surfactant, among others.
4. An increase in microemulsion droplet size, depending on the parameter R ¼ [H2O]/
[surfactant], produces an increase in particle size. However, it should be pointed out
that the relationship between R and the particle size is generally nonlinear because R
also influences film flexibility.
Silica and Titania Nanodispersions 183
In the case of TiO2, uncontrolled aggregation and flocculation was reported for
microemulsion-based methods,132 except at very low reactant concentrations.142 By
contrast, stable, monodisperse SiO2 NPs can be produced because of the relatively low
reactivity of silicon alkoxides, allowing better stability, which increases with the
surfactant-to-water ratio.2 Colloidal stability can be explained by a combination of steric
repulsion with secondary hydration forces.17,18 TiO2 sols, gels, and NPs were produced by
the controlled hydrolysis of tetraisopropyltitanate in AOT reverse micelles. NPs with
diameters less than 10 nm could be produced at relatively high Ti(IV) concentrations
(up to 0.05 mol/dm3). These NPs aggregated into sols, with diameters of 20e300 nm. Lee
et al.143 recently showed that TiO2 NPs can be prepared by the hydrolysis of titanium
tetraisopropoxide in W/O microemulsions consisting of water, nonionic Brij series
surfactants with hydrophilic groups of different lengths, Tween series surfactants with
hydrophobic groups of different lengths, and cyclohexane. Particles have a spherical shape
and show a uniform size distribution, but the shape becomes distorted with a decrease of
hydrophilic group chain length following rapid hydrolysis of water and titanium alkoxide.
TiO2 NPs calcined at 500 C have a stable anatase phase that has no organic surfactants;
the product completely transforms into the anatase phase above 300 C, and the rutile
phase begins to appear at 600 C, regardless of surfactants.
5.3.2 Synthesis in Nano-emulsions
One major disadvantage of microemulsion formation is that it requires larger amounts of
surfactant than conventional emulsions, typically over 20 wt%. This drawback can be
solved by the use of emulsions with droplets in the nanometer range as reaction media,
since formation of small-size emulsions requires lower surfactant concentrations, in
comparison to microemulsions. These emulsions, termed nano-emulsions, miniemulsions, or
ultrafine emulsions, are transparent or translucent (with droplet sizes between 50 and
200 nm) or milky (sizes up to 500 nm)144e148 and exhibit high kinetic stability. Nanoemulsions are a subject of increasing interest in both theoretical discussions and practical
applications because of their singular properties, namely extremely small droplet size,
kinetic stability, and transparency. In addition, they present several advantages over
conventional emulsions, principally because of their similar characteristics to
microemulsions.145 In this sense Porras et al.149 used W/O nano-emulsions as reaction
media for the synthesis of SiO2 and TiO2 ceramic particles. They demonstrated that SiO2
and TiO2NPs, with sizes ranging from 30 to 230 nm, can be obtained by alkoxide
hydrolysis in W/O nano-emulsions. NP sizes can be tuned by controlling the nano-emulsion
droplet sizes. Fig. 5.11 shows examples of TiO2 and SiO2 NPs obtained using this method.
More recently, Wu et al.150 demonstrated a nano-emulsion templating approach for the
synthesis of silica NPs with a compartmentalized, hollow structure, with a uniform diameter
of 12 nm and a cavity diameter of 3.3 nm, via an ultrasound-assisted solgel method.
184 Chapter 5
Figure 5.11
(A) Scanning electron micrograph of silica particles obtained from a nano-emulsion. (B) Atomic
force micrograph of titania particles obtained from a nano-emulsion. Reproduced from Porras, M.;
Martı´nez, A.; Solans, C.; González, C.; Gutiérrez, J. M. Ceramic Particles Obtained Using W/O
Nano-Emulsions as Reaction Media. Colloids Surf. A 2005, 270e271 (1e3), 189e194, with permission
from Elsevier.
5.4 Preparation of Silica and Titania by Other Methods
In general, because of the price of metal alkoxides, the production of NPs via the solgel
process is around 100 times more expensive than production using conventional silica
melting and casting. As a consequence, the production of NPs via the solgel process
makes economic sense only when addressing high value-added products.
The most common raw materials used in the preparation of silica products include
sodium silicate ((Na2O) (SiO2)y), chlorosilanes (RxSiCl4x), and silicon alkoxides
(Si(OR)4). Among these, sodium silicate has the lowest cost on a per-weight basis and is
a commodity chemical available in large quantities. Sodium silicate can be readily
acidified to produce silicic acid (Si(OH)4), from which a wide range of silica
microstructures, ranging from gels with a large surface area to colloidal particles, can be
produced. Silicic acid can be subsequently processed (eg, gelled, precipitated) by
changing the temperature, pH, and/or solid content. Numerous techniques have been
proposed for manufacturing silica sols from soluble silicate, including dialysis,151
electrodialysis,152 peptization,153 acid neutralization,154 and ion exchange.155 The last
one has come to be the most commonly used in the industry. This method is described in
the flow chart in Fig. 5.12.
There are four typical routes to prepare colloidal silica from natural ore. The first among
them is the preparation of colloidal silica from liquid sodium silicate through an ion
Silica and Titania Nanodispersions 185
Figure 5.12
Main methods applied in the industry to obtain colloidal silica. Reproduced with permission from
Lim, H.; Lee, J.; Jeong, J.; Oh, S.; Lee, S. Comparative Study of Various Preparation Methods of Colloidal
Silica. Engineering 2010, 2 (12), 998e1005. Copyright 2010 Scientific Research Publishing Inc.
exchange process. The sodium silicate is obtained from naturally available silica by
melting the silica ore in the presence of alkali flux; subsequently, it is dissolved by heating
under pressure to produce liquid sodium silicate, which is commonly known as water
glass. The liquid sodium silicate has a high viscosity, and therefore it is diluted to a
concentration of 3e5 wt%. Next, it is passed through an ion-exchange resin, and then fed
into the alkali solution to form a silica seed, which is then used to grow silica particles.
The product is concentrated to 30 wt% to obtain the commercial product. The second
method for preparing colloidal silica is from TEOS, which is well known as the Stöber
method8 and was described in detail in Section 5.2. TEOS is a silane monomer prepared
from tetrachlorosilane, which in turn is derived from metallurgical-grade silicon. The
silicon itself is obtained by the reduction of naturally occurring silica ore at temperatures
over 1900 C in the presence of carbon. The third method is that of direct oxidation of
silicon, wherein colloidal silica is prepared by the direct oxidation of metallurgical-grade
silicon without using TEOS. The silicon is treated with water in the presence of alkali
catalysts to produce silica along with the evolution of hydrogen and heat. Finally, the
fourth method for preparing colloidal silica consists of the milling and peptization of silica
(silica gel or fumed silica), which consists of preferentially coacervated or aggregated
primary silica particles. The properties of the colloidal silica prepared by this route largely
depend not only on the milling and peptization process but also on the properties of the
starting silica source, such as purity, shape, and aggregation.
186 Chapter 5
In any case, silica and TiO2 NPs can be obtained by many different methods. The next
section briefly describes the most usual methods (ion exchange, vapor phase,
mechanochemical, hydrothermal/solvothermal, aerosols, and peptization) that are
alternatives to typical solegel processing. Other methods to obtain silica sols, including
dialysis and electrodialysis, have also been proposed, but they were never applied on an
industrial basis and are not in the scope of this chapter.
5.4.1 Preparation of NPs by Ion Exchange Methods
Alkaline silicate is a common and inexpensive source of silica. Silica can be precipitated
by acidification of alkaline silicate solutions with a mineral acid. Sulfuric acid and sodium
silicate solutions are added simultaneously with agitation to water, and silica is
precipitated. The choice of agitation, the duration of precipitation, the addition rate of
reactants, their temperature and concentration, and the pH can vary the properties of the
silica. The formation of a gel stage is avoided by stirring at high temperatures. The
obtained NPs are porous, with a diameter of 5e100 nm. However, this precipitation
method has the disadvantage that control is difficult, with bad reproducibility when
scaling up.
The acidification of alkaline silicate solutions can be improved by ion exchange methods.
The general principle is to remove sodium from silicate via cation exchange with Hþ,
using exchange columns in diluted solutions. When sodium content is reduced at lower pH
values, polymerization takes place in a controlled manner, and silica NPs are nucleated
homogeneously and begin to grow. After reaching the desired size, the stable NP
dispersion is concentrated to the final content. For the preparation of large particles,
smaller particles can be used as seeds for further growth.
The resulting suspensions are commercially available as aqueous dispersions of colloidal
polysilicic acid (resulting from the condensation of two or more molecules of Si(OH)4),
with typical loads up to 50 wt% of solid material. pH and conductivity of the concentrated
sol are adjusted to ensure stability for at least 6 months. Table 5.2 lists examples of
commercially available silica dispersions.
Table 5.2: Selected commercially available silica dispersions.157
Trademark Name
Ludox
Nyacol/Bindzil
Levasil
Kostrosol
Silicadol
Supplier
W. R. Grace & Co.
Eka Chemicals (Akzo Nobel)
Bayer
Chemiewerke Bad Köstritz GmbH
Nippon Chemical Industrial Co.
Silica and Titania Nanodispersions 187
5.4.2 Preparation of SiO2 and TiO2 NPs by Vapor Phase Methods
The vapor phase method is the most important for the production of TiO2 pigment and
fumed silica. It is based on the oxidation of TiCl4 or SiCl4 in a furnace at very high
temperatures or with the use of plasma following Eqs. [5.12] and [5.13]:
TiCl4 þ O2 /TiO2 þ 2Cl2
SiCl4 þ O2 /SiO2 þ 2Cl2
[5.12]
[5.13]
In a study by Park and Park,159 SiCl4 vapor was hydrolyzed with water vapor
at 150 C as the first step to form oxychloride particles. The sample was then
converted into silica particles through further hydrolysis at 1000 C to produce silica
particles in a size range of 250e300 nm. Low-temperature vapor phase hydrolysis
was used more recently for the synthesis of nanostructured silica materials.160 Fumed
silica has a very strong thickening effect. Primary particle size is 5e50 nm. The
particles are nonporous and have a surface area of 50e600 m2/g. The density is
160e190 kg/m3.
In the case of TiO2, this method is performed at a rate of hundreds of tons per year. The
dry process involves a vapor phase oxidation reaction of TiCl4, which leads to the
production of amorphous, nano-sized TiO2. The obtained amorphous, nano-sized TiO2 is
then annealed at different temperatures to get desired crystalline phases such as anatase
or rutile. The resulting powder is nearly monodisperse, with a primary particle size of
w250 nm.
5.4.3 Preparation of SiO2 and TiO2 Particles by Mechanochemical Methods
One interesting alternative synthesis method for inorganic oxides is the mechanochemical
activation of solid-state reactions. The mechanical treatment of powder solids in
high-energy mills significantly increases the number of defects in the solid structure,
leading to much lower crystallinity, smaller crystallite size, and a large surface area.161,162
During the high-energy ball-milling reaction, several types of impact between reactants,
balls, and walls of the container can occur. A schematic of such possible impacts is
depicted in Fig. 5.13.
In this way the reactivity of the activated solids is remarkably enhanced. Ball milling is a
very old method for accelerating solid-state chemical reactions, and it is beeing used as a
tool for synthesizing different materials, such as special alloys, nanocrystalline powders, or
metaleceramic composites.161 An attractive variation of this method allows nanometric
powders of simple metallic oxides to be obtained by solid-state reaction between a
metallic salt (Lewis acid) and a base. The mechanochemically activated reaction produces
188 Chapter 5
Figure 5.13
Three examples of possible manners by which attrition balls inside a mill can collide: head-on
impact (A); oblique impact (B); multiball impact (C). Reproduced with permission from Zhang, D. L.
Processing of Advanced Materials Using High-Energy Mechanical Milling. Prog. Mater. Sci. 2004,
49 (3e4), 537e560. Copyright 2004 Elsevier Inc.
oxides with low crystallinity that are immersed in a water-soluble salt by-product, which is
later removed by means of a simple washing procedure.163,164 Different oxides (Gd2O3,
CeO2, TiO2, etc.)165 and even metallic nanopowders166 have been obtained using this
technique. Compared with the most conventional preparative routes, the main advantages
of the mechanochemical method are the possibility of obtaining relatively large amounts
of product, the simplicity of the process, the low cost of raw materials, and the total
absence of organic solvents.
5.4.4 Preparation of SiO2 and TiO2 by Hydrothermal/Solvothermal Methods
The hydrothermal method has been widely used to prepare TiO2 nanomaterials.167e172
Hydrothermal/solvothermal synthesis is normally conducted in steel pressure vessels,
called autoclaves, under controlled temperature and/or pressure with the reaction in
aqueous/organic solutions. The temperature can be elevated above the boiling point of the
aqueous/organic solvents, reaching the pressure of vapor saturation. The temperature and
the amount of solution added to the autoclave largely determine the internal pressure
produced. The solvothermal method is almost identical to the hydrothermal method except
that a nonaqueous solvent is used. However, the temperature can be elevated much higher
than that in hydrothermal method because a variety of organic solvents with high boiling
points can be chosen. The solvothermal method normally provides better control than the
hydrothermal method over the size, shape distributions, and the crystallinity of TiO2 NPs.
For example, TiO2 nanorods were prepared by hydrothermal treatment of a titanium
trichloride aqueous solution supersaturated with NaCl at 160 C for 2 h.168 Different
surfactants or compositions can be used to tune the morphology of the resulting
Silica and Titania Nanodispersions 189
nanorods.172 When TiO2 powders are put into a 2.5- to 20-mol/L NaOH aqueous solution
and held at 20e110 C for 20 h in an autoclave, TiO2 nanotubes are obtained.171 With or
without the aid of surfactants, solvothermal methods have been used to synthesize
high-quality TiO2 NPs.158,173
5.4.5 Preparation of SiO2 and TiO2 Particles by Aerosol Methods
An aerosol is a colloidal dispersion of liquid droplets in a vapor. Aerosols can be used to
make oxide powders in a variety of ways. A widely practiced technique, generally known
as evaporative decomposition of solutions, involves spraying salt solutions into a furnace,
where the droplets dry and the salts decompose into oxides. Matijevic et al.174 propose a
variant method in which a gas stream passes through a vapor of AgCl to condense nuclei,
then this aerosol is passed over a film of titanium ethoxide or isopropoxide or chloride.
The vapor of the titanium compound condenses on the AgCl nuclei, and the droplets then
are hydrolyzed at a temperature lower than 100 C in a chamber containing water vapor.
Then the droplets are heated at 150 C to complete the reaction. In this method almost
monodisperse spherical particles with a diameter ranging from 60 to 600 nm were
obtained, as shown in Fig. 5.14.
Figure 5.14
(A) Transmission electron micrograph of TiO2 nanoparticles (NPs) prepared by hydrolysis of an
aerosol of titanium ethoxide. (B) Scanning electron micrograph of TiO2 NPs prepared by
hydrolysis of an aerosol of titanium isopropoxide. Reproduced from Matijevic , E.; Budnik, M.; Meites, L.
Preparation and Mechanism of Formation of Titanium Dioxide Hydrosols of Narrow Size Distribution.
J. Colloid Interface Sci. 1977, 61 (2), 302e311, with permission from Elsevier.
190 Chapter 5
5.4.6 Redispersion of NPs by Peptization
Peptization is the process of redispersing a colloid that has been weakly coagulated. This
can be done by adsorbing charged ions (and/or adjusting pH) that reestablish the repulsive
electrostatic double layer.2 An example of silica peptization consists of a two-step pH
adjustment. In this example an acid such as sulfuric or hydrochloric acid is added to a
dilute aqueous solution of sodium silicate while stirring, or while heating as necessary, to
neutralize pH and form a precipitate of silica hydrogel, which contains high electrolyte
concentrations. Next, this crude silica hydrogel is washed with water, removing the
electrolyte, and afterward pH is adjusted to 8.5e10, which is far from the isoelectric point
of silica (pH z 2). The precipitate is then heated for several hours in an autoclave
(120e150 C) to allow the hydrogel to peptize and form a stable colloidal suspension.
During this process, electrostatic repulsion increases and primary coagulated particles are
redispersed, forming a silica slurry. The silica sol prepared using this method can contain
up to 30 wt% SiO2. However, it has the drawback of creating irregular particles, despite
the fact that they can be relatively small (10e20 nm).153
5.5 Properties and Applications
The properties of nanomaterials are usually dependent on size. A nanomaterial often
exhibits unique physical and chemical properties compared with its bulk counterparts. As
described earlier, not much literature is available describing the size-dependent properties
of silica NPs. Some properties such as specific surface area and photoluminescence with
respect to the particle size are rarely reported. Therefore in this section some general
size-dependent properties of nanoceramics are briefly discussed.
5.5.1 Silica NPs
Unique properties of nano-sized SiO2 have led to a marked increase in its application. The
easy preparation methods and versatile properties have led to applications in fields and
products such as biomedicine, catalysis, sensors, coatings, electronic and thermal
insulators, and thin-film substrates.1e5 However, many of these applications are heavily
dependent on the size of the NPs.
A complete review dealing with recent uses of silica-based materials was published in
2013 by Ciriminna et al.,157 in which emerging applications in the controlled release of
fragrance and aromas, active pharmaceutical ingredients and biocides, inks and coatings,
and catalysis for fine chemicals are listed. Many advances have also been achieved in
more conventional applications such as catalysis, additives, and construction and
separation techniques.
Silica and Titania Nanodispersions 191
Catalysis
Silica NPs have attracted attention in catalysis because of their improved efficiency under
mild and environmentally benign conditions in the context of “green” synthesis. Because
of their enormously large and highly reactive surface area, NPs have the potential to
exhibit superior catalytic activity in comparison to their bulk counterparts. Commonly
used metallic/bimetallic nanocatalysts are expensive and toxic. Reports show remarkable
catalytic activity of silica NPs synthesized using the Stöber method in bis-Michael
addition175 and anti-Markovnikov addition of thiols to alkenes.
Mesoporous silica NPs allow the confinement of reactions into the pores. These
nanoreactors provide efficient catalytic materials for producing commodity chemicals and
energy. Furthermore, surface functionalization allows the excellent control of many
reactions.
Additives
By adding SiO2 NPs, many product characteristics are improved, including suspension
stability, thixotropy, adhesion strength between the substrate and coating, and the degree
of finish. For example, in paint formulations SiO2 nanopowder is added to water base
formulations at 0.3e1% of the total weight (fully dispersed). The main advantage is a
significant reduction in paint drying time, which is shortened. Furthermore, other
advantages are that resistance to UV radiation is greatly increased, the scrub resistance is
also increased by several orders of magnitude, and the stain resistance of the coating is
improved significantly. 253,254
SiO2 NPs can be added to plastic, when fully dispersed in polypropylene,
polyvinylchloride (PVC) plastic raw materials can significantly improve strength,
toughness, wear resistance and aging resistance of plastics. For example, nano-modified
polypropylene can match or exceed the performance of nylon 6 in terms of water
absorption, insulation resistance, compressive residual deformation, flexural strength, and
other indicators. 255
Strengthening fillers for concrete
SiO2 NPs can be produced in large quantities and at low cost, allowing application in
concrete. It may replace cement in the mix, which is the most costly and environmentally
unfriendly component in concrete. The use of NPs makes concrete financially more
attractive and reduces the CO2 footprint of the concrete products produced. The NPs also
increase the product properties of the concretedthe workability and the properties in a
hardened statedenabling the development of high-performance concretes for extreme
constructions. That means that a concrete with better performance, lower costs, and an
improved ecological footprint can be designed.
192 Chapter 5
Drug delivery systems
Amorphous silica particles are nontoxic and are regularly used as food additives and
components of vitamin supplements (as colloidal suspensions). Encapsulation in silica has
also been found to prolong the shelf life and preserve the metabolic activities of
enzymes,176 bacteria,177 and mammalian cells.178 This confirms the high compatibility of
silica with biological systems. The use of a solegel process to prepare nanocapsules
allows synthesis at low temperatures and makes silica a good alternative to organic
polymers to be used as a stable, nontoxic platform for biomedical applications such as
drug delivery. Furthermore, the release rates can be precisely controlled by tailoring the
internal structure of the particles for a desired diffusion (release) profile.233,234
Stabilization of pickering emulsions
An interesting application of silica NPs is the stabilization of O/W emulsions, preparing
so-called Pickering emulsions.179 Giermanska-Kahn et al.180 started with classical
surfactant-stabilized O/W emulsions, which are sheared and size fractionated to attain
monodispersity. Then the surfactant is removed by a dialysis method and simultaneously
replaced by silica particles. In this way stable Pickering emulsions are obtained with the
same size distribution as the initial surfactant-stabilized ones. Dulle and Glatter181 and
Sadeghpour et al.182 obtained very stable O/W emulsions of about 200 nm stabilized by
nano-sized hydrophilic silica after a simple surface treatment method with oleic acid,
improving the hydrophobicity of the particles while maintaining their charge and stability.
5.5.2 Titania NPs
The properties of TiO2 NPs strongly depend on the crystallographic phase, structure and
morphology, and particle size. TiO2 exhibits good photocatalytic properties; hence is used
in antiseptic and antibacterial formulations, in degrading organic contaminants and germs,
as a UV-resistant material, in the manufacture of printing ink, self-cleaning ceramics and
glass, coatings, and so on. TiO2 is also used in the manufacturing of cosmetic products
such as sunscreen creams, whitening creams, morning and night cosmetic creams, skin
milks, and toothpastes. Uses of TiO2 are also found in the paper industry for improving
the opacity of paper. Another application is the decontamination of smart textiles, as
described by Senic et al.183,184
The use of TiO2 as photoactive catalyst for solar fuels, photovoltaics, and environmental
remediation constitutes the main advanced research areas for TiO2. In all these
applications TiO2 performs three important functions: first, as a photochemical energy
conversion material because of the electronic structure of photocatalysts; second, as a
photosensitizer substrate because of its surface area and surface stability in artificial
photosynthesis applications; and third, as an electron transport scaffold in photovoltaic
Silica and Titania Nanodispersions 193
applications because of its electrical properties. The research findings derived from these
investigations have then been used for other advanced applications such as sensing,
biomedicine, and electronics.182e193 The next two sections describe in more detail
applications of TiO2 NPs in photocatalysis and photovoltaics.
Photocatalytic applications
The tuning of photocatalytically active materials such as TiO2 has recently attracted
attention because of the enormous significance of these materials in the fields of energy
conversion and chemical synthesis. There are numerous ways to improve the performance
of catalysts, for example, via modifying the composition by doping194e196 or by using
cocatalysts. But optimizing the morphology (surface area and porosity),197 the particle
size,198,199 and the crystallinity200,201 also offers various possibilities.
Because of the presence of a small number of oxygen vacancies, TiO2 is an n-type
semiconductor. The valence band of this oxide is mainly formed by the overlap of the
oxygen 2p orbitals, whereas the lower part of the conduction band is mainly constituted
by the 3d orbitals of Tiþ
4 cations. The band gaps, that is, the void region that extends
from the top of the filled valence band to the bottom of the vacant conduction band, are
3.2 and 3.0 eV for anatase and rutile, respectively. Therefore TIO2 is active under
near-UV light (UVA band from w3.0e3.9 eV). These large values avoid the absorption
of a significant fraction of visible light, resulting in poor solar-to-hydrogen conversion
efficiency.
Heterogeneous photocatalysis belongs to the group of advanced oxidation processes, which
refers to procedures to remove organic (or also inorganic) pollutants from water through
reactions involving hydroxyl radicals. Specifically, the process of photocatalysis (either in a
gaseous or an aqueous medium) is initiated by light energy equal to or greater than the band
gap of the semiconductor. As a consequence, excitation of an electron from the valence band
to the conduction band is produced, thus generating electron(eBC)ehole(hþBV) pairs in
the semiconductor material, as in the case of a TiO2 particle, illustrated in Fig. 5.15. Such
electronehole pairs possess an extremely high capacity to oxidize and thus are able to
completely mineralize a large variety of both organic and inorganic chemicals.
The lifetime of such electronehole pairs is in the nanosecond regime.202 This is sufficient
for the created pair to undergo a charge transfer to adsorbed species on the semiconductor
surface from contact with a solution or gas phase. However, the electronehole can also
recombine with the release of heat. This process can occur either at the surface or in the
particle bulk, and it actually competes with the charge transfer. Recombination depends on
parameters such as the degree of crystallinity.203 At the surface, the semiconductor can
donate electrons to reduce the oxygen, which is a common electron acceptor in aerated
solutions.204 On the other side, the holes can be combined with electrons given by the
194 Chapter 5
Figure 5.15
Different processes that take place in a semiconductor TiO2 particle under near-ultraviolet light
irradiation. Adapted with permission from Linsebigler, A. L.; Lu, G.; Yates, J. T., Jr. Photocatalysis on TiO2
Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95 (3), 735e758.
Copyright 1995 American Chemical Society.
donor species. Two main oxidative paths are recognized today.205 In the first the holes
might directly oxidize the adsorbent compounds, whereas in the second they might oxide
adsorbed water or the hydroxyl groups to form OH radicals, which eventually are
responsible for the mineralization of the adsorbed species.
It should be pointed out that heterogeneous processes are efficient only when the pollutant
concentration is relatively low (up to 100 ppm). Moreover, the reaction is enhanced if the
donor species are preadsorbed onto the catalyst’s surface. The performance of TiO2
photocatalysts strongly depends on the crystal phase, particle size, and surface structure
(eg, surface hydroxyl, oxygen vacancy, specific surface area).207 In this way the
preparation and processing of the photocatalyst have a great influence on the
physicochemical properties, which affect the photocatalytic activity. Heterogeneous
photocatalysis shows some advantages with respect to other advanced oxidation processes,
such as the efficiency in eliminating toxic halogenated chemicals, that additives are not
necessary, and that the photocatalyst remains intact after the reaction.
Photocatalytic activity of TiO2 is particularly interesting for the development of
self-cleaning materials. These materials have recently received substantial interest208
because of their wide range of applications in various fields ranging from indoor
applications of fabrics,209 furnishing materials,210 and window glass211 to exterior
construction materials, roof tiles,212 car mirrors,210 and solar panels.213 These materials
can easily be cleaned by a stream of natural water such as rainfall, which in turn
significantly reduces the cost of routine maintenance.
Silica and Titania Nanodispersions 195
Photovoltaic applications
Photovoltaic applications of TiO2 have been well investigated.214e216 Because of the large
surface-to-volume ratio of NPs, photocatalytic processes are favored in the same way as
increasing the loading of photoactive sensitizers in hybrid solar cells, and they benefit
from optimization at the nanoscale of the diffusion length of electronehole pairs in
hybrid-heterojunction solar devices.217 More specifically, photovoltaic process involving
TiO2 nanostructures are the result of a complex interplay between concurring phenomena
taking place at the nanoscale, which include (1) the interaction of molecules (from small
inorganic to large organic compounds) with the surface of nanocrystals and nanostructures;
(2) the absorption of light, which raises such interacting systems to high-energy-excited
states; and (3) a series of charge transfer or redox processes, which lead the system to
recover its ground state by accomplishing the photovoltaic task.
Different inorganic materials, particularly NPs, are currently mixed with organic dyes and
polymers to enhance a device’s efficiency. These types of cells are called hybrid solar
cells. It is well established that NPs play a significant role in the performance of the
photovoltaic properties of a device. The introduction of trace amounts of NPs in polymer/
organic solar cells can significantly enhance and tune their mechanical, electrical, and
thermal properties.218
5.6 Current Trends and Novel Tendencies
5.6.1 Recent Tendencies on the Preparation of Silica NPs
Research on new alkoxides
Solegel chemistry offers the possibility to design molecular precursors via chemical
modification of metal alkoxides by nucleophilic reagents.80 Most organic ligands are
usually removed during hydrolysis and condensation reactions, while more strongly bound
organics or adsorbed solvent molecules are burnt during thermal treatment, leading to the
formation of oxide glass or ceramics. However, solegel synthesis at room temperature can
keep organic functional groups inside the inorganic gel matrix and lead to hybrid
materials. These new materials offer many opportunities for applications in different fields,
such as patternable optical devices, photonics, sensors, and catalysis. Ormosil, ormocer, or
ceramer219,220 materials can be easily synthesized because SiC bonds are highly covalent
and therefore do not break upon hydrolysis.
Interest in nanocomposites
Silica is a chemically stable material that often requires its combination with other
materials to form composites and obtain functionality.221e229 Novel silica-based
nanocomposites can offer tailor-made properties that target specific applications. An
interesting example is imparting electrical conductivity to porous silica made by solegel
196 Chapter 5
processes.223e226 Three approaches have been developed. The first one is
postfunctionalization, which consists of the synthesis of pure silica followed by surface
deposition of an electrically conductive material,223 obtaining a significant increase in
conductivity. In the second approach, functional species are introduced during the silica
solegel process, before condensation of silanols and formation of siloxane bonds. For
example, formation of silicaegraphene nanocomposites by single-step methods leads to
electrical conductivity values around 0.5 S/cm.225 A third approach, different from
conventional methods, was described by Warren et al.226 In this approach, a metaleligand
solegel precursor is used, incorporating high loads of metal into silica materials.
Uniformly sized and functionalized silica NPs doped with a large amount of metal were
obtained, achieving remarkable electrical conductivities.226 For example, silicaepalladium
nanocomposites exhibited an electrical conductivity only 80 times less than that of bulk
palladium, and exceeded the conductivity of porous carbons by a factor of 100.226 These
high conductivity values enable these materials to be implemented in high-current devices.
Nanocomposites might also have applications in catalysis, where control of surface
functionality is required. Hybrid TiO2e2SiO2 NPs can be obtained with a low degree of
TiO2 crystallinity,227 by reaction of 3-(propyltrimethoxysilyl) acetylacetone with Ti(OiPr)4.
The UV-visible spectrum of the resulting TiO2e2SiO2 materials showed that both fourand six-coordinate titanium atoms are present.
An interesting example of catalysis is the immobilization of amino acids (histidine and
glutamic acid) by amide bonds on the surface of mesoporous silica supports. These
immobilized amino acids were used as ligands for Fe(II), and the obtained biomimetic
iron complexes were tested as catalysts for cyclohexane oxidation at mild reaction
conditions with hydrogen peroxide as the oxidant.228
Other applications of nanocomposites can be found in nonlinear optics and light-emitting
devices. An example is organic/inorganic nanocomposites with lanthanide complexes,
which can possess luminescent properties based on energy transfer from the absorbed
coordinated ligands to the chelated lanthanide ions. Silica-based NPs with such lanthanide
complexes can be formed in situ by solegel chemistry.229 This example illustrates the
importance of solegel processes in the development of new materials for nonlinear optics
and light-emitting devices. Consequently, solegel nanocomposites have prompted great
academic interest in further investigations, and extensive research is focused on the
development of novel materials.
Research on encapsulation
There has recently been an increased interest in the use of solegel silica particles for
encapsulation of active molecules and controlled release.230e234 Silica particles offer an
interesting alternative to organic delivery systems. Their intrinsic hydrophilicity and
Silica and Titania Nanodispersions 197
biocompatibility, as well as the excellent protection they provide for their internal payload,
makes them perfect candidates for controlled drug delivery applications. Silica
microspheres (with monodisperse dimensions typically between 10 and 100 mm) or silica
matrix particles share the same basic advantages: enhanced mechanical strength, chemical
and physical stability, biocompatibility, and an environmentally benign nature. The release
rates for a given drug can be designed to range from very slow to very fast, as porosity,
and thus diffusion through different porous structures, can be easily controlled in any
solegel silicate. Furthermore, the surface of the particles can be functionalized to
minimize protein interaction and enhance blood circulation for active targeting.
5.6.2 Recent Tendencies on TiO2 NPs
Future advanced uses of TiO2 clearly depend on the development of more complex
materials with advanced functionalities. Materials that are purpose-built to perform specific
tasks are the targets of such current research. For example, for photocatalysis, TiO2 has the
advantage of good photocatalytic activity and photostability and low cost.235 However, one
limitation of TiO2 as a photocatalyst is its wide band gap, making TiO2 sensitive to only
UV light, which covers less than 5% of the solar spectrum. To overcome this intrinsic
disadvantage, two strategies have been pursued. The first is to broaden the active spectrum
of TiO2 by a variety of chemical modifications that adjust its electronic structure. The
other strategy, aimed at improving TiO2 utility, is to create TiO2 with a large surface area.
In photocatalysis the photocatalytic conversion scales with the surface area. This implies a
great interest in small NPs. For example, TiO2 NPs with sizes around 2e10 nm can be
synthesized using a modified solegel method at around 50e100 C.236e241
As mentioned before, TiO2 nanomaterials, including NPs, nanorods, nanowires, and
nanotubes, are widely investigated for various applications in photocatalysis, photovoltaics,
batteries, photonic crystals, sensors, UV blockers, “smart” surface coatings, pigments, and
paints. Various methods such as solegel, sol, hydrothermal/solvothermal, physical/
chemical vapor deposition, and electrodeposition have been successfully used in making
TiO2 nanomaterials. New physical and chemical properties emerge at the nanometer scale,
and they vary with the sizes and shapes of the nanomaterials. The movement of electrons
and holes in semiconductor nanomaterials is governed by the well-known quantum
confinement, the transport properties related to phonons and photons are greatly affected
by the size and geometry of the materials, and the specific surface area and surface-tovolume ratio increase dramatically as the size of a material decreases. The large surface
area brought about by small particle size is beneficial to most TiO2-based devices because
it facilitates reaction/interaction between the devices and interacting media, which mainly
occurs on the surface and depends on the surface area. As the size, shape, and crystal
structure of TiO2 nanomaterials change, not only does surface stability vary but the
198 Chapter 5
transitions between different phases of TiO2 under pressure or heat become size dependent
as well.
Research on nanocomposites or doped NPs
A serious limitation of TiO2-based photocatalysis is its low efficiency. The most important
limitations are probably the low rate of electron transfer to oxygen and the high
recombination rate of electronehole pairs, which limits the rate of photo-oxidation of
organic compounds on the surface of the catalyst. For this reason, many recent studies
have been carried out to produce doped TiO2 NPs or nanocomposites. Addition of noble
metals such as gold, silver, and platinum to TiO2, which may enhance the overall
photocatalytic efficiency, has been reported by many investigators. TiO2 modified by metal
clusters exhibits a reduced electronehole recombination rate as a result of the Schottky
barrier at the metaleTiO2 interface.206 Moreover, silver is known to have some
antimicrobial properties.242,243 Therefore TiO2 modified by silver NPs is a promising
material for antimicrobial applications.
Among the many various modification strategies for visible lightedriven titania, TiO2 NPs
can be hosts for metal or nonmetal dopants.244e246 The band gap of TiO2 can be modified
effectively by doping with transition metal ions (eg, vanadium247 or selenium248 at
titanium sites).
To enhance the visible light efficiency, much research has been undertaken to modify the
electron structure of TiO2 and to narrow its band gap by doping with either anions or
cations; this body of work has been reported and reviewed elsewhere.249,250
Research on coreeshell NPs
Applications of coreeshell materials can be found in the control of surface wetting.251 A
new type of surface with supported Ag@TiO2 (core@shell) particles has been described.
When irradiated with UV light, these surfaces become superhydrophilic, reducing the
water contact angle to zero.251 Surfaces can be converted back again to the
superhydrophobic state, with a water contact angle larger than 150 degrees, when samples
are heated or illuminated with visible light for a sufficient period of time. This
superhydrophobicesuperhydrophilic transition is fully reversible.
Other applications of coreeshell NPs can be found in enhanced photocatalytic processes.
Metal NPs can be anchored on the TiO2 surface, as isolated “nanoislands,” obtaining
heterointerfaces that are able to trap photogenerated electrons. Therefore, TiO2@metal
core@shell particles might have applications in the photodegradation of pollutants.
However, a typical problem is that metal particles may be corroded or dissolved via
photocatalytic reactions. On account of this problem, heterostructures with metal cores and
semiconductor shells (metal@TiO2) have attracted much deeper attention. Nevertheless,
Silica and Titania Nanodispersions 199
the semiconductor shells in metal@TiO2 NPs might be unstructured, exposing crystal
facets with low reactivity or having small specific surface areas. Consequently, it is
important to control the structures and morphologies of shells in metal@TiO2 particles,
and this objective has recently become an interesting research field. An example is the
development of Au@TiO2 NPs with a truncated wedge-shaped morphology,252 which can
be synthesized by a relatively simple hydrothermal method, obtaining interesting
photocatalytic properties. This example illustrates that the availability of new chemical
routes for the synthesis of heteronanostructures is renewing the interest of coreeshell
particles for catalytic applications.
5.7 Concluding Remarks
Silica and TiO2 NPs have been a focus of interest for both fundamental research and
technological developments. Tremendous effort has been devoted to the synthesis of SiO2
and TiO2 NPs with narrow size distributions, the formation of hybrid and/or
nanocomposite structures, and the control of properties. As a consequence, a rich database
of scientific information on tailor-made NPs has been established.
Many synthetic routes have already been investigated by the scientific community,
including precipitation in solution, synthesis in microemulsions and nano-emulsions,
nucleation and growth by ion exchange, synthesis in the vapor phase, synthesis in
aerosols, and synthesis using mechanochemical and solvothermal methods.
Fundamental research has often been focused on precipitation, microemulsion, and
nano-emulsion methods because of their precise control of particle properties,
whereas industry has centered its attention mainly on ion exchange and vapor phase
methods, since these methods allow NPs to be produced on a very large scale and at a
low cost.
In the scientific literature alkoxides are the most commonly used precursors because of
their chemical stability, controlled reactivity, and high purity. Consequently, the literature
on some alkoxides (eg, TEOS for the synthesis of silica) is extremely wide and
comprehensive. However, alkoxides are not often used in industry because of higher costs;
halides and water-soluble salts (such as TiCl4 or sodium silicate) are widely preferred in
production plants. Therefore some synthetic strategies and precursors of great industrial
interest are underrepresented in the scientific literature. For example, ion exchange
methods have been widely used in industry for many years at a very large scale, but
nevertheless, nucleation/growth processes, as well as nanoparticle stability, have been less
studied from a fundamental point of view. In conclusion, a number of synthetic pathways
have not been explored as systematically as others, implying that there is room for deep
fundamental research.
200 Chapter 5
References
1. Iler, R. K. The chemistry of Silica: Solubility, Polymerization, Colloid and Surface Properties and
Biochemistry of Silica; John Wiley & Sons: New York, 1979.
2. Brinker, C. J.; Scherer, G. W. SoleGel Science: The Physics and Chemistry of SoleGel Processing;
Academic Press: London, 1990.
3. Unger, K. D. Porous Silica; Elsevier Scientific Publishing Co.: Amsterdam, The Netherlands, 1979.
4. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature
1972, 238 (5358), 37e38.
5. Hernández-Alonso, M. D.; Fresno, F.; Suárez, S.; Coronado, J. M. Development of Alternative
Photocatalysts to TiO2: Challenges and Opportunities. Energy Environ. Sci. 2009, 2 (12), 1231e1257.
6. Zhang, H.; Banfield, J. F. Thermodynamic Analysis of Phase Stability of Nanocrystalline Titania.
J. Mater. Chem. 1998, 8 (9), 2073e2076.
7. Geffcken, W.; Berger, E.; German Patent 736411, 1939.
8. Stöber, W.; Fink, A.; Bohn, E. Controlled Growth of Monodisperse Silica Spheres in the Micron Size
Range. J. Colloid Interface Sci. 1968, 26 (1), 62e69.
9. Roy, D. M.; Roy, R. An Experimental Study of the Formation and Properties of Synthetic Serpentines and
Related Layer Silicate Minerals. Am. Mineral. 1954, 39, 957.
10. Roy, R. Aids in Hydrothermal Experimentation: II, Methods of Making Mixtures for Both “Dry” and
“Wet” Phase Equilibrium Studies. J. Am. Ceram. Soc. 1956, 39 (4), 145e146.
11. Roy, S. K. Characterization of Porosity in Porcelain-Bonded Porous Alumina Ceramics. J. Am. Ceram.
Soc. 1969, 52 (10), 543e548.
12. Mackenzie, J. D. State of the Art and Prospects of Glass Science. J. Non-Cryst. Solids 1982, 52 (1e3), 1e8.
13. Mackenzie, J. D. Glasses from Melts and Glasses from Gels, a Comparison. J. Non-Cryst. Solids 1982,
48 (1), 1e10.
14. Mackenzie, J. D. Ultrasonic Processing of Ceramics, Glasses and Composites. In Ultrasonic Processing of
Ceramics, Glasses and Composites; Hench, L. L., Ulrich, D. R., Eds.; Wiley: New York, 1984; p 1.
15. Hench, L. L.; Wilson, M. J. R.; Balaban, C.; Nogues, J. L. Processing of Large Silica Optics. In 4th
International Conference on Ultrastructure Processing of Ceramics, Glasses and Composites, Tucson, AZ; 1989.
16. Hench, L. L.; Wang, S. H. In Multifunctional Materials 1988, Vol. 878; p 76.
17. Moran, P. D.; Bartlett, J. R.; Bowmaker, G. A.; Woolfrey, J. L.; Cooney, R. P. Formation of TiO2 Sols,
Gels and Nanopowders From Hydrolysis of Ti(OiPr)4 in AOT Reverse Micelles. J. Sol-Gel Sci. Technol.
1999, 15 (3), 251e262.
18. Moran, P. D.; Bartlett, J. R.; Woolfrey, J. L.; Bowmaker, G. A.; Cooney, R. P. Formation and Gelation of
Titania Nanoparticles from AOT Reverse Micelles. J. Sol-Gel Sci. Technol. 1997, 8 (1e3), 65e69.
19. Pang, Y. X.; Bao, X. Aluminium Oxide Nanoparticles Prepared by Water-in-Oil Microemulsions.
J. Mater. Chem. 2002, 12 (12), 3699e3704.
20. López-Quintela, M. A. Synthesis of Nanomaterials in Microemulsions: Formation Mechanisms and
Growth Control. Curr. Opin. Colloid. Interface Sci. 2003, 8 (2), 137e144.
21. Kumar, P.; Pillai, V.; Bates, S. R.; Shah, D. O. Preparation of YBa2Cu3O7x Superconductor by Coprecipitation
of Nanosize Oxalate Precursor Powder in Microemulsions. Mater. Lett. 1993, 16 (2e3), 68e74.
22. Bergna, E. H. Colloid Chemistry of Silica. In The Colloid Chemistry of Silica, Vol. 234; American
Chemical Society, 1994; p 1.
23. Carman, P. C. Constitution of Colloidal Silica. Trans. Faraday Soc. 1940, 36, 964e973.
24. Pohl, E. R.; Osterholtz, F. D. In Molecular Characterisation of Composites Interfaces; Kruna, G.,
Ishida, H., Eds.; Plenum: New York, 1985.
25. McNeil, K. J.; DiCaprio, J. A.; Walsh, D. A.; Pratt, R. F. Kinetics and Mechanism of Hydrolysis of a
Silicate Triester, Tris(2-Methoxyethoxy)Phenylsilane. J. Am. Chem. Soc. 1980, 102 (6), 1859e1865.
26. Aelion, R.; Loebel, A.; Eirich, F. Hydrolysis of Ethyl Silicate. J. Am. Chem. Soc. 1950, 72 (12),
5705e5712.
Silica and Titania Nanodispersions 201
27. Kelts, L. W.; Effinger, N. J.; Melpolder, S. M. SoleGel Chemistry Studied by 1H and 29Si Nuclear
Magnetic Resonance. J. Non-Cryst. Solids 1986, 83 (3), 353e374.
28. Keefer, K. D. In Silicon Based Polymer Science: A Comprehensive Resource; Zeigler, J. M.,
Fearon, F. W. G., Eds.; American Chemical Society: Washington, DC, 1990; pp 227e240.
29. Assink, R. A.; Kay, B. D. SoleGel Kinetics I. Functional group Kinetics. J. Non-Cryst. Solids 1988,
99 (2e3), 359e370.
30. Klein, L. C. SoleGel Processing of Silicates. Ann. Rev. Mater. Sci. 1985, 15, 227e248.
31. Engelhardt, L. G.; Zeigan, D.; Jancke, H.; Wieker, W.; Hoebbel, D. 29Si-NMR-Spektroskopie an
Silicatlösungen. II. Zur Abhängigkeit der Struktur der Silicatanionen in wäßrigen Natriumsilicatlösungen
vom Na: Si-Verhältnis. Z. Anorg. Allg. Chem. 1975, 418 (1), 17e28.
32. Van Helden, A. K.; Jansen, J. W.; Vrij, A. Preparation and Characterization of Spherical Monodisperse
Silica Dispersions in Nonaqueous Solvents. J. Colloid Interface Sci. 1981, 81 (2), 354e368.
33. Bogush, G. H.; Tracy, M. A.; Zukoski, C. F., IV Preparation of Monodisperse Silica Particles: Control of
Size and Mass Fraction. J. Non-Cryst. Solids 1988, 104 (1), 95e106.
34. Bogush, G. H.; Zukoski, C. F., IV Studies of the Kinetics of the Precipitation of Uniform Silica Particles
Through the Hydrolysis and Condensation of Silicon Alkoxides. J Colloid Interface Sci. 1991, 142 (1), 1e18.
35. Bogush, G. H.; Zukoski, C. F., IV Uniform Silica Particle Precipitation: An Aggregative Growth Model.
J. Colloid Interface Sci. 1991, 142 (1), 19e34.
36. van Blaaderen, A.; Vrij, A. Synthesis and Characterization of Monodisperse Colloidal Organo-Silica
Spheres. J. Colloid Interface Sci. 1993, 156 (1), 1e18.
37. Sakka, S.; Kamiya, K.; Makita, K.; Yamamoto, Y. Formation of Sheets and Coating Films from Alkoxide
Solutions. J. Non-Cryst. Solids 1984, 63 (1e2), 223e235.
38. Khadikar, C. The Effect of Adsorbed Poly(Vin 1 Alcohol) on the Properties of Model Silica Suspensions
(Ph.D. Dissertation); University of Florida: Gainesville, FL, 1988.
39. Lamer, V. K.; Dinegar, R. H. Theory, Production and Mechanism of Formation of Monodispersed
Hydrosols. J. Am. Chem. Soc. 1950, 72 (11), 4847e4854.
40. Turgeon, J. C.; LaMer, V. K. The Kinetics of the Formation of the Carbinol of Crystal Violet. J. Am.
Chem. Soc. 1952, 74 (23), 5988e5995.
41. Matsoukas, T.; Gulari, E. Self-Sharpening Distributions Revisited-Polydispersity in Growth by Monomer
Addition. J. Colloid Interface Sci. 1991, 145 (2), 557e562.
42. Matsoukas, T.; Gulari, E. Monomer-Addition Growth with a Slow Initiation Step: A Growth Model for
Silica Particles from Alkoxides. J. Colloid Interface Sci. 1989, 132 (1), 13e21.
43. Matsoukas, T.; Gulari, E. Dynamics of Growth of Silica Particles from Ammonia-Catalyzed Hydrolysis of
Tetra-Ethyl-Orthosilicate. J. Colloid Interface Sci. 1988, 124 (1), 252e261.
44. Lee, K.; Sathyagal, A. N.; McCormick, A. V. A Closer Look at an Aggregation model of the Stober
process. Colloids Surfaces A Physicochem Eng Aspects 1998, 144 (1e3), 115e125.
45. Okudera, H.; Hozumi, A. The Formation and Growth Mechanisms of Silica Thin Film and Spherical
Particles through the Stöber Process. Thin Solid Films 2003, 434 (1e2), 62e68.
46. Green, D. L.; Lin, J. S.; Lam, Y. F.; Hu, M. Z. C.; Schaefer, D. W.; Harris, M. T. Size, Volume Fraction,
and Nucleation of Stober Silica Nanoparticles. J. Colloid Interface Sci. 2003, 266 (2), 346e358.
47. Allen, L. H.; Matijevic, E. Stability of Colloidal Silica. II. Ion Exchange. J. Colloid Interface Sci. 1970,
33 (3), 420e429.
48. Allen, L. H.; Matijevı́c, E.; Meites, L. Exchange of Naþ for the Silanolic Protons of Silica. J. Inorg.
Nucl. Chem. 1971, 33 (5), 1293e1299.
49. Rahman, I. A.; Vejayakumaran, P.; Sipaut, C. S.; Ismail, J.; Bakar, M. A.; Adnan, R.; Chee, C. K. An
Optimized SoleGel Synthesis of Stable Primary Equivalent Silica Particles. Colloids Surf. A 2007,
294 (1e3), 102e110.
50. Ramsay, J. D. F.; Booth, B. O. Determination of Structure in Oxide Sols and Gels from Neutron
Scattering and Nitrogen Adsorption Measurements. J. Chem. Soc. Faraday Trans. 1 1983, 79 (1),
173e184.
202 Chapter 5
51. Bailey, J. K.; Mecartney, M. L. Formation of Colloidal Silica Particles from Alkoxides. Colloids Surf.
1992, 63 (1e2), 131e138.
52. Boukari, H.; Lin, J. S.; Harris, M. T. Probing the Dynamics of the Silica Nanostructure Formation and
Growth by SAXS. Chem. Mater. 1997, 9 (11), 2376e2384.
53. Yamane, M.; Inoue, S.; Yasumori, A. SoleGel Transition in the Hydrolysis of Silicon Methoxide.
J. Non-Cryst. Solids 1984, 63 (1e2), 13e21.
54. Yaroslavsky, N. G. Dissertation, Cand. Phys.-Math. Sc.; GOI: Leningrad, 1948.
55. Yaroslavsky, N. G.; Terenin, A. N. Dokl. Akad. Nauk SSSR 66, 1949.
56. Yaroslavsky, N. G. Zh. Fiz. Khim 1950, 24, 68.
57. Kurbatov, L. N.; Neuymin, G. G. Dokl. Akad. Nauk SSSR 68, 1949; p 34.
58. Kiselev, A. V. In Surface Chemical Compounds and Their Role in Adsorption Phenomena; Kiselev, A. V.,
Ed.; Moscow State University Press: Moscow, 1957; p. 90 and p. 199.
59. Zhuravlev, L. T. The Surface Chemistry of Amorphous Silica. Zhuravlev Model. Colloids Surf. A 2000,
173 (1e3), 1e38.
60. Kickelbick, G. Concepts for the Incorporation of Inorganic Building Blocks into Organic Polymers on a
Nanoscale. Prog. Polym. Sci. (Oxf.) 2003, 28 (1), 83e114.
61. Vansant, E. F.; Voort, P. V. D.; Vrancken, K. C. Characterization and Chemical Modification of the Silica
Surface; Elsevier Science: New York, NY, USA, 1995.
62. Yu, Y. Y.; Chen, C. Y.; Chen, W. C. Synthesis and Characterization of OrganiceInorganic Hybrid Thin
Films from Poly(Acrylic) and Monodispersed Colloidal Silica. Polymer 2002, 44 (3), 593e601.
63. Pfaff, G.; Reynders, P. Angle-Dependent Optical Effects Deriving from Submicron Structures of Films
and Pigments. Chem. Rev. 1999, 99 (7), 1963e1981.
64. Zallen, R.; Moret, M. P. The Optical Absorption Edge of Brookite TiO2. Solid State Commun. 2006,
137 (3), 154e157.
65. Salvador, A.; Pascual-Martı́, M. C.; Adell, J. R.; Requeni, A.; March, J. G. Analytical Methodologies for
Atomic Spectrometric Determination of Metallic Oxides in UV Sunscreen Creams. J. Pharm. Biomed.
Anal. 2000, 22 (2), 301e306.
66. Braun, J. H.; Baidins, A.; Marganski, R. E. TiO2 Pigment Technology: A Review. Prog. Org. Coatings
1992, 20 (2), 105e113.
67. Yuan, S.; Chen, W.; Hu, S. Fabrication of TiO2 Nanoparticles/Surfactant Polymer Complex Film on
Glassy Carbon Electrode and Its Application to Sensing Trace Dopamine. Mater. Sci. Eng. C 2005,
25 (4), 479e485.
68. Arnal, P.; Corriu, R. J. P.; Leclercq, D.; Mutin, P. H.; Vioux, A. A Solution Chemistry Study of
Nonhydrolytic SoleGel Routes to Titania. Chem. Mater. 1997, 9 (3), 694e698.
69. Sugimoto, T.; Zhou, X.; Muramatsu, A. Synthesis of Uniform Anatase TiO2 Nanoparticles by GeleSol
Method: 1. Solution Chemistry of Ti(OH)(4n)þ
Complexes. J. Colloid Interface Sci. 2002, 252 (2), 339e346.
n
70. Trung, T.; Cho, W. J.; Ha, C. S. Preparation of TiO2 Nanoparticles in Glycerol-Containing Solutions.
Mater. Lett. 2003, 57 (18), 2746e2750.
71. Kim, C. S.; Moon, B. K.; Park, J. H.; Chung, S. T.; Son, S. M. Synthesis of Nanocrystalline TiO2 in
Toluene by a Solvothermal Route. J. Cryst. Growth 2003, 254 (3e4), 405e410.
72. Nian, J. N.; Teng, H. Hydrothermal Synthesis of Single-Crystalline Anatase TiO2 Nanorods with
Nanotubes as the Precursor. J. Phys. Chem. B 2006, 110 (9), 4193e4198.
73. Kolen’ko, Y. V.; Churagulov, B. R.; Kunst, M.; Mazerolles, L.; Colbeau-Justin, C. Photocatalytic
Properties of Titania Powders Prepared by Hydrothermal Method. Appl. Catal. B Environ. 2004, 54 (1),
51e58.
74. Xiaolan, C.; Xiaofei, W. Preparation of the Al-CNT (Carbon Nanotubes) Compound Material by High
Energy Milling. Curr. Nanosci. 2012, 8 (1), 33e37.
75. Kamei, M.; Mitsuhashi, T. Hydrophobic Drawings on Hydrophilic Surfaces of Single Crystalline Titanium
Dioxide: Surface Wettability Control by Mechanochemical Treatment. Surf. Sci. 2000, 463 (1),
L609eL612.
Silica and Titania Nanodispersions 203
76. Guimarães, J. L.; Abbate, M.; Betim, S. B.; Alves, M. C. M. Preparation and Characterization of
TiO2 and V2O5 Nanoparticles Produced by Ball-Milling. J. Alloys Compd. 2003, 352 (1e2),
16e20.
77. Oh, S. M.; Ishigaki, T. Preparation of Pure Rutile and Anatase TiO2 Nanopowders Using RF Thermal
Plasma. Thin Solid Films 2004, 457 (1), 186e191.
78. Matsubara, M.; Yamaki, T.; Itoh, H.; Abe, H.; Asai, K. Preparation of TiO2 Nanoparticles by Pulsed
Laser Ablation: Ambient Pressure Dependence of Crystallization. Jpn. J. Appl. Phys. Part 2 Lett. 2003,
42 (5 A), L479eL481.
79. Zhang, T.; Bao, Y.; Gawne, D. T.; Liu, B.; Karwattzki, J. Computer Model to Simulate the Random
Behaviour of Particles in a Thermal-Spray Jet. Surf. Coatings Technol. 2006, 201 (6), 3552e3563.
80. Sanchez, C.; In, M. Molecular Design of Alkoxide Precursors for the Synthesis of Hybrid
OrganiceInorganic Gels. J. Non-Cryst. Solids 1992, 147e148 (0), 1e12.
81. Sanchez, C.; Livage, J.; Henry, M.; Babonneau, F. Chemical Modification of Alkoxide Precursors.
J Non-Cryst. Solids 1988, 100 (1e3), 65e76.
82. Livage, J.; Henry, M.; Sanchez, C. SoleGel Chemistry of Transition Metal Oxides. Prog. Solid State
Chem. 1988, 18 (4), 259e341.
83. Watthanaarun, J.; Pavarajarn, V.; Supaphol, P. Titanium (IV) Oxide Nanofibers by Combined SoleGel and
Electrospinning Techniques: Preliminary Report on Effects of Preparation Conditions and Secondary
Metal Dopant. Sci. Technol. Adv. Mater. 2005, 6 (3e4 special issue), 240e245.
84. Bischoff, B. L.; Anderson, M. A. Peptization Process in the SoleGel Preparation of Porous Anatase
(TiO2). Chem. Mater. 1995, 7 (10), 1772e1778.
85. Hunziker, J. C.; Frey, M.; Clauer, N.; Dallmeyer, R. D.; Friedrichsen, H.; Flehmig, W.; Hochstrasser, K.;
Roggwiler, P.; Schwander, H. The Evolution of Illite to Muscovite: Mineralogical and Isotopic Data from
the Glarus Alps, Switzerland. Contrib. Mineral. Petrol. 1986, 92 (2), 157e180.
86. Shannon, R. D.; Pask, J. A. Kinetics of the Anatase-Rutile Transformation. J. Am. Ceram. Soc. 1965,
48 (8), 391e398.
87. Rao, C. N. R.; Yoganarasimhan, S. R.; Faeth, P. A. Studies on the Brookite-Rutile Transformation. Trans.
Faraday Soc. 1961, 57, 504e510.
88. Kikkawa, H.; O’Regan, B.; Anderson, M. A. The Photoelectrochemical Properties of Nb-Doped TiO2
Semiconducting Ceramic Membrane. J. Electroanal. Chem. Interfacial Electrochem. 1991, 309 (1e2),
91e101.
89. Sanchez, C.; Ribot, F.; Rozes, L.; Alonso, B. Design of Hybrid OrganiceInorganic Nanocomposites
Synthesized Via SoleGel Chemistry. Mol. Cryst. Liq. Cryst. Sci. Technol. Sect. A Mol. Cryst. Liq. Cryst.
2000, 354, 143e158.
90. Barringer, E. A.; Bowen, H. K. High-Purity, Monodisperse TiO2 Powders by Hydrolysis of Titanium
Tetraethoxide. A. Synthesis and Physical Properties. Langmuir 1985, 1 (4), 414e420.
91. Vorkapic, D.; Matsoukas, T. Effect of Temperature and Alcohols in the Preparation of Titania
Nanoparticles from Alkoxides. J. Am. Ceram. Soc. 1998, 81 (11), 2815e2820.
92. Ding, X. Z.; Qi, Z. Z.; He, Y. Z. Effect of Hydrolysis Water on the Preparation of Nano-Crystalline
Titania Powders via a SoleGel Process. J. Mater. Sci. Lett. 1995, 14 (1), 21e22.
93. Jean, J. H.; Ring, T. A. Nucleation and Growth of Monosized TiO2 Powders from Alcohol Solution.
Langmuir 1986, 2 (2), 251e255.
94. Look, J. L.; Zukoski, C. F. Alkoxide-Derived Titania Particles: Use of Electrolytes to Control Size and
Agglomeration Levels. J. Am. Ceram. Soc. 1992, 75 (6), 1587e1595.
95. Look, J. L.; Zukoski, C. F. Colloidal Stability and Titania Precipitate Morphology: Influence of
Short-Range Repulsions. J. Am. Ceram. Soc. 1995, 78 (1), 21e32.
96. Lijzenga, C.; Zaspalis, V. T.; Ransijn, C. D.; Kumar, K. P.; Keizer, K.; Burggraaf, A. J. Nanostructure
Characterization of Titania Membranes. Key Eng. Mater. 1991, 61-62, 379e382.
97. Anderson, M. A.; Gieselmann, M. J.; Xu, Q. Titania and Alumina Ceramic Membranes. J. Membr. Sci.
1988, 39 (3), 243e258.
204 Chapter 5
98. Xu, Q.; Gieselmann, M. J.; Anderson, M. A. Colloid Chemistry of Ceramic Membranes, Polymeric
Materials Science and Engineering. In Proceedings of the ACS Division of Polymeric Materials Science
and Engineering; 1989; pp 889e893.
99. Brinker, C. J. Hydrolysis and Condensation of Silicates: Effects on Structure. J. Non-Cryst. Solids 1988,
100 (1e3), 31e50.
100. Rozes, L.; Steunou, N.; Fornasieri, G.; Sanchez, C. Titanium-Oxo Clusters, Versatile Nanobuilding Blocks
for the Design of Advanced Hybrid Materials. Monatsh. Chem. 2006, 137 (5), 501e528.
101. Blanchard, J.; Ribot, F.; Sanchez, C.; Bellot, P. V.; Trokiner, A. Structural Characterization of
Titanium-Oxo-Polymers Synthesized in the Presence of Protons or Complexing Ligands as Inhibitors.
J. Non-Cryst. Solids 2000, 265 (1), 83e97.
102. Soloviev, A.; Ivanov, D.; Tufeu, R.; Kanaev, A. V. Nanoparticle Growth During the Induction Period of
the SoleGel Process. J. Mater. Sci. Lett. 2001, 20 (10), 905e906.
103. Simonsen, M. E.; Søgaard, E. G. Sol-gel reactions of titanium alkoxides and water: Influence of pH and
alkoxy group on cluster formation and properties of the resulting products. J. Sol-Gel Science Tech. 2010,
53 (3), 485e497.
104. Nagpal, V. J.; Davis, R. M.; Riffle, J. S. In Situ Steric Stabilization of Titanium Dioxide Particles
Synthesized by a SoleGel Process. Colloids Surf. A 1994, 87 (1), 25e31.
105. Jean, J.-H.; Ring, T. A. Processing Monosized TiO2 Powders Generated with HPC Dispersant. Am.
Ceram. Soc. Bull. 1986, 65 (12), 1574e1577.
106. Harris, M. T.; Byers, C. H. Effect of Solvent on the Homogeneous Precipitation of Titania by Titanium
Ethoxide Hydrolysis. J. Non-Cryst. Solids 1988, 103 (1), 49e64.
107. Watanabe, H.; Kimura, T.; Yamaguchi, T. Sintering of Platelike Bismuth Titanate Powder Compacts with
Preferred Orientation. J. Am. Ceram. Soc. 1991, 74 (1), 139e147.
108. Day, V. W.; Eberspacher, T. A.; Frey, M. H.; Klemperer, W. G.; Liang, S.; Payne, D. A. Barium Titanium
Glycolate: A New Barium Titanate Powder Precursor. Chem. Mater. 1996, 8 (2), 330e332.
109. Babonneau, F.; Coury, L.; Livage, J. Aluminum Sec-Butoxide Modified with Ethylacetoacetate: An Attractive
Precursor for the SoleGel Synthesis of Ceramics. J. Non-Cryst. Solids 1990, 121 (1e3), 153e157.
110. Léaustic, A.; Babonneau, F.; Livage, J. Structural Investigation of the Hydrolysis-Condensation Process of
Titanium Alkoxides Ti(OR)4 (OR ¼ OPri, OEt) modified by Acetylacetone. 1. Study of the Alkoxide
Modification. Chem. Mater. 1989, 1 (2), 240e247.
111. Bahnemann, D.; Henglein, A.; Spanhel, L. Detection of the Intermediates of Colloidal TiO2-Catalysed
Photoreactions. Faraday Discuss. Chem. Soc. 1984, 78, 151e163.
112. Su, C.; Hong, B. Y.; Tseng, C. M. SoleGel Preparation and Photocatalysis of Titanium Dioxide. Catal.
Today 2004, 96 (3), 119e126.
113. Bartlett, J. R.; Gazeau, D.; Zemb, T.; Woolfrey, J. L. Structure of Multicomponent (Titania/Zirconia)
Colloids. Langmuir 1998, 14 (13), 3538e3544.
114. Friberg, S. E.; Ma, Z. Hydrolysis of Tetraethoxysilane in a Liquid Crystal, in a Microemulsion and in a
Solution. J. Non-Cryst. Solids 1992, 147e148 (C), 30e35.
115. Friberg, S. E.; Ahmed, A. U.; Yang, C. C.; Ahuja, S.; Bodesha, S. S. Gelation of a Microemulsion by
Silica Formed in situ. J Mater Chem 1992, 2 (2), 257e258.
116. Boutonnet, M.; Kizling, J.; Stenius, P.; Maire, G. The Preparation of Monodisperse Colloidal Metal
Particles from Microemulsions. Colloids Surf. 1982, 5 (3), 209e225.
117. Kurihara, K.; Kizling, J.; Stenius, P.; Fendler, J. H. Laser and Pulse Radiolytically Induced Colloidal Gold
Formation in Water and in Water-in-Oil Microemulsions. J. Am. Chem. Soc. 1983, 105 (9), 2574e2579.
118. Kandori, K.; Kishi, K.; Ishikawa, T. Preparation of Uniform Silica Hydrogel Particles by SPG Filter
Emulsification Method. Colloids Surf. 1992, 62 (3), 259e262.
119. Ayyub, P.; Maitra, A. N.; Shah, D. O. Formation of Theoretical-Density Microhomogeneous YBa2Cu3O7x
Using a Microemulsion-Mediated Process. Phys C 1990, 168 (5e6), 571e579.
120. Robinson, B. H.; Towey, T. F.; Zourab, S.; Visser, A. J. W. G.; van Hoek, A. Characterisation of
Cadmium Sulphide Colloids in Reverse Micelles. Colloids Surf. 1991, 61 (C), 175e188.
Silica and Titania Nanodispersions 205
121. López-Quintela, M. A.; Rivas, J. Chemical Reactions in Microemulsions: A Powerful Method to Obtain
Ultrafine Particles. J. Colloid Interface Sci. 1993, 158 (2), 446e451.
122. Lisiecki, I.; Pileni, M. P. Synthesis of Copper Metallic Clusters Using Reverse Micelles as Microreactors.
J. Am. Chem. Soc. 1993, 115 (10), 3887e3896.
123. Yamauchi, H.; Ishikawa, T.; Kondo, S. Surface Characterization of Ultramicro Spherical Particles of
Silica Prepared by W/O Microemulsion Method. Colloids Surfaces 1989, 37 (C), 71e80.
124. Esquena, J.; Tadros, T. F.; Kostarelos, K.; Solans, C. Preparation of Narrow Size Distribution Silica
Particles Using Microemulsions. Langmuir 1997, 13 (24), 6400e6406.
125. Silber, J. J.; Biasutti, A.; Abuin, E.; Lissi, E. Interactions of Small Molecules with Reverse Micelles. Adv.
Colloid Interface Sci. 1999, 82 (1), 189e252.
126. Cushing, B. L.; Kolesnichenko, V. L.; O’Connor, C. J. Recent Advances in the Liquid-Phase Syntheses of
Inorganic Nanoparticles. Chem. Rev. 2004, 104 (9), 3893e3946.
127. Holmberg, K. Surfactant-Templated Nanomaterials Synthesis. J. Colloid Interface Sci. 2004, 274 (2),
355e364.
128. López-Quintela, M. A.; Tojo, C.; Blanco, M. C.; Garcı́a Rio, L.; Leis, J. R. Microemulsion Dynamics and
Reactions in Microemulsions. Curr. Opin. Colloid Interface Sci. 2004, 9 (3e4), 264e278.
129. Eriksson, S.; Nylén, U.; Rojas, S.; Boutonnet, M. Preparation of Catalysts from Microemulsions and
Their Applications in Heterogeneous Catalysis. Appl. Catal. A 2004, 265 (2), 207e219.
130. Capek, I. Preparation of Metal Nanoparticles in Water-in-Oil (W/O) Microemulsions. Adv. Colloid
Interface Sci. 2004, 110 (1e2), 49e74.
131. Lazzari, M.; Arturo López-Quintela, M. Block Copolymers as a Tool for Nanomaterial Fabrication. Adv.
Mater. 2003, 15 (19), 1583e1594.
132. Osseo-Asare, K.; Arriagada, F. J. Preparation of SiO2 Nanoparticles in a Non-Ionic Reverse Micellar
System. Colloids Surf. 1990, 50 (C), 321e339.
133. Arriagada, F. J.; Osseo-Asare, K. Silica Nanoparticles Produced in Aerosol OT Reverse Microemulsions:
Effect of Benzyl Alcohol on Particle Size and Polydispersity. J. Dispersion Sci. Technol. 1994, 15 (1),
59e71.
134. Friberg, S. E.; Sjöblom, J. In Industrial Applications of Microemulsions; Solans, C., Kunieda, H., Eds.;
Marcel Dekker: New York, 1997.
135. Sjöblom, E.; Henriksson, U. In Surfactants in Solution, Vol. 3, Mittal, K. L., Lindman, B., Eds.; Plenum
Press: New York, 1984; p 1987.
136. Di Meglio, M.; Dvolaitzky, M.; Taupin, C. In Microemulsion Systems; Rosano, H. L., Clause, M., Eds.;
Marcel Dekker: New York, 1987.
137. Bellocq, A. M.; Biais, J.; Bothorel, P.; Clin, B.; Fourche, G.; Lalanne, P.; Lemaire, B.; Lemanceau, B.;
Roux, D. Microemulsions. Adv. Colloid Interface Sci. 1984, 20 (3e4), 167e272.
138. Roux, D.; Bellocq, A. M.; Bothorel, P. Effect of the Molecular Structure of Components on Micellar
Interactions in Microemulsions. Prog. Colloid Polym. Sci. 1984, 69, 1e11.
139. Tojo, C.; Blanco, M. C.; López-Quintela, M. A. The Influence of Reactant Excess and Film Flexibility on
the Mechanism of Nanoparticle Formation in Microemulsions: A Monte Carlo Simulation. Langmuir
1998, 14 (24), 6835e6839.
140. Ramesh Kumar, A.; Hota, G.; Mehra, A.; Khilar, K. C. Modeling of Nanoparticles Formation by Mixing
of Two Reactive Microemulsions. AIChE J. 2004, 50 (7), 1556e1567.
141. Quintillán, S.; Tojo, C.; Blanco, M. C.; López-Quintela, M. A. Effects of the Intermicellar Exchange on
the Size Control of Nanoparticles Synthesized in Microemulsions. Langmuir 2001, 17 (23), 7251e7254.
142. Hirai, T.; Sato, H.; Komasava, I. Mechanism of Formation of Titanium Dioxide Ultrafine Particles in
Reverse Micelles by Hydrolysis of Titanium Tetrabutoxide. Ind. Eng. Chem. Res. 1993, 32 (12),
3014e3019.
143. Lee, M. S.; Park, S. S.; Lee, G. D.; Ju, C. S.; Hong, S. S. Synthesis of TiO2 Particles by Reverse
Microemulsion Method Using Nonionic Surfactants with Different Hydrophilic and Hydrophobic Group
and Their Photocatalytic Activity. Catal. Today 2005, 101 (3e4 special issue), 283e290.
206 Chapter 5
144. Lee, G. W. J.; Tadros, T. F. Formation and Stability of Emulsions Produced by Dilution of Emulsifiable
Concentrates. Part I. An Investigation of the Dispersion on Dilution of Emulsifiable Concentrates
Containing Cationic and Non-Ionic Surfactants. Colloids Surf. 1982, 5 (2), 105e115.
145. Forgiarini, A.; Esquena, J.; González, C.; Solans, C. Formation of Nano-Emulsions by Low-Energy
Emulsification Methods at Constant Temperature. Langmuir 2001, 17 (7), 2076e2083.
146. Nakajima, H. In Industrial Applications of Microemulsions; Solans, C., Kunieda, H., Eds.; Marcel
Dekker: New York, 1996; pp 157e197.
147. Sagitani, H. In Organized Solutions; Friberg, S. E., Lindman, B., Eds.; Marcel Dekker: New York, 1992;
pp 259e271.
148. Sudol, E. D.; El-Aasser, M. S. In Emulsion Polymerization and Emulsion Polymers; Lovell, P. A.,
El-Aasser, M. S., Eds.; Wiley & Sons: New York, 1997; pp 699e722.
149. Porras, M.; Martı́nez, A.; Solans, C.; González, C.; Gutiérrez, J. M. Ceramic Particles Obtained Using
W/O Nano-Emulsions as Reaction Media. Colloids Surf. A 2005, 270e271 (1e3), 189e194.
150. Wu, Z.; Wu, Q.; Du, L.; Jiang, C.; Piao, L. Progress in the Synthesis and Applications of Hierarchical
Flower-Like TiO2 Nanostructures. Particuology 2014, 15, 61e70.
151. Graham, T. On the Properties of Silicic Acid and Other Analogous Colloidal Substances. J. Chem. Soc.
1864, 17 (0), 318e327.
152. Sanchez, M. G. Canadian Patent, 586, 261, 1959.
153. White, J.F., US Patent 2, 375, 1945.
154. Alexander, G. B.; Iler, R. K. US Patent, 2.601.235, 1952.
155. Alexander, G.B. U.S. Patent 2,750,345, 1956.
156. Lim, H.; Lee, J.; Jeong, J.; Oh, S.; Lee, S. Comparative Study of Various Preparation Methods of
Colloidal Silica. Engineering 2010, 2 (12), 998e1005.
157. Ciriminna, R.; Fidalgo, A.; Pandarus, V.; Béland, F.; Ilharco, L. M.; Pagliaro, M. The SoleGel Route to
Advanced Silica-Based Materials and Recent Applications. Chem. Rev. 2013, 113 (8), 6592e6620.
158. Wang, H.; Liu, P.; Cheng, X.; Shui, A.; Zeng, L. Effect of Surfactants on Synthesis of TiO2
Nano-Particles by Homogeneous Precipitation Method. Powder Technol. 2008, 188 (1), 52e54.
159. Park, H. K.; Park, K. Y. Vapor-Phase Synthesis of Uniform Silica Spheres Through Two-Stage Hydrolysis
of SiCl4. Mater. Res. Bull. 2008, 43 (11), 2833e2839.
160. Chen, X.; Jiang, J.; Yan, F.; Tian, S.; Li, K. A Novel Low Temperature Vapor Phase Hydrolysis Method
for the Production of Nano-Structured Silica Materials Using Silicon Tetrachloride. RSC Adv. 2014,
4 (17), 8703e8710.
161. Zhang, D. L.; Liu, Z. G.; Koch, C. C. Applications of High Energy Mechanical Milling in Processing
Advanced Materials: An Overview of the Understanding. J. Metastable Nanocryst. Mater. 2004, 20e21,
103e110.
162. Zhang, D. L. Processing of Advanced Materials Using High-Energy Mechanical Milling. Prog. Mater. Sci.
2004, 49 (3e4), 537e560.
163. Tsuzuki, T.; McCormick, P. G. Synthesis of Ultrafine Ceria Powders by Mechanochemical Processing.
J. Am. Ceram. Soc. 2001, 84 (7), 1453e1458.
164. Tsuzuki, T.; McCormick, P. G. Mechanochemical Synthesis of Nanoparticles. J. Mater. Sci. 2004,
39 (16e17), 5143e5146.
165. Tsuzuki, T.; Pirault, E.; McCormick, P. G. Mechanochemical Synthesis of Gadolinium Oxide
Nanoparticles. Nanostruct. Mater. 1999, 11 (1), 125e131.
166. Ding, J.; Miao, W. F.; McCormick, P. G.; Street, R. Mechanochemical Synthesis of Ultrafine Fe Powder.
Appl. Phys. Lett. 1995, 67, 3804.
167. Cot, F.; Larbot, A.; Nabias, G.; Cot, L. Preparation and Characterization of Colloidal Solution Derived
Crystallized Titania Powder. J. Eur. Ceram. Soc. 1998, 18 (14), 2175e2181.
168. Yang, J.; Mei, S.; Ferreira, J. M. F. Hydrothermal Synthesis of Nanosized Titania Powders: Influence of
Peptization and Peptizing Agents on the Crystalline Phases and Phase Transitions. J. Am. Ceram. Soc.
2000, 83 (6), 1361e1368.
Silica and Titania Nanodispersions 207
169. Feng, X.; Zhai, J.; Jiang, L. The Fabrication and Switchable Superhydrophobicity of TiO2 Nanorod Films.
Angew. Chem. Int. Ed. 2005, 44 (32), 5115e5118.
170. Bavykin, D. V.; Gordeev, S. N.; Moskalenko, A. V.; Lapkin, A. A.; Walsh, F. C. Apparent
Two-Dimensional Behavior of TiO2 Nanotubes Revealed by Light Absorption and Luminescence. J. Phys.
Chem. B 2005, 109 (18), 8565e8569.
171. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Formation of Titanium Oxide Nanotube.
Langmuir 1998, 14 (12), 3160e3163.
172. Wang, X.; Zhuang, J.; Peng, Q.; Li, Y. A general Strategy for Nanocrystal Synthesis. Nature 2005,
437 (7055), 121e124.
173. Yang, S.; Gao, L. Fabrication and Characterization of Nanostructurally Flowerlike Aggregates of TiO2 via
a Surfactant-Free Solution Route: Effect of Various Reaction Media. Chem. Lett. 2005, 34 (7),
1044e1045.
174. Matijevic, E.; Budnik, M.; Meites, L. Preparation and Mechanism of Formation of Titanium Dioxide
Hydrosols of Narrow Size distribution. J. Colloid Interface Sci. 1977, 61 (2), 302e311.
175. Banerjee, S.; Das, J.; Alvarez, R. P.; Santra, S. Silica Nanoparticles as a Reusable Catalyst: A
Straightforward Route for the Synthesis of Thioethers, Thioesters, Vinyl Thioethers and Thio-Michael
Adducts Under Neutral Reaction Conditions. New J. Chem. 2010, 34 (2), 302e306.
176. Kataoka, K.; Harada, A.; Nagasaki, Y. Block Copolymer Micelles for Drug Delivery: Design,
Characterization and Biological Significance. Adv. Drug Deliv. Rev. 2001, 47 (1), 113e131.
177. Gill, I.; Ballesteros, A. Encapsulation of Biologicals within Silicate, Siloxane, and Hybrid SoleGel
Polymers: An Efficient and Generic Approach. J. Am. Chem. Soc. 1998, 120 (34), 8587e8598.
178. Boninsegna, S.; Bosetti, P.; Carturan, G.; Dellagiacoma, G.; Dal Monte, R.; Rossi, M. Encapsulation of
Individual Pancreatic Islets by SoleGel SiO2: A Novel Procedure for Perspective Cellular Grafts.
J. Biotechnol. 2003, 100 (3), 277e286.
179. Binks, B. P.; Horozov, T. S. Colloidal Particles at Liquid Interfaces; University Press: Cambridge, 2006.
180. Giermanska-Kahn, J.; Schmitt, V.; Binks, B. P.; Leal-Calderon, F. A New Method to Prepare
Monodisperse Pickering Emulsions. Langmuir 2002, 18 (7), 2515e2518.
181. Dulle, M.; Glatter, O. Internally Self-Assembled Submicrometer Emulsions Stabilized with a Charged
Polymer or with Silica Particles. Langmuir 2012, 28 (2), 1136e1141.
182. Sadeghpour, A.; Pirolt, F.; Glatter, O. Submicrometer-Sized Pickering Emulsions Stabilized by Silica
Nanoparticles with Adsorbed Oleic Acid. Langmuir 2013, 29 (20), 6004e6012.
183. Senic, Z.; Bauk, S.; Vitorovic, M.; Pajic, N.; Samolov, A.; Rajic, D. Application of TiO2 Nanoparticles
for Obtaining Self Decontaminating Smart Textiles. Sci. Tech. Rev. 2011, 61 (3e4), 63e72.
Bauk, S.; Simic, D.; Vitorovic-Todorovic, M.; Markovic, T.; Radojkovic, A.; et al. The
184. Senic, Z.;
Preliminary Comparative Analysis of Different Routes for TiO2 Nanoparticles Synthesis and Their
Deposition on Textiles. The Methyl-Orange Degradation and VX Detoxication Study. Dig. J. Nanomater.
Biostructures 2013, 8 (2), 711e719.
185. Sundström, V. Light in Elementary Biological Reactions. Prog. Quantum Electron. 2000, 24 (5),
187e238.
186. Maliakal, A.; Katz, H.; Cotts, P. M.; Subramoney, S.; Mirau, P. Inorganic Oxide Core, Polymer Shell
Nanocomposite as a High K Gate Dielectric for Flexible Electronics Applications. J. Am. Chem. Soc.
2005, 127 (42), 14655e14662.
187. Majewski, L. A.; Schroeder, R.; Grell, M. Low-Voltage, High-Performance Organic Field-Effect
Transistors with an Ultra-Thin TiO2 Layer as Gate Insulator. Adv. Funct. Mater. 2005, 15 (6),
1017e1022.
188. Shrestha, N. K.; Macak, J. M.; Schmidt-Stein, F.; Hahn, R.; Mierke, C. T.; Fabry, B.; Schmuki, P.
Magnetically Guided Titania Nanotubes for Site-Selective Photocatalysis and Drug Release. Angew.
Chem. Int. Ed. 2009, 48 (5), 969e972.
189. Song, Y. Y.; Schmidt-Stein, F.; Bauer, S.; Schmuki, P. Amphiphilic TiO2 Nanotube Arrays: An Actively
Controllable Drug Delivery System. J. Am. Chem. Soc. 2009, 131 (12), 4230e4232.
208 Chapter 5
190. Cai, R.; Kubota, Y.; Shuin, T.; Sakai, H.; Hashimoto, K.; Fujishima, A. Induction of Cytotoxicity by
Photoexcited TiO2 Particles. Cancer Res. 1992, 52 (8), 2346e2348.
191. Heinlaan, M.; Ivask, A.; Blinova, I.; Dubourguier, H. C.; Kahru, A. Toxicity of Nanosized and Bulk ZnO,
CuO and TiO2 to Bacteria Vibrio fischeri and Crustaceans Daphnia magna and Thamnocephalus
platyurus. Chemosphere 2008, 71 (7), 1308e1316.
192. Zhu, Y.; Shi, J.; Zhang, Z.; Zhang, C.; Zhang, X. Development of a Gas Sensor Utilizing
Chemiluminescence on Nanosized Titanium Dioxide. Anal. Chem. 2002, 74 (1), 120e124.
193. Varghese, O. K.; Gong, D.; Paulose, M.; Ong, K. G.; Grimes, C. A. Hydrogen Sensing Using Titania
Nanotubes. Sens. Actuators B 2003, 93 (1e3), 338e344.
194. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped
Titanium Oxides. Science 2001, 293 (5528), 269e271.
195. Choi, W.; Termin, A.; Hoffmann, M. R. The Role of Metal Ion Dopants in Quantum-Sized TiO2:
Correlation Between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 1994,
98 (51), 13669e13679.
196. Djerdj, I.; Arcon, D.; Jaglicic, Z.; Niederberger, M. Nonaqueous Synthesis of Metal Oxide Nanoparticles:
Short Review and Doped Titanium Dioxide as Case Study for the Preparation of Transition Metal-Doped
Oxide Nanoparticles. J. Solid State Chem. 2008, 181 (7), 1571e1581.
197. Pajonk, G. M. Catalytic Aerogels. Catal. Today 1997, 35 (3), 319e337.
198. Maira, A. J.; Yeung, K. L.; Soria, J.; Coronado, J. M.; Belver, C.; Lee, C. Y.; Augugliaro, V. Gas-Phase
Photo-Oxidation of Toluene Using Nanometer-Size TiO2 Catalysts. Appl. Catal. B 2001, 29 (4), 327e336.
199. Maira, A. J.; Yeung, K. L.; Lee, C. Y.; Yue, P. L.; Chan, C. K. Size Effects in Gas-Phase Photo-Oxidation
of Trichloroethylene Using Nanometer-Sized TiO2 Catalysts. J. Catal. 2000, 192 (1), 185e196.
200. Gao, L.; Zhang, Q. Effects of Amorphous Contents and Particle Size on the Photocatalytic Properties of
TiO2 Nanoparticles. Scr. Mater. 2001, 44 (8e9), 1195e1198.
201. Leung, D. Y. C.; Fu, X.; Wang, C.; Ni, M.; Leung, M. K. H.; Wang, X.; Fu, X. Hydrogen Production
Over Titania-Based Photocatalysts. ChemSusChem 2010, 3 (6), 681e694.
202. Chen, X.; Mao, S. S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications and
Applications. Chem. Rev. 2007, 107 (7), 2891e2959.
203. Tanaka, K.; Capule, M. F. V.; Hisanaga, T. Effect of Crystallinity of TiO2 on Its Photocatalytic Action.
Chem. Phys. Lett. 1991, 187 (1e2), 73e76.
204. Fox, M. A.; Dulay, M. T. Heterogeneous Photocatalysis. Chem. Rev. 1993, 93 (1), 341e357.
205. Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W. Environmental Applications of
Semiconductor Photocatalysis. Chem. Rev. 1995, 95 (1), 69e96.
206. Linsebigler, A. L.; Lu, G.; Yates, J. T., Jr. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and
Selected Results. Chem. Rev. 1995, 95 (3), 735e758.
207. Guohui, T.; Honggang, F.; Liqiang, J.; Baifu, X.; Kai, P. Preparation and Characterization of Stable
Biphase TiO2 Photocatalyst with High Crystallinity, Large Surface Area, and Enhanced Photoactivity.
J. Phys. Chem. C 2008, 112 (8), 3083e3089.
208. Banerjee, S.; Dionysiou, D. D.; Pillai, S. C. Self-Cleaning Applications of TiO2 by Photo-Induced
Hydrophilicity and Photocatalysis. Appl. Catal. B 2015, 176e177, 396e428.
209. Qi, K.; Daoud, W. A.; Xin, J. H.; Mak, C. L.; Tang, W.; Cheung, W. P. Self-Cleaning Cotton. J. Mater.
Chem. 2006, 16 (47), 4567e4574.
210. Takata, Y.; Hidaka, S.; Cao, J. M.; Nakamura, T.; Yamamoto, H.; Masuda, M.; Ito, T. Effect of Surface
Wettability on Boiling and Evaporation. Energy 2005, 30 (2e4), 209e220.
211. Parkin, I. P.; Palgrave, R. G. Self-Cleaning Coatings. J. Mater. Chem. 2005, 15 (17), 1689e1695.
212. Bondioli, F.; Taurino, R.; Ferrari, A. M. Functionalization of Ceramic Tile Surface by SoleGel
Technique. J. Colloid Interface Sci. 2009, 334 (2), 195e201.
213. Andrews, R. W.; Pollard, A.; Pearce, J. M. A New Method to Determine the Effects of Hydrodynamic
Surface Coatings on the Snow Shedding Effectiveness of Solar Photovoltaic Modules. Sol. Energy Mater.
Sol. Cells 2013, 113, 71e78.
Silica and Titania Nanodispersions 209
214. Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Dye-Sensitized Solar Cells. Chem. Rev.
2010, 110 (11), 6595e6663.
215. Bouclé, J.; Ackermann, J. Solid-State Dye-Sensitized and Bulk Heterojunction Solar Cells Using TiO2 and
ZnO Nanostructures: Recent Progress and New Concepts at the Borderline. Polym. Int. 2012, 61 (3), 355e373.
216. Grätzel, M. Recent Advances in Sensitized Mesoscopic Solar Cells. Accounts Chem. Res. 2009, 42 (11),
1788e1798.
217. Fujishima, A.; Zhang, X.; Tryk, D. A. TiO2 Photocatalysis and Related Surface Phenomena. Surf. Sci.
Rep. 2008, 63 (12), 515e582.
218. Wong, H. C.; Cabral, J. T. Nanoparticle Aggregation Behaviour in Polymer Nanocomposites: Bulk vs.
Thin Films. J. Phys. 2010, 247.
219. Schmidt, H.; Seiferling, B. Chemistry and Applications of InorganiceOrganic Polymers (Organically
Modified Silicates). In Materials Research Society Symposia Proceedings; 1986; pp 739e750.
220. Wang, B.; Wilkes, G. L.; Smith, C. D.; McGrath, J. E. High Refractive Index Hybrid Ceramer Materials
Prepared from Titanium Tetraisopropoxide and Poly(Arylene Ether Phosphine Oxide) Through SoleGel
Processing. Polym. Commun. Guildf. 1991, 32 (13), 400e402.
221. Templin, M.; Franck, A.; Du Chesne, A.; Leist, H.; Zhang, Y.; Ulrich, R.; Schädler, V.; Wiesner, U.
Organically Modified Aluminosilicate Mesostructures from Block Copolymer Phases. Science 1997,
278 (5344), 1795e1798.
222. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Ordered Mesoporous Molecular
Sieves Synthesized by a LiquideCrystal Template Mechanism. Nature 1992, 359 (6397), 710e712.
223. Ryan, J. V.; Berry, A. D.; Anderson, M. L.; Long, J. W.; Stroud, R. M.; Cepak, V. M.; Browning, V. M.;
Rolison, D. R.; Merzbacher, C. I. Electronic Connection to the Interior of a Mesoporous Insulator with
Nanowires of Crystalline RuO2. Nature 2000, 406 (6792), 169e172.
224 Morris, C. A.; Anderson, M. L.; Stroud, R. M.; Merzbacher, C. I.; Rolison, D. R. Silica Sol as a
Nanoglue: Flexible Synthesis of Composite Aerogels. Science 1999, 284 (5414), 622e624.
225. Watcharotone, S.; Diking, D. A.; Stankovich, S.; Pinery, R.; Jung, I.; Dommett, G. H. B.; Evmenenko, G.;
Wu, S. E.; Chen, S. F.; Liu, C. P.; et al. Graphene-Silica Composite Thin Films as Transparent
Conductors. Nano Lett. 2007, 7 (7), 1888e1892.
226. Warren, S. C.; Perkins, M. R.; Adams, A. M.; Kamperman, M.; Burns, A. A.; Arora, H.; Herz, E.;
Suteewong, T.; Sai, H.; Li, Z.; et al. A Silica SoleGel Design Strategy for Nanostructured Metallic
Materials. Nat. Mater. 2012, 11 (5), 460e467.
227. Rupp, W.; Hüsing, N.; Schubert, U. Preparation of Silica-Titania Xerogels and Aerogels by SoleGel
Processing of New Single-Source Precursors. J. Mater. Chem. 2002, 12 (9), 2594e2596.
228. Luechinger, M.; Kienhöfer, A.; Pirngruber, G. D. Immobilized Complexes of Metals with Amino Acid
Ligands e A First Step Toward the Development of New Biomimetic Catalysts. Chem. Mater. 2006,
18 (5), 1330e1336.
229. Dong, D.; Jiang, S.; Men, Y.; Ji, X.; Jiang, B. Nanostructured Hybrid OrganiceInorganic Lanthanide
Complex Films Produced In Situ via a SoleGel Approach. Adv. Mater. 2000, 12 (9), 646e649.
230. Radin, S.; Chen, T.; Ducheyne, P. The Controlled Release of Drugs from Emulsified, Sol Gel Processed
Silica Microspheres. Biomaterials 2009, 30 (5), 850e858.
231. Ciriminna, R.; Pagliaro, M. SoleGel Microencapsulation of Odorants and Flavors: Opening the Route to
Sustainable Fragrances and Aromas. Chem. Soc. Rev. 2013, 42 (24), 9243e9250.
232. Veith, S. R.; Pratsinis, S. E.; Perren, M. Aroma Retention in SoleGel-Made Silica Particles. J. Agric.
Food Chem. 2004, 52 (19), 5964e5971.
233. Barbé, C.; Bartlett, J.; Kong, L.; Finnie, K.; Lin, H. Q.; Larkin, M.; Calleja, S.; Bush, A.; Calleja, G.
Silica Particles: A Novel Drug-Delivery System. Adv. Mater. 2004, 16 (21), 1959e1966.
234. Barbé, C. J.; Kong, L.; Finnie, K. S.; Calleja, S.; Hanna, J. V.; Drabarek, E.; Cassidy, D. T.;
Blackford, M. G. SoleGel Matrices for Controlled Release: From Macro to Nano Using Emulsion
Polymerisation. J. Sol-Gel Sci. Technol. 2008, 46 (3), 393e409.
210 Chapter 5
235. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects.
Jpn. J. Appl. Phys. Part 1 2005, 44 (12), 8269e8285.
236. Pottier, A.; Cassaignon, S.; Chanéac, C.; Villain, F.; Tronc, E.; Jolivet, J. P. Size Tailoring of TiO2
Anatase Nanoparticles in Aqueous Medium and Synthesis of Nanocomposites. Characterization by Raman
Spectroscopy. J. Mater. Chem. 2003, 13 (4), 877e882.
237. Abbas, Z.; Holmberg, J. P.; Hellström, A. K.; Hagström, M.; Bergenholtz, J.; Hassellöv, M.; Ahlberg, E.
Synthesis, Characterization and Particle Size Distribution of TiO2 Colloidal Nanoparticles. Colloids Surf.
A 2011, 384 (1e3), 254e261.
238. Chuang, H. Y.; Chen, D. H. Catalyst-Free Low Temperature Synthesis of Discrete Anatase Titanium
Dioxide Nanocrystals with Highly Thermal Stability and UVC-Cut Capability. J. Nanoparticle Res. 2008,
10 (Suppl. 1), 233e241.
239. Dhage, S. R.; Choube, V. D.; Samuel, V.; Ravi, V. Synthesis of Nanocrystalline TiO2 at 100 C. Mater.
Lett. 2004, 58 (17e18), 2310e2313.
240. Hart, J. N.; Bourgeois, L.; Cervini, R.; Cheng, Y. B.; Simon, G. P.; Spiccia, L. Low Temperature Crystallization
Behavior of TiO2 Derived from a SoleGel Process. J. Sol-Gel Sci. Technol. 2007, 42 (2), 107e117.
241. Li, G.; Li, L.; Boerio-Goates, J.; Woodfield, B. F. High Purity Anatase TiO2 Nanocrystals: Near
Room-Temperature Synthesis, Grain Growth Kinetics, and Surface Hydration Chemistry. J. Am. Chem.
Soc. 2005, 127 (24), 8659e8666.
242. Bellantone, M.; Coleman, N. J.; Hench, L. L. Bacteriostatic Action of a Novel Four-Component Bioactive
Glass. J. Biomed. Mater. Res. 2000, 51 (3), 484e490.
243. Gristina, A. G. Biomaterial-Centered Infection: Microbial Adhesion versus Tissue Integration. Science
1987, 237 (4822), 1588e1595.
244. Anpo, M.; Takeuchi, M. The Design and Development of Highly Reactive Titanium Oxide Photocatalysts
Operating under Visible Light Irradiation. J. Catal. 2003, 216 (1e2), 505e516.
245. Das, K.; Sharma, S. N.; Kumar, M.; De, S. K. Morphology Dependent Luminescence Properties of Co
Doped TiO2 Nanostructures. J. Phys. Chem. C 2009, 113 (33), 14783e14792.
246. Kumar, S. G.; Devi, L. G. Review on Modified TiO2 Photocatalysis under UV/Visible Light: Selected
Results and Related Mechanisms on Interfacial Charge Carrier Transfer Dynamics. J. Phys. Chem. A
2011, 115 (46), 13211e13241.
247. Bhattacharyya, K.; Varma, S.; Tripathi, A. K.; Bharadwaj, S. R.; Tyagi, A. K. Effect of Vanadia Doping
and Its Oxidation State on the Photocatalytic Activity of TiO2 for Gas-Phase Oxidation of Ethene.
J. Phys. Chem. C 2008, 112 (48), 19102e19112.
248. Rockafellow, E. M.; Haywood, J. M.; Witte, T.; Houk, R. S.; Jenks, W. S. Selenium-Modified TiO2 and
Its Impact on Photocatalysis. Langmuir 2010, 26 (24), 19052e19059.
249. Bellardita, M.; Addamo, M.; Di Paola, A.; Palmisano, L. Photocatalytic Behaviour of Metal-Loaded TiO2
Aqueous Dispersions and Films. Chem. Phys. 2007, 339 (1e3), 94e103.
250. Nakano, Y.; Morikawa, T.; Ohwaki, T.; Taga, Y. Origin of Visible-Light Sensitivity in N-Doped TiO2
Films. Chem. Phys. 2007, 339 (1e3), 20e26.
251. Borras, A.; Barranco, A.; González-Elipe, A. R. Reversible Superhydrophobic to Superhydrophilic
Conversion of Ag@TiO2 Composite Nanofiber Surfaces. Langmuir 2008, 24 (15), 8021e8026.
252. Wu, X. F.; Song, H. Y.; Yoon, J. M.; Yu, Y. T.; Chen, Y. F. Synthesis of Core-Shell Au@TiO2
Nanopartides with Truncated Wedge-Shaped Morphology and Their Photocatalytic Properties. Langmuir
2009, 25 (11), 6438e6447.
253. Zhao, G.; Mo, R.; Wang, B.; Zhang, L.; Sun, K. Enhanced cyclability of Li-O2 batteries based on TiO2
supported cathodes with no carbon or binder. Chemistry of Materials 2014, 26 (8), 2551e2556.
254. Teng, T. C.; Teng, T. P.; Pan, S. I. Effects of titania and silica nanoparticle on paint film. Adv. Mat. Res.
2011, 287 (290), 1437e1442.
255. Lingaraju, D.; Ramji, K.; Devi, M. P.; Lakshmi, U. R. Mechanical and tribological studies of polymer
hybrid nanocomposites with nano reinforcements. Bull. Mat. Science 2011, 34 (4), 705e712.
Download