Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

10.1007@s10971 015 3696 2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

J Sol-Gel Sci Technol

DOI 10.1007/s10971-015-3696-2

ORIGINAL PAPER

The influence of addition of a catalyst and chelating agent


on the properties of titanium dioxide synthesized via the sol–gel
method
Katarzyna Siwińska-Stefańska1 • Jakub Zdarta1 • Dominik Paukszta1 •

Teofil Jesionowski1

Received: 15 July 2014 / Accepted: 27 March 2015


Ó The Author(s) 2015. This article is published with open access at Springerlink.com

Abstract Titanium dioxide powders were prepared via using Fourier transform infrared spectroscopy. In addition,
the sol–gel method, using titanium alkoxide as a precursor. for selected TiO2 systems, their photocatalytic activity in
It was investigated how the addition of a catalyst (ammo- the decomposition of C.I. Basic Blue 9 was investigated.
nia) and chelating agent (acetylacetone), as well as the Our research results show that the presence of chelating
temperature of calcination, affects the properties of the agent in the synthesis of titanium dioxide caused significant
resulting TiO2 powder. The physicochemical properties of changes in dispersive properties, crystalline structure and
samples were determined, including the dispersion, mor- porous structure parameters. The novel feature of this work
phology and microstructure of the systems (particle size is the proposed method of synthesis of highly photoactive
distribution, TEM images), crystalline structure (XRD), titanium dioxide with desirable physicochemical properties
characteristics of porous structure (BET), and thermal in the presence of a chelating agent such as acetylacetone.
stability (TGA/DTA). The samples were also analyzed Graphical Abstract

Keywords Titania  Sol–gel method  Chelating agent 


& Katarzyna Siwińska-Stefańska Photocatalytic activity
katarzyna.siwinska-stefanska@put.poznan.pl
1
Faculty of Chemical Technology, Institute of Chemical
Technology and Engineering, Poznan University of
Technology, Berdychowo 4, 60965 Poznan, Poland

123
J Sol-Gel Sci Technol

1 Introduction evolution, porous structure parameters and photocatalytic


capability of the resulting TiO2 powders. To examine the
The synthesis of titanium dioxide is one of the major re- photocatalytic activity of the synthesized titanium dioxide,
search areas in ‘green chemistry.’ Titania is a chemically photodecomposition of C.I. Basic Blue 9 was carried out.
inert, thermally stable, insoluble, biocompatible, non-toxic
material and an excellent absorber of destructive UV ra-
diation. Because of these properties, titanium dioxide has 2 Experimental
for some time enjoyed a great and still growing popularity
in many applications [1–3]. Titania-based photocatalytic 2.1 Materials
systems are used for a variety of applications, such as de-
composition of unwanted and toxic organic compounds, Titanium dioxide powders were synthesized employing the
degradation of pollutants from contaminated water and air sol–gel method, in which titanium tetraisopropoxide
and destruction of harmful bacteria and cancer cells [4–9]. [Ti(OC3H7)4—TTIP, 97 %, Sigma-Aldrich] was used as
The properties of TiO2 are determined by the morphology the precursor of titania, ammonia aqueous (NH3H2O,
of its particles, the size of its crystals and its crystalline 25 %, Chempur) as a catalyst, propan-2-ol (C3H7OH—
structure, which depend on the choice of method for its IPA, 99.5 %, Chempur) as the solvent and acetylacetone
synthesis and final heat treatment [10]. Nanocrystalline (C5H8O2—AcAc, 99 %, Sigma-Aldrich) as a chelating
TiO2 particles are usually obtained by crystallization agent. All reagents were used without any further
(chemical precipitation) [11], the microemulsion method purification.
(reverse micelles) [12], the sol–gel method [13–17] or
hydrothermal crystallization [18–22]. Each of these meth- 2.2 Synthesis of TiO2 via the sol–gel method
ods has its own advantages and drawbacks, but a common
factor that connects them is the ability to obtain materials Titanium dioxide powders were prepared via a sol–gel
with strictly defined properties. process, in which titania sol was prepared by mixing tita-
The sol–gel method is a very flexible route for the nium tetraisopropoxide, propan-2-ol and ammonia, with or
synthesis of advanced materials in a wide variety of forms, without chelating agent, at room temperature. In the first
such as spherical or ultrafine shaped powders, fibers, thin approach, TiO2 sol was prepared without the modifier
film coatings, dense or porous materials including high- (AcAc); TTIP as the starting material was introduced at a
purity inorganic oxides and hybrid (inorganic–organic) constant rate of 1 mL/min using an ISM833A peristaltic
materials [23–27]. The sol–gel process is a useful synthetic pump (Ismatec) into a conical flask filled with a solution
method for the preparation of amorphous, as well as prepared by adding an appropriate amount of ammonia to
structurally ordered, materials. It offers advantages such as IPA. The initially clear solution turned into a white dis-
the possibility of obtaining homogeneous hybrid materials persion. This solution was vigorously stirred for 1 h using a
at low temperature, thereby enabling the incorporation of a high-speed stirrer T25 basic type (IKA Werke GmbH),
variety of compounds [28–34]. A typical sol–gel method is working at 800 rpm. The entire system was transferred to a
based on hydrolysis and condensation of the precursor round-bottomed flask and placed in a vacuum rotary
(inorganic metal salts or metal organic compounds such as evaporator (Rotavapor RII, Büchi Labortechnik GmbH), in
metal alkoxides) [24]. This process can be employed for order to remove the solvent IPA (water bath temperature
the synthesis of functionalized inorganic oxides with con- 60 °C, pressure 136 mbar). The next stage involved fil-
trolled particle size and shape [35–42]. In the sol–gel tration of the mixture under reduced pressure. The sample
method, control of the reactivity of the metal alkoxides obtained in this way was washed with distilled water to
used as precursors is very important in order to obtain sols eliminate IPA. At the final stage, the sample was dried by
and gels with desirable properties. In sol–gel processing, convection at 105 °C for 18 h (SEL-I3 chamber drier,
for better control of the hydrolysis and condensation pro- Memmert). The resulting samples were then calcined at
cess, many different modifiers of alkoxide precursors can 600, 700 or 800 °C for 2 h (Nabertherm type Controller
be used, including acetylacetone [43–45], acetic acid [43, P320).
46] and other complex ligands. In the second approach, the TiO2 sol was prepared in the
The aim of this work is to study the correlation between presence of acetylacetone. First, an appropriate amount of
the photocatalytic activity of TiO2 powders and various of acetylacetone was mixed with TTIP using an IKAMAG
its properties, including crystalline structure, surface area R05 magnetic stirrer (IKA Werke GmbH) for 40 min. In
and particle size. It was investigated how certain conditions the sol–gel method, AcAc was used as a stabilizing agent
of preparation (addition of catalyst and chelating agent, and to control the rate of hydrolysis and condensation of the
temperature of calcination) affect the microstructural metal alkoxide. Acetylacetone has a reactive hydroxyl

123
J Sol-Gel Sci Technol

The morphology and microstructure of the TiO2 pow-


ders obtained was analyzed using a Jeol 1200 EX II
transmission electron microscope, at an accelerating volt-
age of 100 kV. Microscopic photographs were taken using
a direct reflection method, which involves covering a net
with the firmware foil and then with a thin layer of carbon.
On the surface prepared in this way, powder from the
suspension was introduced. All samples of the prepared
systems were tested in analogous electron conditions with
appropriate magnification.
X-ray diffraction (XRD) analysis was employed to deter-
mine the crystalline structure of the TiO2 materials. The mea-
surements were performed using a TUR-M62 diffractometer,
operating at 30 kV and 25 mA, with CuKa (a = 1.5418 Å)
radiation, Ni filtered. The XRD pattern data were collected in
step-scanning mode with steps of D2h = 0.04°. Based on the
XRD analysis results, it was possible to calculate crystallites
size as well as to evaluate the relative content of rutile and
anatase phases in prepared samples. Scherrer’s equation [48],
(101) and (110) reflections of anatase and rutile planes, re-
spectively, were used for this purpose.
An ASAP 2020 porosimetry analyzer (Micromeritics
Fig. 1 Interaction of acetylacetone with TTIP Instrument Co.) was used to determine the porous structure
parameters of the TiO2 materials, including Brunauer–
Emmett–Teller (BET) surface area, pore size and distribu-
group which reacts easily with the metal alkoxide and tion in the mesoporous range, using low-temperature N2
causes the transfer of an acidic proton from the acety- sorption. The samples were first degassed for 4 h at 120 °C.
lacetone to an alkoxy ligand, resulting in the corresponding The thermal stability of the obtained TiO2 samples was
alcohol and a modified alkoxide precursor, as presented in determined by thermogravimetric analysis (TGA/DTA)
Fig. 1 [45, 47]. (Jupiter STA 449 F3, Netzsch). Measurements were carried
The reaction of metal alkoxide with AcAc is exothermic out under flowing nitrogen at a heating rate of 10 °C/min
and produces a yellow solution, whereas both precursors and in a temperature range of 30–1000 °C, with an initial
were colorless liquids. The resulting precursor was then sample weight of approximately 5 mg.
introduced at a constant rate of 1 mL/min, using an To identify the characteristic groups present on the
ISM833A peristaltic pump (Ismatec), into the solution surface of the TiO2 systems, the samples were subjected to
prepared by adding an appropriate amount of ammonia to FT-IR analysis using an IFS 66v/s spectrophotometer
IPA. The products were analyzed in the same way as in the (Bruker). The samples were studied in the form of KBr
case of the preparation of pure titanium dioxide. tablets, as KBr crystals are inactive in the IR range. The
analysis was performed over a range of 4000–400 cm-1.
2.3 Determination of physicochemical properties
2.4 Evaluation of photocatalytic activity
To determine the effect of the addition of the catalyst and
chelating agent, and of the temperature of calcination, on The photocatalytic activity of selected samples was tested via
the product’s physicochemical properties, the obtained photodegradation of the organic dye C.I. Basic Blue 9. In order
TiO2 samples were comprehensively analyzed using the to establish the adsorption/desorption equilibrium between
most advanced analytical methods and techniques. the dye and the catalyst surface, an appropriate amount of a
A Zetasizer Nano ZS (Malvern Instruments Ltd.) was TiO2 sample (the catalyst, 0.04 g) was dispersed in an aqueous
used to determine the particle size distributions in the range solution of C.I. Basic Blue 9 with a concentration of 5 mg/L
0.6–6000 nm, based on the noninvasive backscattering and stirred using an IKAMAG R05 magnetic stirrer (IKA
technique (NIBS). Each sample was prepared by dispersing Werke GmbH) for 15 min in the dark. Next, the suspension
0.01 g of the tested product in 25 mL of propan-2-ol. The was irradiated with UV light for 15 min (Fig. 2), utilizing
system was stabilized in an ultrasonic bath for 15 min, and Laboratory-UV-Reactor system 2 (Heraeus). The UV light
it was then placed in a cuvette and analyzed. source was a 150W Hg lamp at room temperature. The

123
J Sol-Gel Sci Technol

synthetic TiO2 obtained via the sol–gel method with dif-


ferent molar ratios of catalyst (ammonia).
The results of the dispersive analysis (Table 1) show
that both the addition of a catalyst and the conditions of the
thermal process have a significant effect on the dispersion
of the resulting materials. The data show that, irrespective
of the temperature of calcination, samples with the most
favorable dispersion were obtained at the molar ratio TTIP/
NH3H2O = 1.48. All of the products have monomodal
particle size distributions. Samples T1, T4 and T7, obtained
using the molar ratio TTIP/NH3H2O = 1.48 and calcined,
respectively, at 600 °C, 700 °C and 800 °C, contain par-
ticles in the ranges 122–615, 220–712 and 396–1110 nm.
These bands correspond to the presence of primary parti-
cles and agglomerates. These samples also exhibit high
homogeneity, as is shown by the low values of the poly-
Fig. 2 Evaluation of photocatalytic activity of TiO2 catalyst dispersity index (respectively, 0.256, 0.113 and 0.187). It
was also observed for all of the analyzed materials that,
intermediate solutions were separated by filtration, and the with an increase in the quantity of ammonia used as a
concentration of C.I. Basic Blue 9 in aqueous solution (both catalyst in the process of obtaining TiO2 by the sol–gel
after exposure and in the unexposed solution) was monitored method, there is an increase in the particle diameters of the
by measuring light absorption at 661 nm with a SPEKOL UV- various systems. It can be assumed that an increase in the
1201 spectrometer (Shimadzu), using water as reference. The quantity of catalyst used in the sol–gel process leads to an
spectra obtained made it possible to evaluate the concentration increase in the degree of nucleation and hydrolysis, re-
of decomposed dye using the Lambert–Beer law (1): sulting in particles of larger diameter and with a significant
tendency to agglomerate. The results of the dispersive
A
ct ¼ ð1Þ analysis also showed that the calcination temperature sig-
ðe  lÞ
nificantly affects the dispersion of the resulting titanium
where ct is the concentration of the dye after irradiation; A dioxide. As the heat processing temperature increases, the
is the absorbance; e is the molar absorption coefficient, particle diameters are found to shift toward higher values,
emax = 66,700 L/molcm at a wavelength of k = 661 nm this being an effect of the sintering and greater agglom-
for C.I. Basic Blue 9; and l is the thickness of the absorbing eration of the particles (Table 1).
layer (on the path of the radiation passing through the The TEM microphotographs of selected titanium diox-
cuvette solution, l = 1 cm). ide samples (samples T1, T4 and T7, obtained at the molar
The photocatalytic activity of titanium dioxide in the ratio TTIP/NH3H2O = 1.48 and calcined at temperatures
photodegradation of the organic dye was determined by of 600 °C, 700 °C and 800 °C, respectively) are presented
calculating the yield of dye degradation (W), using the in Fig. 3.
formula (2): TEM images of samples obtained (Fig. 3) inform about
  their inhomogeneous structure and the presence of primary
ct
W ð%Þ ¼ 1   100 % ð2Þ particles and secondary agglomerates. Particles of irregular
c0
shapes make larger clusters. The TEM microphotographs
where c0 is the concentration of the substrate prior to ir- of the studied samples confirm the presence of particles of
radiation and ct is the concentration of the substrate after precisely designed diameter (corresponding to those indi-
irradiation. cated in the particle size distributions).
At the next stage, it was investigated how the addition of
chelating agent affects the dispersion and microstructure of
3 Results titanium dioxide obtained by the sol–gel method. It was
first determined how the addition of acetylacetone affects
3.1 Dispersive and microstructural properties the dispersive properties of samples additionally calcined
of TiO2 at 600 °C. Dispersive analysis of these samples of TiO2
(Table 2) showed that the addition of AcAc leads to
The aim of the first stage of the study was to determine the products with smaller particle diameter values. The most
dispersive and microstructural properties of the inorganic, favorable dispersion was observed in samples obtained

123
J Sol-Gel Sci Technol

with the addition of chelating agent in a quantity of quantity of AcAc decreases. It is well known that the
0.0096 mol, irrespective of the quantity of ammonia used presence of chelating agent in the synthesis of titanium
as catalyst (samples T19, T20 and T21). The analyzed dioxide decreases the rate of hydrolysis and results in the
samples have monomodal particle size distributions, in the formation of particles of smaller diameter [46].
ranges 164–396 nm (sample T19) and 164–459 nm (sam- Figure 4 shows TEM microphotographs of selected sam-
ples T20 and T21). The addition of chelating agent leads ples (T19, T22 and T25) obtained with the addition of
not only to products characterized by the presence of par- chelating agent in an amount of 0.0096 mol and calcined at
ticles of smaller diameter, but also a visible shift in the different temperatures, which confirm the presence of parti-
diameter of the particles with the maximum volume con- cles of nanosized diameters (corresponding to those in the
tribution toward smaller diameter values. Among the ana- indicated particle size ranges). The particles are almost
lyzed samples, the maximum volume contribution comes spherical in shape and show only a slight tendency to form
from particles of diameter 255 nm (samples T19 and T21) agglomerate structures. The TEM images of TiO2 particles
and 295 nm (sample T20). However, analysis of the values prepared with AcAc addition show spherical particles and
of the polydispersity index shows that the addition of a narrow size distribution. TiO2 particles obtained without
stabilizer in a quantity of 0.0096 mol leads to a product chelating ligand (Fig. 3) are highly agglomerated, and most of
which is less homogeneous than a sample obtained without them are non-spherical. Those differences in morphology and
the addition of AcAc. particle sizes indicate that the addition of the nucleophilic
Table 2 also shows the dispersive properties of samples ligand in preparation of TiO2 samples leads to fine powders
of titanium dioxide obtained with or without the addition of containing homogeneous as well as spherical particles.
stabilizer, calcined at 700 and 800 °C, respectively. Dis-
persive analysis of the samples obtained with the addition 3.2 Structural characteristic of synthetic titanium
of acetylacetone shows that the proposed method of syn- dioxide
thesis leads to a product in which particles with smaller
diameter values are present. In most cases, it also leads to XRD patterns of TiO2 samples calcined at different tem-
products with poorer homogeneity. The samples with the perature (600, 700 and 800 °C) are shown in Fig. 5. XRD
most favorable dispersion were obtained by adding measurement of titanium dioxide calcined at 600 °C (the
chelating agent in a quantity of 0.0096 mol (samples T22– sample labeled as T3) shows a strong peak occurring at
T24 and T25–T27, calcined at 700 and 800 °C, respec- 2h = 25.3°, which corresponds to the (101) reflection,
tively). In summary, it was confirmed that the addition of a while other characteristic peaks—(004), (200), (105) and
stabilizer causes a shift in particle diameters toward lower (211)—correspond to different crystalline planes. These
values, irrespective of the further thermal processing of the peaks confirm the presence of the polycrystalline anatase
TiO2. Our observations are in agreement with previous structure of TiO2. The rutile TiO2 phase is not observed in
research [44]. The results of dispersive analysis showed an this sample. The XRD patterns of titanium dioxide powders
apparent downward trend in the particle diameter as the calcined at higher temperatures (700 and 800 °C—samples

Table 1 Dispersive properties of titanium dioxide powders obtained at different amount of catalyst
Sample Amount of reactants (mol) Particle size Maximum Polydispersity
distributions volume index
IPA TTIP NH3H2O by volume (nm) contribution (%)

Calcined at 600 °C
T1 1.304 0.049 0.033 122–615 164 nm—14.4 0.256
T2 0.049 255–1480 459 nm—13.1 0.359
T3 0.065 459–1480 825 nm—24.2 0.311
Calcined at 700 °C
T4 1.304 0.049 0.033 220–712 396 nm—23.8 0.113
T5 0.049 255–955 531 nm—21.6 0.257
T6 0.065 459–1480 825 nm—23.4 0.159
Calcined at 800 °C
T7 1.304 0.049 0.033 396–1110 712 nm—28.4 0.187
T8 0.049 531–1480 955 nm—27.7 0.431
T9 0.065 531–2300 1480 nm—15.5 0.321

123
J Sol-Gel Sci Technol

Fig. 3 TEM images of titanium


dioxide powders: a T1, b T4
and c T7

T6 and T9) were found also to be crystalline, but with measure at its half-high utilizing the Scherrer equation.
diffraction peaks corresponding to both anatase and rutile Sample T3, calcined at 600 °C, contains only anatase phase
forms. For both of these samples, there was a perceptible with an average crystallite size of about 39 nm. On the
weakening in the intensity of the diffraction peaks (101), other hand, sample calcined at 700 °C—T6—is composed
(004), (200) and (211) that correspond to the anatase form, of anatase crystallites of 47 nm and the transformed rutile
while the characteristic peaks (110), (101), (200), (111), phase that appears with crystallites size of about 39 nm.
(210), (211) and (220) that correspond to the rutile struc- Sample T9 (calcined at 700 °C) contains both anatase and
ture were also observed. The formation of a rutile structure rutile phases with crystallites of 47 nm in size. The relative
was observed in TiO2 samples calcined above 700 °C percentage content of anatase and rutile phases in obtained
(samples T6 and T9). The calcination process is necessary samples was calculated using equation given in the lit-
to eliminate water and organic species arising from the erature [49]. Sample T3 is composed only of anatase, and
initial reactants. Moreover, it was found that the tem- sample T6 contains 25 and 75 % and sample T9 contains
perature of calcination significantly affects the crystalline 11 and 89 % of anatase and rutile, respectively. The
structure of the titanium dioxide. Calcination at 600 °C crystallites size as well as the relative percentage content of
leads to a product with a pure anatase structure, whereas a rutile increases together with increasing temperature of
higher calcination temperature leads to a product with a calcination.
mixed rutile and anatase structure, in which rutile In the next stage of the physicochemical analysis, the
nonetheless dominates. The XRD analysis shows that the effect of the addition of chelating agent (acetylacetone) on
contribution of rutile structure increases with increasing the crystalline structure of calcined titanium dioxide was
temperature of calcination. examined (Fig. 6). Figure 6a shows diffractograms of
Table 3 presents crystallite sizes and relative percentage samples of titanium dioxide obtained with the addition of
content of anatase and rutile in obtained titanium dioxide various amounts of acetylacetone and calcined at 600 °C.
samples, additionally calcined at various temperatures. The In the case of the samples obtained with the use of the
average crystallite sizes of TiO2 samples were calculated chelating agent (samples T12 and T21), the XRD patterns
using width of the XRD peak at 2h = 25.3° and 27.5° show a weak peak occurring at 2h = 27.4°, which

123
J Sol-Gel Sci Technol

Table 2 Dispersive properties of titanium dioxide powders obtained at different amount of catalyst and chelating agent and calcined at different
temperature
Sample Amount of reactants (mol) Particle size Maximum Polydispersity
distributions volume index
IPA TTIP NH3H2O AcAc by volume (nm) contribution
(%)

Calcined at 600 °C
T1 1.304 0.049 0.033 – 122–615 164 nm—14.4 0.256
T10 0.0193 255–615 396 nm—29.7 0.285
T19 0.0096 164–396 255 nm—34.3 0.808
T2 0.049 – 255–1480 459 nm—13.1 0.359
T11 0.0193 255–459 342 nm—37.1 0.742
T20 0.0096 164–459 295 nm—21.1 0.190
T3 0.065 – 459–1480 825 nm—24.2 0.311
T12 0.0193 220–712 342 nm—24.7 0.374
T21 0.0096 164–459 255 nm—26.5 0.411
Calcined at 700 °C
T4 1.304 0.049 0.033 – 220–712 396 nm—23.8 0.113
T13 0.0193 122–255 164 nm—30.4 0.643
T22 0.0096 106–255 142 nm—39.1 0.588
T5 0.049 – 255–955 531 nm—21.6 0.257
T14 0.0193 255–712 459 nm—31.4 0.195
T23 0.0096 122–396 190 nm—24.0 0.418
T6 0.065 – 459–1480 825 nm—23.4 0.159
T15 0.0193 220–825 459 nm—23.8 0.141
T24 0.0096 190–615 295 nm—25.3 0.524
Calcined at 800 °C
T7 1.304 0.049 0.033 – 396–1110 712 nm—28.4 0.187
T16 0.0193 164–712 190 nm—16.2 0.921
T25 0.0096 164–531 295 nm—23.5 0.173
T8 0.049 – 531–1480 955 nm—27.7 0.431
T17 0.0193 255–1110 531 nm—22.6 0.145
T26 0.0096 122–396 190 nm—19.8 0.589
T9 0.065 – 531–2300 1480 nm—15.5 0.321
T18 0.0193 190–615 342 nm—23.3 0.392
T27 0.0096 142–459 220 nm—21.5 0.217

corresponds to the (110) reflection of the rutile structure, as rutile phases with crystallite size of 47 and 39 nm, respec-
compared to the sample obtained without AcAc (sample tively. It means that the presence of chelating agent in
T3). Analysis of the data leads to the conclusion that the preparation process leads to decrease in temperature of
addition of acetylacetone at the stage of titanium dioxide transformation of anatase to rutile. Moreover, anatase phase
synthesis, followed by calcination, leads to a mixed anatase is characterized with higher crystallite sizes as compared to
and rutile structure, the former type being dominant. the sample T3. The higher the temperature of calcination
However, diffractograms obtained for the samples calcined (700 and 800 °C—samples T15, T24 and T18, T27, re-
at higher temperatures show a reduction or vanishing of the spectively), the increase in rutile phase content together
characteristic anatase diffraction bands (101), (004), with decrease in amount of anatase was observed. Samples
(200)—this applies to samples T15, T24 and T18, T27, prepared with the smallest addition of AcAc are charac-
calcined, respectively, at 700 and 800 °C. terized with the biggest crystallites of 59 nm. Our obser-
Preparation of TiO2 samples with the addition of AcAc vations are in agreement with Chang et al.’s studies [49].
caused significant changes in crystallites characteristic. Based on the results of XRD analysis, it was concluded
Sample T12 (calcined at 600 °C) contains both anatase and that the preparation of titanium dioxide in the presence of a

123
J Sol-Gel Sci Technol

Fig. 4 TEM images of titanium


dioxide obtained with the
addition of AcAc: a T19, b T22
and c T25

3.3 Porous structure

In order to investigate the structural properties of TiO2, se-


lected samples were analyzed using a physisorption analyzer.
Figures 7 and 8 show the fundamental parameters used to
determine the structural properties of the samples obtained
(BET surface area, total volume and mean size of pores).
The nitrogen isotherms recorded for obtained TiO2
samples indicate the non-porous or macroporous character
of the materials (Fig. 7). Those isotherms were classified as
type II with hysteresis loops type H3. Sample T3 (calcined
at 600 °C) is characterized with the highest surface area
(BET) of 13.9 m2/g.
At the next stage, the influence of chelating agent ad-
dition on the parameters of the porous structure of titanium
Fig. 5 XRD patterns of titanium dioxide calcined at 600, 700 and dioxide obtained by the sol–gel method (Fig. 7b) was in-
800 °C vestigated. Analysis of the data presented in Fig. 7b shows
that the smaller the amount of the chelating agent, the
chelating agent (AcAc) influences the crystalline structure higher the surface area parameter. As compared to sample
of the TiO2 and causes a decrease in the temperature of T3, the highest surface area of 15.7 m2/g was noted for T21
transformation of the anatase phase to rutile. With an in- powder.
crease in the quantity of the chelating agent introduced at Analysis of the porous structure of titanium dioxide
the stage of synthesis of TiO2, there is observed a decrease obtained with appropriate amounts of ammonia and cal-
in the intensity of the diffraction peaks characteristic of cined at different temperatures (Fig. 8) shows that sample
anatase (2h = 25.3°), irrespective of the temperature of T3 (calcined at 600 °C) has the highest surface area
thermal processing. (13.9 m2/g), while its pore volume is 0.010 cm3/g and

123
J Sol-Gel Sci Technol

Table 3 Crystallites size and Sample Amount of reactants (mol) Crystallite size (nm) Content (%)
relative percentage content of
anatase and rutile in titanium IPA TTIP NH3H2O AcAc Anatase Rutile Anatase Rutile
dioxide powders obtained at
different amount of chelating Calcined at 600 °C
agent and calcined at different T3 1.304 0.049 0.065 – 39 – 100 –
temperature T12 0.0193 47 39 95 5
T21 0.0096 47 47 93 7
Calcined at 700 °C
T6 1.304 0.049 0.065 – 39 39 25 75
T15 0.0193 47 59 7 93
T24 0.0096 59 59 13 87
Calcined at 800 °C
T9 1.304 0.049 0.065 – 47 47 11 89
T18 0.0193 47 59 3 97
T27 0.0096 59 59 3 97

Fig. 6 XRD patterns of


titanium dioxide obtained with
AcAc addition and calcined at:
a 600 °C, b 700 °C and
c 800 °C

mean pore diameter 2.8 nm. Sample T9 (calcined at area of 15.7 m2/g, a total pore volume of 0.011 cm3/g and a
800 °C) has the lowest surface area (7.4 m2/g), its pore pore diameter of 2.9 nm. In the case of sample T27 (cal-
volume and mean pore diameter being 0.004 cm3/g and cined at 800 °C), the total pore volume is 0.002 cm3/g, the
4.5 nm, respectively. pore diameter 2.7 nm and the BET surface area 9.2 m2/g.
Based on analysis of the data for TiO2 samples obtained The addition of a chelating agent in the amount 0.0096 mol
with the addition of acetylacetone in an amount of leads to an increase in the surface area and pore volume of
0.0096 mol (samples T21, T24 and T27), it is found that the products, irrespective of the temperature of calcination.
these products have slightly higher values for the BET The values for pore diameter, however, were found to
surface area than those of TiO2 obtained in the traditional decrease. The increase in the BET surface area for the
manner. Sample T21 (calcined at 600 °C) has a surface samples obtained using the chelating agent (AcAc) in an

123
J Sol-Gel Sci Technol

Fig. 7 Nitrogen adsorption/desorption isotherms for titanium dioxide obtained: a without chelating agent and calcined at different temperature
and b with AcAc addition and calcined at 600 °C

Fig. 8 Parameters of the porous


structures of titanium dioxide
obtained with or without
addition of chelating agent and
calcined at different
temperature: a BET surface
area, b pore volume and, c pore
size versus AcAc amount

amount of 0.0096 mol at the titanium dioxide synthesis 0.0193 mol. The addition of AcAc (0.0193 mol) also re-
stage may be directly related to the dispersive nature of the sulted in an increase in the mean size of pores relative to
materials obtained. These samples contained particles with those of native TiO2. It was also observed that the BET
smaller diameters than those of pure TiO2, this being di- surface area decreased with increasing temperature of the
rectly linked to the porous structure parameters of the calcination process.
products of synthesis. Addition of the chelating agent during the prepara-
However, the addition of AcAc in an amount of tion of TiO2 caused significant changes in the porous
0.0193 mol causes significant deterioration in the porous structure parameters of the resulting samples. The re-
structure parameters of the products. A considerable de- sults of BET analysis of the samples confirmed that by
crease in the BET surface area and pore volume relative to selecting an appropriate quantity of modifier (acety-
the pure TiO2 sample was observed for all samples ob- lacetone), it is possible to control the porous structure
tained with the addition of chelating agent in an amount of of the product.

123
J Sol-Gel Sci Technol

In contrast to the results of dispersive characteristics,


determination of the porous structure parameters confirmed
the surface area increase with the corresponding increase in
the amount of primary particles in prepared TiO2 samples.

3.4 Thermal analysis

The thermal stability of selected samples was estimated


using TGA/DTA analysis. On the basis of this analysis,
thermograms were obtained showing thermogravimetric
(TGA) and differential thermal analysis (DTA) curves,
indicating physical and/or chemical changes that occur in a
sample’s structure during heat treatment.
The results of TGA of TiO2 samples (before calcination
process), prepared with various amounts of catalyst and Fig. 9 TGA/DTA curves of titanium dioxide samples obtained with
AcAc addition, are shown in Fig. 9. All TGA/DTA curves and without AcAc addition before calcination process
show endothermic peaks and the sample weight loss up to
130 °C, which are attributed to the removal of physically
adsorbed water. The exothermic peak at 270 °C and a However, analysis of the TGA curves for the samples
broad one at 300–400 °C are related to decomposition of obtained with the addition of AcAc and calcined at 600 °C
organic compounds, residual hydroxyl groups and AcAc (Fig. 10b) shows the major weight loss of about 0.5–0.8 %
ligands. Additionally, the exothermic peak observed at to occur in the temperature range 30–600 °C. This weight
390 °C correlates with the crystallization of the amorphous loss is mainly associated with the local elimination of water
phase into the anatase. For sample obtained with chelating bonded with the surface of the sample. In the temperature
agent addition, this peak appears at 395 °C, different as in range 600–1000 °C, the TGA curves show an additional
the case of TiO2 powder prepared without AcAc (310 °C), weight loss of about 0.7 and 1.0 % for T12 and T21, re-
which is in agreement with [43]. It can be assumed that spectively, as a result of titania phase transformation as
above 400 °C the powder completely transforms into well as decomposition of the chelating agent.
crystalline form, which is confirmed by weight stabiliza-
tion. Moreover, presented TGA curves show two different 3.5 FT-IR analysis
steps of samples degradation—first in the range of
80–120 °C related to free water loss and second in the In order to identify the functional groups present on the
range of 120–400 °C related to the following sample surface of the TiO2 prepared by the sol–gel method, with or
decomposition. without chelating agent, selected samples were subjected to
Figure 10a shows thermograms of titanium dioxide FT-IR analysis.
systems obtained without the chelating agent and calcined Figure 11 shows the FT-IR spectra of samples obtained
at different temperatures (600, 700 and 800 °C). The re- with and without AcAc (samples T3, T12 and T21) and
sulting systems were analyzed to determine their thermal calcined at 600 °C. The broad absorption peaks around
stability. In all of the systems, the main mass loss is ob- 800–500 cm-1 are characteristic of a Ti–O–Ti network.
served within a temperature range of 30–600 °C, corre- The broad absorption peaks at approximately 3400 cm-1
sponding to the loss of physically and chemically bound are ascribed to the stretching vibrations of O–H bonds
water. At this point, the mass loss is slightly above 0.7, 0.6 indirectly related to water physically adsorbed on the sur-
and 0.9 %, respectively, for samples T3, T6 and T9 and is face. The band at 1430 cm-1 is associated with symmetric
almost equal to the total mass loss (*0.9, 0.7 and *1.2 % vibrations of the carboxyl (C–O) group. Analysis of the
for T3, T6 and T9, respectively). There was also noted a FT-IR spectra of the samples obtained with AcAc and
small exothermic effect associated with the transformation calcined at 600 °C reveals a few peaks characteristic of the
of the anatase form to the rutile form of titanium dioxide. chelating agent. The FT-IR spectra do not show any
DTA analysis of the TiO2 samples calcined at different characteristic absorption peaks for free acetylacetone at
temperatures shows exothermic peaks in a temperature around 1707 and 1726 cm-1 due to the stretching C–C and
range of 100–600 °C, which also corresponds to crystal- C–O vibrations which are assigned of keto tautomeric C–O
lization of the amorphous phase into the anatase or rutile group. When AcAc reacted completely with TTIP, the FT-
phase. IR spectra showed a sharp absorption peak at 1580 cm-1,

123
J Sol-Gel Sci Technol

Fig. 10 TGA/DTA curves of


titanium dioxide obtained:
a without chelating agent and
calcined at different temperature
and b with AcAc and calcined at
600 °C

rising from the split of enol form of C–O at 1630 cm-1. 600–800 °C (samples T3, T6 and T9) show higher photo-
This proves the chelation occurring between AcAc and catalytic activity in the decomposition of C.I. Basic Blue 9.
TTIP [47, 50]. A peak was observed at about 1360 cm-1, The highest efficiency of C.I. Basic Blue 9 degradation
which was one of the characteristics of C–O deformation (over 98.3 %) was observed in the presence of TiO2 ob-
(for samples T12 and T21). tained without AcAc and calcined at 800 °C (sample T9).
Moreover, it was observed that photocatalytic activity in-
3.6 Photocatalytic activity creases with increasing temperature of calcination of tita-
nium dioxide. This confirms certain literature reports
Figure 12 shows the photocatalytic degradation of C.I. stating that a mixture of rutile and anatase has better
Basic Blue 9 by UV irradiation in the presence of TiO2 photocatalytic properties than pure anatase [52, 53].
prepared by the sol–gel method with or without chelating The TiO2 samples obtained with the addition of chelating
agent. agent and calcined at temperatures of 600–800 °C (samples
The efficiency of C.I. Basic Blue 9 photodegradation in T21, T24 and T27) show lower activity in the photocatalytic
the presence of the obtained samples was measured with decomposition of C.I. Basic Blue 9. The degradation effi-
reference to A11 commercial titanium dioxide (Chemical ciency was 95.9 % in the presence of sample T21 and
Works Police SA). Moreover, the decolorization of C.I. slightly lower (93.8 and 89.0 %) in the case of photo-
Basic Blue 9 under UV irradiation was performed, with catalysis using samples T24 and T27.
25 % degradation efficiency, which corresponded to the
literature [51]. The measurements showed the commercial
anatase titanium dioxide (A11) to have very good pho- 4 Discussion
tooxidation activity (the efficiency of its degradation of C.I.
Basic Blue 9 was 97 %). The TiO2 samples obtained The alkoxides are very air-sensitive and rapidly precipitate
without AcAc and calcined at temperatures in the range with moisture, leading to the component segregation and
agglomeration of large-sized complex hydroxide particles;

Fig. 11 FT-IR spectra of titanium dioxide obtained with and without Fig. 12 Efficiency of C.I. Basic Blue 9 photodegradation in the
AcAc and calcined at 600 °C presence of synthesised TiO2 samples

123
J Sol-Gel Sci Technol

hence, adding a chelating agent such as binary alcohol or was made of the influence of pH on the morphology and
organic acids to the alkoxide solution is an effective way to crystalline structure of titania obtained by the sol–gel method
control the rate of hydrolysis. Several literature reports in the presence of acetic acid as chelating agent. It was found
have confirmed the influence of different modified alkoxide that the presence of chelating agent (AcOH) or ligands (H2O2)
precursors on the physicochemical properties of the re- and their sequence of addition have a considerable influence on
sulting titania. the phase formation and morphology of TiO2. The authors
Attar et al. [43] reported the effect of modifier ligands noted that the pH of the solution critically determines the
such as acetylacetone and acetic acid on the formation of composition of the hydrolyzed product, which in turn influ-
titanium dioxide via the sol–gel method. The authors studied ences the titania phase type. The addition of AcOH/TTIP or
the influence of chelating agent on the morphology, thermal TTIP/AcOH solution (with pH in the range 3.9–4.1) to water or
stability and crystalline structure of the resulting titania. The vice versa gave a solution in the pH range 3.2–3.5, which is
SEM images of titania prepared without chelating ligands favorable for anatase nucleation. When TTIP was introduced
show that some particles are non-spherical and highly ag- to H2O2 solution, the pH changed from 3.4–3.5 to 1.1, and
glomerated. Morphological analysis of samples obtained in nucleation of both anatase and rutile was observed in the same
the presence of chelating agent shows that the addition of proportions. The addition of TTIP/AcOH to H2O2 solution
complex ligands leads to products with particles of smaller caused the pH to drop from 3.9–4.1 to 1.5, rather slowly,
diameter (about 20–25 nm for AcAc and 25–30 nm fol- leading to predominantly anatase (92 %) with traces of rutile.
lowing the addition of acetic acid) compared with native Huang et al. [45] obtained titanium dioxide via a hy-
TiO2. Samples prepared with chelating agent are character- drothermal method, using titanium tetrabutoxide as the
ized by particles of spherical shape. The type of chelating source material and acetylacetone as the chelating agent.
agent was found to play an important role in the thermal They studied the influence of the chelating agent on the
stability. Titanium dioxide prepared with acetylacetone or crystalline structure and morphology of the resulting TiO2.
without chelating agent showed higher weight loss (22 and The titanium dioxide obtained with the addition of
17 %, respectively) than that synthesized with acetic acid chelating agent had particles with diameters ranging from
(15 %). Titanium dioxide obtained without chelating agent several nanometers to 20 nm. The authors noted that par-
is crystalline after calcination at 300 °C, in contrast to ticle size and average crystallite size decreased gradually
samples prepared with acetylacetone and acetic acid, which when the molar ratio of AcAc/Ti was increased to 0.2, and
prevent the crystallization of titania at temperatures below after that, they increased slightly up to a molar ratio of 0.4.
400 °C. XRD analysis showed that the presence of chelating Choi et al. [54] synthesized thin films and membranes of
agent caused an increase in the temperature of transition of TiO2 with enhanced catalytic activity and better structural
the amorphous phase to anatase. properties using a novel simple sol–gel route, employing
You et al. [44] studied the influence of the amount of acetic acid and TweenÒ 80 as surfactant. The resulting titania
acetylacetone (from 0 to 1 mL) and the reaction time (from 2 had high surface area (147 m2/g) and high porosity (46 %), a
to 32 h) on the morphology and particle size, surface area, narrow pore size distribution ranging from 2 to 8 nm, ho-
chemical composition, crystalline structure and thermal mogeneity without cracks and pinholes, as well as enhanced
stability of TiO2 prepared by a single-step swelling process catalytic properties such as active anatase phase, and small
of a polystyrene template. They noted that the amount of crystallite size (9 nm). The synthesized TiO2 photocatalysts
AcAc and the reaction time affect the physicochemical were highly efficient for the destruction of C.I. Basic Blue 9
properties of the synthesized materials. The particle size of and creatinine in water. The prepared photocatalytic TiO2/
the TiO2/PS composite before the polymer was removed was Al2O3 composite membranes exhibited high water perme-
larger than the particle size after calcination. Dispersive ability and effective organic retention. The proposed mod-
analysis showed an apparent upward trend in the diameter of ified sol–gel method was useful for the preparation of
particles as the amount of AcAc and reaction time increased. nanostructured TiO2 films and membranes with high pho-
The titanium dioxide particles were spherical with a diameter tocatalytic activity and desired pore structure, as well as for
of 450 nm when the amount of chelating agent approached the synthesis of similar crystal nanostructures of other oxide
0.1 mL and the reaction time reached 8 h. The results of BET materials for applications in catalysis and separation.
analysis showed that titanium dioxide with large BET sur-
face area and pore volume was obtained as the reaction time
and quantity of chelating agent increased. 5 Conclusions
Chang et al. [46] obtained titania nanocrystals from aqueous
solutions of peroxo titanium complex starting from titanium The results of the dispersive analysis presented above show
tetraisopropoxide (TTIP), acetic acid and hydrogen peroxide that an increase in the amount of catalyst used in the pro-
in water/propan-2-ol media by a facile sol–gel process. A study cess of obtaining titanium dioxide via the sol–gel method

123
J Sol-Gel Sci Technol

leads to an increase in the degree of nucleation and hy- 3. Siwińska-Stefańska K, Krysztafkiewicz A, Ciesielczyk F, Pauk-
drolysis, resulting in an increase in particle diameter and a szta D, Sójka-Ledakowicz J, Jesionowski T (2010) Physico-
chemical and structural properties of TiO2 precipitated in an
significant tendency for the particles to agglomerate. emulsion system. Physicochem Probl Miner Process 44:231–244
Moreover, in the case of titania obtained with the addition 4. Beydoun D, Amal R, Low G, Mcevoy S (1999) Role of
of chelating agent, the diameter of particles tended to de- nanoparticles in photocatalysis. J Nanopart Res 1:439–458
crease with a decreasing quantity of AcAc. The presence of 5. Blake DM, Maness PC, Huang Z, Wolfrum EJ, Huang J, Jacoby
WA (1999) Application of the photocatalytic chemistry of tita-
chelating agent in the synthesis of titanium dioxide de- nium dioxide to disinfection and the killing of cancer cells. Sep
creases the rate of hydrolysis and results in the formation of Purif Methods 28:1–50
particles with smaller diameter. 6. Mozia S, Bubacz K, Janus M, Morawski AW (2012) Decompo-
The XRD analysis shows that, in the preparation of ti- sition of 3-chlorophenol on nitrogen modified TiO2 photo-
catalysts. J Hazard Mater 203–204:128–136
tanium dioxide without addition of chelating agent (AcAc), 7. Dolat D, Mozia S, Ohtani B, Morawski AW (2013) Nitrogen,
the contribution of the rutile structure increases with in- iron-single modified (N–TiO2, Fe–TiO2) and co-modified (Fe, N–
creasing temperature of calcination. Moreover, XRD ana- TiO2) rutile titanium dioxide as visible-light active photo-
lysis showed that, by selecting a suitable amount of catalysts. Chem Eng J 225:358–364
8. Sthathatos E, Tsiourvas D, Lianos P (1999) Titanium dioxide
chelating agent (AcAc) at the stage of titanium dioxide films made from reverse micelles and their use for the photo-
synthesis, followed by thermal processing, it is possible to catalytic degradation of adsorbed dyes. Colloid Surf A 149:49–56
control its crystalline structure, causing a decrease in the 9. Comparelli R, Fanizza E, Curri ML, Cozzoli PD, Mascolo G,
temperature of transformation of anatase to rutile. Passino R, Agostiano A (2005) Photocatalytic degradation of azo
dyes by organic-capped anatase TiO2 nanocrystals immobilized
The addition of chelating agent in the process of TiO2 onto substrates. Appl Catal B-Environ 55:81–91
preparation caused significant changes in the porous 10. Hadjiivanov KI, Klissurski DG (1996) Surface chemistry of titania
structure parameters of the resulting samples. Analysis of (anatase) and titania-supported catalysts. Chem Soc Rev 25:61–69
the porous structure of TiO2 obtained via the proposed 11. Scolan A, Sanchez C (1998) Synthesis and characterization of
surface-protected nanocrystalline titania particles. Chem Mater
method (with the addition of AcAc in an amount of 10:3217–3223
0.0096 mol) showed these powders to have a higher sur- 12. Lal M, Chhabra V, Ayyub P, Maitra A (1998) Preparation and
face area than native titania. A considerable decrease in the characterization of ultrafine TiO2 particles in reverse micelles by
BET surface area and pore volume relative to that of the hydrolysis of titanium di-ethylhexyl sulfosuccinate. J Mater Res
13:1249–1254
pure TiO2 sample was observed for all samples obtained 13. Su C, Hong B-Y, Tseng C-M (2004) Sol–gel preparation and
with the addition of chelating agent in an amount of photocatalysis of titanium dioxide. Catal Today 96:119–126
0.0193 mol. 14. Yang P, Lu C, Hua N, Du Y (2002) Titanium dioxide nanopar-
The FT-IR analysis confirmed the formation of a chelating ticles co-doped with Fe3? and Eu3? ions for photocatalysis.
Mater Lett 57:794–801
bond between titanium alkoxide and acetylacetone. 15. Bessekhouad Y, Robert D, Weber JV (2003) Synthesis of pho-
The TiO2 systems prepared by the sol–gel method with tocatalytic TiO2 nanoparticles: optimization of the preparation
or without the addition of chelating agent exhibit relatively conditions. J Photochem Photobiol A 157:47–53
high photocatalytic activity in the decomposition of C.I. 16. Kormann C, Bahnemann DW, Hoffmann R (1988) Preparation
and characterization of quantum-size titanium dioxide. J Phys
Basic Blue 9. The results clearly indicate the possibility of Chem 92:5196–5201
carrying out degradation of this type of organic compound 17. Li B, Wang X, Yan M, Li L (2002) Preparation and charac-
using powder materials prepared via the proposed method. terization of nano-TiO2 powder. Mater Chem Phys 78:184–188
18. Kim C-S, Moon BK, Park J-H, Chung ST, Son S-M (2003)
Acknowledgments This work was supported by Polish National Synthesis of nanocrystalline TiO2 in toluene by a solvothermal
Centre of Science research Grant No. 2011/01/B/ST8/03961. route. J Cryst Growth 254:405–410
19. Wang C, Deng Z-X, Zhang G, Fan S, Li Y (2002) Synthesis of
Open Access This article is distributed under the terms of the nanocrystalline TiO2 in alcohols. Powder Technol 125:39–44
Creative Commons Attribution 4.0 International License (http:// 20. Kang M, Kim B-J, Cho SM, Chung C-H, Kim B-W, Han GY,
creativecommons.org/licenses/by/4.0/), which permits unrestricted Yoon KJ (2002) Decomposition of toluene using an atmospheric
use, distribution, and reproduction in any medium, provided you give pressure plasma/TiO2 catalytic system. J Mol Catal A-Chem
appropriate credit to the original author(s) and the source, provide a 180:125–132
link to the Creative Commons license, and indicate if changes were 21. Kominami H, Kohno M, Takada Y, Inoue M, Inui T, Kera Y
made. (1999) Hydrolysis of titanium alkoxide in organic solvent at high
temperatures: a new synthetic method for nanosized, thermally
stable titanium(IV) oxide. Ind Eng Chem Res 38:3925–3931
22. Kolen’ko YV, Burukhin AA, Churagulov BR, Oleynikov NN
References (2003) Synthesis of nanocrystalline TiO2 powders from aqueous
TiOSO4 solutions under hydrothermal conditions. Mater Lett
1. Wypych G (1999) Handbook of fillers, 2nd edn. ChemTec Pub- 57:1124–1129
lishing, Toronto 23. Milea CA, Bogatu C, Duţă A (2011) The influence of parameters
2. Braun JH (1997) Titanium dioxide a review. J Coat Tech in silica sol–gel process. Bull Transilv Univ Braş Ser I Eng Sci
69:59–72 4:59–66

123
J Sol-Gel Sci Technol

24. Brinker CJ, Scherer GW (1990) Sol–Gel science: the physics and modified by methacrylate groups. J Sol–Gel Sci Technol
chemistry of sol–gel processing. Academic Press, Boston 43:21–26
25. Chan Y-L, Pung S-Y, Sreekantan S (2013) Degradation of or- 41. Ricci GP, Rocha ZN, Nakagaki S, Castro KADF, Crotti AEM,
ganic dye using ZnO nanorods based continuous flow water pu- Calefi PS, Nassar EJ, Ciuffi KJ (2011) Iron-alumina materials
rifier. J Sol–Gel Sci Technol 66:399–405 prepared by the non-hydrolytic sol–gel route: synthesis, charac-
26. Wu L-A, Jiang X, Wu S, Yao R, Qiao X, Fan X (2014) Synthesis terization and application in hydrocarbons oxidation using hy-
of monolithic zirconia with macroporous bicontinuous structure drogen peroxide as oxidant. Appl Catal A-Gen 389:147–154
via epoxide-driven sol–gel process accompanied by phase 42. Matos MG, Pereira PFS, Calefi PS, Ciuffi KJ, Nassar EJ (2009)
separation. J Sol–Gel Sci Technol 69:1–8 Preparation of a GdCaAl3O7 matrix by the non-hydrolytic sol–gel
27. Ciesielczyk F, Przybysz M, Zdarta J, Piasecki A, Paukszta D, route. J Lumin 129:1120–1124
Jesionowski T (2014) The sol–gel approach as a method of 43. Attar AS, Ghamsari MS, Hajiesmaeilbaigi F, Mirdamadi S (2008)
synthesis of xMgOySiO2 powder with defined physicochemical Modifier ligands effects on the synthesized TiO2 nanocrystals.
properties including crystalline structure. J Sol–Gel Sci Technol. J Mater Sci 43:1723–1729
doi:10.1007/s10971-014-3398-1 44. You JH, Hsu KY (2010) Influence of chelating agent and reaction
28. Vrancken KC, Possemiers K, Voort PVD, Vansant EF (1995) time on the swelling process for preparation of porous TiO2
Surface modification of silica gels with aminoorganosilanes. particles. J Eur Ceram Soc 30:1307–1315
Colloid Sur A 98:235–241 45. Huang T, Huang W, Zhou C, Situ Y, Huang H (2012) Super-
29. Mark JE, Lee CYC, Bianconi PA (1995) Hybrid organic–inor- hydrophilicity of TiO2/SiO2 thin films: synergistic effect of SiO2
ganic composites. American Chemical Society, Washington and chase-separation-induced porous structure. Surf Coat Tech
30. Cerveau G, Corriu RJP, Lepeytre C, Mutin PH (1998) Influence 213:126–132
of the nature of the organic precursor on the textural and chemical 46. Chang JA, Vithal M, Baek IC, Seok SI (2009) Morphological and
properties of silsesquioxane materials. J Mater Chem phase evolution of TiO2 nanocrystals prepared from peroxo ti-
12:2707–2714 tanate complex aqueous solution: influence of acetic acid. J Solid
31. Corriu R (1998) A new trend in metal-alkoxide chemistry: the State Chem 182:749–756
elaboration of monophasic organic–inorganic hybrid materials. 47. Zhou C, Ouyang J, Yang B (2013) Retarded hydrolysis con-
Polyhedron 17:925–934 densing reactivity of tetra butyl titanate by acetylacetone and the
32. Corriu RJP, Leclerq D (1996) Recent developments of molecular application in dye sensitized solar cells. Mater Res Bull
chemistry for sol–gel processes. Angew Chem Int Ed 35:1420–1436 48:4351–4356
33. Shea KJ, Loy DA, Webster O (1992) Arylsilsesquioxane gels and 48. Cullity BD (1978) Elements of X-ray diffraction. Addison
related materials. New hybrids of organic and inorganic net- Wesley Publishing Company Inc, Menlo Park
works. J Am Chem Soc 114:6700–6710 49. Zhang W, Chen S, Yu S, Yin Y (2007) Experimental and theo-
34. Jackson CL, Bauer BJ, Nakatami AI, Barnes J (1996) Synthesis retical investigation of the pH effect on the titania phase trans-
of hybrid organic–inorganic materials from interpenetrating formation during the sol–gel process. J Cryst Growth
polymer network chemistry. Chem Mater 8:727–733 308:122–129
35. Nassar EJ, Ciuffi KJ, Ribeiro SJL, Messaddeq Y (2003) 50. Chen H-J, Wang L, Chiu W-Y (2007) Chelation and solvent
Europium incorporated in the silica matrix obtained by sol–gel: effect on the preparation of titania colloids. Mater Chem Phys
luminescent materials. Mat Res 6:557–562 101:12–19
36. Beari F, Brand M, Jenkner P, Lehnert R, Metternich HJ, Mon- 51. Djellabi R, Ghorab MF, Cerrato G, Morandi S, Gatto S, Oldani V,
kiewicz J, Siesler H (2001) Organofunctional alkoxysilanes in Di Michele A, Bianchi CL (2015) Photoactive TiO2-montmoril-
dilute aqueous solution: new accounts on the dynamic structural lonite composite for degradation of organic dyes in water.
mutability. J Orgmomet Chem 625:208–216 J Photochem Photobiol A 295:57–63
37. Nassar EJ, Neri CR, Calefi PS, Serra OA (1999) Functionalized 52. Watson SS, Beydoun D, Scott JA, Amal R (2003) The effect of
silica synthesized by sol–gel process. J Non-Cryst Solids preparation method on the photoactivity of crystalline titanium
247:124–128 dioxide particles. Chem Eng J 95:213–220
38. Stöber W, Fink A, Bohn E (1968) Controlled growth of 53. Mills A, Lee SK, Lepre A (2003) Photodecomposition of ozone
monodisperse silica spheres in the micron size range. J Colloid sensitized by a film of titanium dioxide on glass. J Photochem
Interface Sci 26:62–69 Photobiol A 155:199–205
39. Papacı́dero AT, Rocha LA, Caetano BL, Molina EF, Sacco C, 54. Choi H, Stathatos E, Dionysiou DD (2006) Sol–gel preparation of
Nassar EJ, Martinelli Y, Mello C, Nakagaki S, Ciuffi KJ (2006) mesoporous photocatalytic TiO2 films and TiO2/Al2O3 composite
Preparation and characterization of spherical silica–porphyrin membranes for environmental applications. Appl Catal B-Envi-
catalysts obtained by the sol–gel methodology. Colloid Surf A ron 63:60–67
275:27–35
40. Nassar EJ, Nassor ECO, Ávila LR, Pereira PFS, Cestari A, Luz
LM, Ciuffi KJ, Calefi PS (2007) Spherical hybrid silica particles

123

You might also like