Sei sulla pagina 1di 7

Powder Technology 253 (2014) 2228

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

Inuence of pH on the formulation of TiO2 nano-crystalline powders with


high photocatalytic activity
Andreia Molea a, Violeta Popescu a,b,, Neil A. Rowson c, Adrian M. Dinescu b
a
b
c

Technical University of Cluj-Napoca, Faculty of Material and Environmental Engineering, Physics and Chemistry Department, No.103-105 Muncii avenue, 400641 Cluj-Napoca, Romania
National Institute for Research and Development in Microtechnologies, IMT, 126A Erou Iancu Nicolae Street, 077190, Bucharest, Romania
University of Birmingham, School of Chemical Engineering, Edgbaston, Birmingham B15 2TT, United Kingdom

a r t i c l e

i n f o

Article history:
Received 15 May 2013
Received in revised form 16 September 2013
Accepted 21 October 2013
Available online 29 October 2013
Keywords:
Titanium dioxide
Hydrolysis
pH effect
Catalysis
Photodegradation process
Methylene Blue dye

a b s t r a c t
This paper describes the effect of synthesis conditions on the formation of anatase and rutile crystalline phases
and photocatalytic activity of synthesised TiO2 phase. The synthesised powders were characterised by X-ray
diffraction, Raman microscopy, Scanning Electron Microscopy and UVVis spectroscopy. Using these characterisation techniques, the structural, morphological and optical properties as a function of formulation pH were
determined. Since photocatalysis is a surface process, the mass surface charge of the powders was also measured
using a Faraday Cage connected to an electrometer. The structural, morphological, optical and surface properties
were correlated with the photocatalytic activity of the formulated TiO2 powders.
The inuence of synthesis condition on the photocatalytic activity of TiO2 powders was determined by the
degradation of Methylene Blue dye under both UV-A and visible light.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Titanium dioxide (TiO2) has potential applications in environmental
elds such as wastewater treatment and more specically in cases
where the water has been contaminated with azo dyes from textile
and oil spillages [16]. Titanium dioxide occurs in nature as anatase, rutile and brookite mineral phases. Anatase is stabilised by heat treatment
at 400600 C, whilst, rutile, the most thermodynamically stable crystalline phase of titanium dioxide, is stabilised at 900 C. Between 600
and 900 C, both anatase and rutile phases co-exist [7,8].
Photocatalysis is a surface process; therefore, structural, morphological and optical properties are critical parameters for controlling the photocatalytic activity of the synthesised materials [912].
According to Cassaignon et al. [13], using titanium tri-chloride as a
precursor, in an acid medium, at pH between 3 and 4, a rutile phase is
stabilised after 24 h at 60 C. In an alkaline medium, at pH N 6.5 and
the same synthesis conditions, anatase is the main phase (65%) accompanied by brookite.

Corresponding author at: Technical University of Cluj-Napoca, Faculty of Material and


Environmental Engineering, Physics and Chemistry Department, No.103-105 Muncii avenue, 400641 Cluj-Napoca, Romania.
E-mail addresses: andreia.molea@chem.utcluj.ro (A. Molea),
violeta.popescu@chem.utcluj.ro (V. Popescu), n.a.rowson@bham.ac.uk (N.A. Rowson),
andrian.dinescu@imt.ro (A.M. Dinescu).
0032-5910/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.powtec.2013.10.040

Anatase and brookite are transformed by heat treatment into rutile,


but the synthesis conditions inuence this transition temperature. As a
result, if the material is prepared under acidic conditions and after heat
treatment at moderate temperature, i.e. at 400 C, a rutile phase is
stabilised [14]. However, in alkaline environments, (pH N 7) an anatase
phase is stabilised, even at 800 C [15]., However, only a few studies
have characterised the formulation of TiO2 powders at different pHs
with respect to photocatalytic activity [16].
According to some scientic literature, an anatase phase has the
highest photocatalytic activity [5,15], whereas other researches have indicated that a mixture of crystalline phases such as anatase and rutile
[1,17,18] or anatase, brookite and rutile [19] exhibits a higher photocatalytic activity than pure anatase. Xie et al. [18] studied the photocatalytic activity of TiO2 catalysts containing various rutile mass fractions using
the degradation of benzene under 24 mW/cm2 UV radiation. The sample with a higher concentration of rutile exhibits the highest photocatalytic activity, due to the lower potential of the conduction band of
rutile compared to anatase, giving more holes on the anatase surface
for oxidation reactions. Lopez et al. [19] studied the photocatalytic activity of TiO2 synthesised at various pHs. Also the authors [19] demonstrated that the photodegradation process depends on the energy band gap
(Eg). Bulk anatase has Eg = 3.2 eV, while rutile has Eg = 3 eV. The
highest activity corresponds to catalysts that contained anatasebrookiterutile and anataserutile, respectively. This correlates to the lowest
energy band gap of the rutile crystalline phase compared with the energy band gap of pure anatase phase. As such, the material could absorb a

A. Molea et al. / Powder Technology 253 (2014) 2228

23

Fig. 1. XRD patterns of TiO2 powders synthesised at different pHs and heat treatment at 400 C, 1 h ( anatase, rutile).

2. Materials and method

both UV-A and visible radiation. The radiation was emitted by two
lamps, a 6 W lamp with emission in 300400 nm UV-A wavelength
and a 9 W lamp with emission of visible wavelengths between 400
and 700 nm. A volume of 50 ml of photocatalyst suspension, which
contains 0.01 g catalyst, was mixed with 50 ml of a solution containing
5.5 103 mg/ml MB. The experiments took place at the natural pH
of the MB solution (pH 6.8). The mixture was maintained in dark conditions for 1 h, in order to establish the adsorption/desorption equilibrium, followed by irradiation for 300 min with both UV-A and visible
radiation. The intensity of the radiation was measured with a Mavolux
5032C lux meter which gave a value of 0.183 mW/cm2 at the solution
surface. A sample was prepared without catalyst (blank sample) for
benchmarking purposes.

2.1. Synthesis of TiO2 powders at different pH

2.3. Characterisation

TiO2 powders were synthesised by the hydrolysis of a titanium trichloride precursor. The solution was prepared from 5 ml TiCl3 (10% containing 15% HCl - Merck) and 5 ml hydrogen peroxide 3%, added into
5 ml distilled water. Hydrogen peroxide was used in order to achieve
the rapid oxidation of titanium. The initial solution pH was 1. The pH
was adjusted with ammonia solution (NH4OH ~25%) to 3, 8.5 and 10.5,
respectively. The solutions were stirred until a white precipitate of titanium hydroxide was formed. The precipitates were ltered, washed several
times with distilled water and subjected to heat treatment for 1 h at 400
C.

The titanium dioxide powders were characterised by X-ray diffraction


in order to determine the structural properties of the samples. A Bruker
AXS D8 diffractometer working with CuK radiation ( = 1.5406 )
was used. Structural parameters were calculated using the Powder Cell
software [21].
Raman spectra of specimens were registered using a WiTec Alpha
300 R (LOT Oriel, UK) operating a 0.3 W single frequency 785 nm
diode laser (Toptica Photonics, Germany) and an Acton SP2300 triple
grating monochromatic/spectrograph (Princeton Instruments, USA).
Confocal Raman spectroscope was used to determine the characteristic
TiO bonds for the crystalline phases.
Scanning electron microscopy was utilised to characterise the TiO2
powder morphology. The samples were examined using a eld emission
scanning electron microscope (FE-SEM) Raith e_Line with in-lens
electron detection capabilities.

wider wavelength range of the incident radiation, generating electron


hole pairs.
In this study, nanocrystalline TiO2 powders were synthesised by a
hydrolysis technique using titanium tri-chloride, in an aqueous solution
obtained from distilled water in the presence of hydrogen peroxide.
Inorganic TiCl3 was used because, when using organic precursors as a
titanium source, it is difcult to remove any organic residues from the
particle surface [20]. The inuences of pH on structural, morphological
and optical properties were studied. The photocatalytic activity of the
synthesised powders was determined by the degradation of Methylene
Blue under low intensity UV-A light and visible radiation.

2.2. Photodegradation experiment


In order to assess the performance of the formulated TiO2 powders,
Methylene Blue (MB) dye was degraded, at room temperature, under

Table 1
Structural parameters calculated by Rietveld renement based on XRD data, using the Powder Cell software.
Sample

pH 3
pH 8.5
pH 10.5

Crystalline phases/
Fraction mass [%]

Average crystallite
size [nm]

Strain

Anatase

Anatase

Anatase

29.76
100
100

Rutile

70.24

10
16
13

Rutile

10.6

0.009549
0.003001
0.004261

Lattice parameters []
Rutile

0.000001

Anatase
PDF 21-1272

Rutile
PDF 21-1276

3.7852

9.5139

4.5933

2.9592

3.7787
3.7709
3.7684

9.4801
9.4559
9.4470

4.5854

2.9466

24

A. Molea et al. / Powder Technology 253 (2014) 2228

Fig. 2. Raman spectra of TiO2 powders synthesised at different pHs. Insert: Raman shift of anatase peak position from 144 cm1.

A Faraday Cage connected to an electrometer, model Keithley 6514


was used to measure the surface charge of the materials.
Lambda 35 UVVis spectrophotometer with an integrated sphere
was used to determine the optical properties of the TiO2 powders. The
total transmittance of the samples was measured. Based on the absorption coefcient (), the energy band gaps of the samples were determined, using Tauc's relation [22]:

m
h A hEg

where: h is photon energy, A is a constant and m is an integer depending on the nature of electronic transitions. For the direct allowed transitions, m has a value of 1/2 while for indirect allowed transitions, m = 2
[22]. The absorption coefcient was calculated with the formula [23]:

Where: C0MB is the initial concentration of Methylene Blue and CMB is the
concentration of Methylene Blue at a certain irradiation time.
The kinetics of the degradation process of Methylene Blue was studied. The rate constant, k, was obtained by plotting the natural logarithm
of the ratio between initial concentration and the concentration at a certain irradiation time of Methylene Blue versus irradiation time assuming a rst order reaction [26,27]:
0

ln

cMB
kt
cMB

3. Results and discussion


3.1. X-ray diffraction

103 A 
2:303 
cl

Where: A is the absorbance, is the TiO2 bulk density ( = 3.84 [g/cm3]),


c is the concentration of the TiO2 suspensions (c = 1.2 103 [g/cm3]),
and l is the path length (l = 1 [cm]).
The Urbach energy (EU) has been estimated based on the following
equation [24]:
ln 0

1
h
Eu

Where: is the absorption coefcient, 0 is a constant. The Eu was determined from the inverse of the slope of the linear portion of the plot
ln versus h of the following equation [24]:The variation of MB concentration as a function of irradiation time, under both UV-A light and
visible irradiation, was determined using the UVVis spectroscopy
based on a calibration curve. The efciency of the degradation process
was calculated with the relation [25]:
Efficiency %

C 0Mb C Mb
 100
C 0Mb

The TiO2 diffraction patterns (Fig. 1) revealed that after heat treatment, in an acid medium, anatase and rutile crystalline phases were obtained, but rutile is the predominant phase, even if the temperature of
heat treatment was 400 C. In an alkaline medium environment however, only an anatase phase was stabilised. According to some scientic literature [13], at pH conditions between 2.5 and 4.5, the transformation
of titanium tri-chloride occurs by oxidation of the precursor into a rutile
intermediate
crystalline phase due to the formation of a Ti(OH)(OH)2+
5
compounds. At pH N 4.5, the oxidation of the precursor is very fast,
leading to the formation of a Ti(OH)3 + x compound, which, after 24 h
at 60 C is transformed into an anatase phase [13]. The average crystallite size (Table 1) increased with increase of the pH, from 10 nm for the
sample synthesised in pH 3 (anatase phase) to 16 nm for the sample
synthesised at pH 8.5 and then the crystallite size tends to decrease to
13 nm, for the sample synthesised at pH 10.5. This has been attributed
to an increase of the micro-strain. According to Shao et al. [7], a decrease
of crystallite size is linked to an increase of the number of defects, thus
leading to an increase in the number of lattice deformations. Parameter
cell calculations indicate narrowing of the elementary cell due to the
oxygen vacancies or site-disorder [28]. In an acid environment, the crystallite size was smaller (for rutile phase) because the acid acts as an

A. Molea et al. / Powder Technology 253 (2014) 2228

25

3.2. Raman microscopy


Fig. 2 shows the Raman signature in the range of 80700 cm1 of
TiO2 samples obtained at different pH, after heat treatment at 400 C
for 1 h. The symmetric model of a tetragonal anatase phase was identied at ~144 cm1 (Eg), 197 cm1 (Eg), 398 cm1 (B1g), 515 cm1
(B1g) and 638 cm1 (Eg) [29] for the samples synthesised in an alkaline
medium. However, for the sample prepared in an acid medium, fundamental vibration modes for rutile phase at 446 cm1 and 609 cm1
were also observed [30]. The Raman peak position changed from
144 cm1 as presented in Fig. 3. It was noted that the sample synthesised in acid medium exhibits a higher wavenumber (148 cm1). The
peak shift can be associated with small crystallite size and a higher
value of micro-strain [31]. An increased broadening of the peak due to
the quantum size effect was also noticed [29,31].

3.3. SEM microscopy


The SEM images of the TiO2 powders are presented in Fig. 3. At pH 3,
most of the crystals have elongated shapes with a length ranged between 20 and 44 nm and the cross section between 11 and 14 nm. Crystal size is quite uniform and the agglomeration tendency is small. The
sample obtained at pH of 8.5 is formed from nanoparticles smaller
than 20 nm with irregular shapes. The increase of the pH to 10.5 assured
the formation of well formed particles of 23 to 53 nm, with an average
of 32 nm.

3.4. Surface charge


The surface charge of the catalyst also plays a signicant role on the
degradation of the dye. It is known that MB dye has a positively charged
surface generated from a nitrogen centre on its organic framework [32].
Following measurement of the electrical charge of the particles it was
observed that for all the TiO2 samples, the surfaces are negatively
charged. The TiO2 prepared at pH 3 had a surface charge of 0.25
nC/g, whilst for TiO2 formulated at pH 8 and 10.5, the surface charge
was 0.15 nC/g and 0.1 nC/g. The value of negative electric charge
of the particles decreased with the increase of the pH of the solution.

3.5. UVVis spectroscopy

Fig. 3. SEM micrographs of TiO2 powders, formulated at a) pH 3, b) pH 8.5 and c) pH 10.5,


respectively.

electrolyte and prevents the particle growth and/or agglomeration


through electrostatic repulsion [9]. Y.C. Lee et al. [10] have obtained similar results.

Fig. 4 shows the UVVis absorption spectra of TiO2, obtained at different pH levels. With an increase of the pH from 8.5 to 10.5, a blue
shift of the absorption edge can be noticed due to the decrease of the
crystallite size from 16 nm to 13 nm. A higher absorption edge was
noted in the sample prepared at low pH due to the rutile content of
the sample, which has an absorption band at higher wavelengths compared with the anatase phase [19].
The determination of the energy band gap using Tauc's relationship
(Eq. (1)) [22] is presented in Fig. 5. It can be observed that the energy
band gap increased from 3.25 eV for the sample synthesised in an acid
medium to 3.31 eV for the sample prepared at pH 8.5 due to the rutile
content of the sample synthesised at pH 3. Rutile has a lower energy
band gap when compared to pure anatase [11]. In the case of the samples obtained in an alkaline medium, the energy band gap increased
from 3.31 eV for the sample prepared at pH 8.5 to 3.39 eV for the sample prepared at pH 10.5. The increase in Eg can be correlated to the
reduction of the crystallite size that determined quantum size effect,
which induce a blue shift of the absorption edge in the optical absorbance [12].
The plots used for the determination of Urbach energy are presented
in the insert of Fig. 5. Urbach energy decreased with increase of pH
(Table 2) being linked to the decrease of the micro-strain calculated
from XRD data.

26

A. Molea et al. / Powder Technology 253 (2014) 2228

Fig. 4. Absorption spectra of TiO2 samples synthesised at different pHs, after heat treatment at 400 C for 1 h.

3.6. Photodegradation process


The results of the photo degradation process for an MB dye in
the presence of the TiO2 suspensions, under both UV-A light and
visible irradiation, are shown in Fig. 6. It was observed that the
photodegradation process of MB took place faster in the presence of
TiO2, synthesised at pH 3 but decreased for the sample synthesised at
pH 8.5 and 10.5. The increase of pH gave an obvious decrease of the
photodegradation efciency that can be correlated with the increase of
the band gap.

Since MB is a cationic dye and has a positively charged surface, the


molecules are attracted on TiO2 catalyst surface negatively charged.
The attraction is stronger in the case of the catalyst formulated at
pH 3 also explaining the highest photoactivity of the sample and the
higher value of the negative charge.
The efciency of the photodegradation process increased from 17%
in the presence of TiO2 synthesised at pH 10.5 to 47% in the presence
of TiO2 synthesised at pH 3 after 300 min of irradiation. These results
are in good agreement with literature data [19]. No activity was detected for the blank sample.

Fig. 5. The determination of the energy band gap of TiO2 powders. Insert - determination of Urbach energy.

A. Molea et al. / Powder Technology 253 (2014) 2228


Table 2
Energy band gap and Urbach energy of the TiO2 samples synthesised at different pHs.
Sample

Energy band gap [eV]

Urbach energy [meV]

TiO2 pH 3 treated at 400 C, 1 h


TiO2 pH 8.5 treated at 400 C, 1 h
TiO2 pH 10.5 treated at 400 C, 1 h

3.25
3.31
3.39

980
551
620

At low concentrations of dye, the photodegradation reactions exhibited rst order kinetics mechanism [26,27]. Fig. 7 shows that the photo
degradation reaction of MB in the presence of TiO2 synthesised at various pH exhibit rst order kinetics. By plotting the ln(CMB/C0MB) versus irradiation time, the rate constant, k, was determined. This decreased
from 2.27 103 [min1] in the presence of TiO2 synthesised at pH 3
to 0.73 103 [min1] for the sample synthesised at pH 10.5. The kinetic model shows that the sample synthesised at pH 3 has a photocatalytic
activity three times higher than the sample synthesised at pH 10.5 due to
the lower energy band gap [19].
It should be noted that the TiO2 samples exhibit photocatalytic activity
even if the radiation intensity of the lamp was very low (0.183 mW/cm2).

4. Conclusion
Titanium dioxide powders were successfully synthesised by a hydrolysis method, using a titanium tri-chloride inorganic precursor. The
inuence of pH on the formation of TiO2 crystalline phases and on photocatalytic activity was studied. Based on XRD measurements it was
established that at high level of the pH, only the anatase phase of TiO2
has been obtained whilst under acidic conditions rutile and anatase
co-exist, but rutile is the predominant phase. TiO2 samples synthesised
at various pHs exhibited a trend towards higher wavenumber of the
Raman peak from 144 cm1, due to the quantum size effect, which

27

implies also an increase of the micro-strain. SEM images reveal that


the samples have the particle size between 20 and 50 nm.
The photodegradation experiments indicated that all the TiO2
samples exhibit some photocatalytic activity, even if the intensity of
the lamp radiation was low. However the TiO2 sample synthesised at
pH 3, which contained both anatase and rutile crystalline phases, exhibited the highest photocatalytic activity due to the lower energy band
gap and higher value of the negative charge on the surface. The efciency of Methylene Blue photo degradation in the presence of TiO2 prepared at pH 3 was 47% under both UV-A light and visible irradiation,
after 300 min.

Acknowledgments
This paper was supported by the project SIDOC, contract no.
POSDRU/88/1.5/S/60078 and project Human Resource Development
by Postdoctoral Research on Micro and Nanotechnologies, Contract
POSDRU/89/1.5/S/63700.
The authors acknowledge Gabriela Buda from Technical University
of Cluj-Napoca, for surface charge measurements. The authors would
also like to thank to Jacqueline Deans from School of Chemistry, University of Birmingham for assistance during XRD measurements and to
Dr James Bowen from Laboratory of Advanced Materials 2, School of
Chemical Engineering, University of Birmingham for assistance with
Raman microscopy. The equipment used in XRD measurement was
utilised through the Science City Advanced Materials Project: Creating
and Characterising Next Generation Advanced Materials, with support
from Advantage West Midlands (AWM) and part funded by the
European Regional Development Fund (ERDF). The Confocal Raman Microscope used in this research was obtained, through the Birmingham
Science City: Innovative Uses for Advanced Materials in the Modern
World (West Midlands Centre for Advanced Materials Project 2), with
support from Advantage West Midlands (AWM) and part funded by
the European Regional Development Fund (ERDF).

Fig. 6. Photodegradation of Methylene Blue in the presence of TiO2 catalysts. Insert - the efciency of photodegradation process, under low UV-A and visible irradiation, after 300 min.

28

A. Molea et al. / Powder Technology 253 (2014) 2228

Fig. 7. The rst order kinetics of the degradation of Methylene Blue dye with irradiation time.

References
[1] Y. Lin, C. Ferronato, N. Deng, J.-M. Chovelon, Study of benzylparaben photocatalytic
degradation by TiO2, Appl. Catal. B Environ. 104 (2011) 353360.
[2] X. Zeng, Y.X. Gan, E. Clark, L. Su, Amphiphilic and photocatalytic behaviors of TiO2
nanotube arrays on Ti prepared via electrochemical oxidation, J. Alloys Compd.
509 (2011) L221L227.
[3] A. Molea, V. Popescu, N.A. Rowson, Effects of I-doping content on the structural, optical and photocatalytic activity of TiO2 nanocrystalline powders, Powder Technol.
230 (2012) 203211.
[4] X.-L. Peng, M.-M. Yao, F. Li, X.-H. Sun, Microstructures and photocatalytic
properties of S doped nanocrystalline TiO2 lms, Part. Sci. Technol. 30 (2012)
8191.
[5] Y. Liu, L. Hua, S. Li, Photocatalytic degradation of Reactive Brilliant Blue KN-R by
TiO2/UV process, Desalination 258 (2010) 4853.
[6] Y.Y. Hsu, T.L. Hsiung, H.P. Wang, Y. Fukushima, Y.L. Wei, J.E. Chang, Photocatalytic
degradation of spill oils on TiO2 nanotube thin lms, Mar. Pollut. Bull. 57 (2008)
873876.
[7] Y. Shao, D. Tang, J. Sun, Y. Lee, W. Xiong, Lattice deformation and phase transformation from nano-scale anatase to nano-scale rutile TiO2 prepared by sol-gel technique, China Part. 2 (2004) 119123.
[8] G. Li, L. Li, J. Boerio-Goates, B.F. Woodeld, High purity anatase TiO2 nanocrystals:
near room-temperature synthesis, grain growth kinetics, and surface hydration
chemistry, J. Am. Chem. Soc. 127 (2005) 86598666.
[9] K. Yu, J. Zhao, Y. Guo, X. Ding, H. Bala, Y. Liu, Z. Wang, Solgel synthesis and hydrothermal processing of anatase nanocrystals from titanium n-butoxide, Mater. Lett.
59 (2005) 25152518.
[10] Y.C. Lee, Y.J. Jung, P.Y. Park, K.H. Ko, Preparation of TiO2 powder by modied
two-stage hydrolysis, J. Sol-Gel Sci. Technol. 30 (2004) 2128.
[11] C.K. Lee, D.K. Kim, J.H. Lee, J.H. Sung, I. Kim, K.H. Lee, J.W. Park, Y.K. Lee, Preparation
and characterization of peroxo titanic acid solution using TiCl3, J. Sol-Gel Sci.
Technol. 31 (2004) 6772.
[12] S. Lee, I.-S. Cho, J.-H. Noh, K.S. Hong, G.S. Han, H.S. Jung, S. Jeong, C. Lee, H. Shin, Correlation of anatase particle size with photocatalytic properties, Phys. Status Solidi A
207 (10) (2010) 22882291.
[13] S. Cassaignon, M. Koelsch, J.-P. Jolivet, From TiCl3 to TiO2 nanoparticles (anatase,
brookite and rutile): thermohydrolysis and oxidation in aqueous medium, J. Phys.
Chem. Solids 68 (2007) 695700.
[14] X. Bokhimi, A. Morales, F. Pedraza, Crystallography and crystallite morphology of rutile synthesized at low temperature, J. Solid State Chem. 169 (2002)
176181.
[15] A.L. Castro, M.R. Nunes, A.P. Carvalho, F.M. Costa, M.H. Florencio, Synthesis of anatase TiO2 nanoparticles with high temperature stability and photocatalytic activity,
Solid State Sci. 10 (2008) 602606.
[16] Z. Chen, G. Zhao, H. Li, G. Han, B. Song, Effects of water amount and pH on the crystal
behavior of a TiO2 nanocrystalline derived from a solgel process at a low temperature, J. Am. Ceram. Soc. 92 (2009) 10241029.

[17] M.A. Behnajady, H. Eskandarloo, N. Modirshahla, M. Shokri, Investigation of the


effect of solgel synthesis variables on structural and photocatalytic properties of
TiO2 nanoparticles, Desalination 278 (2011) 1017.
[18] H. Xie, L. Zhu, L. Wang, S. Chen, D. Yang, L. Yang, G. Gao, H. Yuan, Photodegradation
of benzene by TiO2 nanoparticles prepared by ame CVD process, Particuology 9
(2011) 7579.
[19] T. Lopez, R. Gomez, E. Sanchez, F. Tzompantzi, L. Vera, Photocatalytic activity in the
2,4-dinitroaniline decomposition over TiO2 sol-gel derived catalysts, J. Sol-Gel Sci.
Technol. 22 (2001) 99107.
[20] Z. Abbas, J. Perez Holmberg, A.K. Hellstriom, M. Hagstrom, J. Bergenholtz, M.
Hassellov, E. Ahlberg, Synthesis, characterization and particle size distribution of
TiO2 colloidal nanoparticles, Colloids Surf., A 384 (2011) 254261.
[21] W. Kraus, G. Nolze, POWDER CELL - a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns,
J. Appl. Crystallogr. 29 (1996) 301303.
[22] J. Tauc, R. Grigorovici, A. Vancu, Optical properties and electronic structure of amorphous germanium, Phys. Status Solidi 15 (1966) 627637.
[23] N. Serpone, D. Lawless, R. Khairutdinovt, Size effects on the photophysical properties
of colloidal anatase Ti02 particles: size quantization or direct transitions in this indirect semiconductor? J. Phys. Chem. 99 (1995) 1664616654.
[24] S.H. Kang, J.W. Lim, H.S. Kim, J.Y. Kim, Y.H. Chung, Y.E. Sung, Photo and electrochemical characteristics dependent on the phase ratio of nanocolumnar structured TiO2
lms by RF magnetron sputtering technique, Chem. Mater. 21 (2009) 27772788.
[25] C.H. Ao, S.C. Lee, Indoor air purication by photocatalyst TiO2 immobilized on an activated carbon lter installed in an air cleaner, Chem. Eng. Sci. 60 (2005) 103109.
[26] M. Mohseni, F. Taghipour, Experimental and CFD analysis of photocatalytic gas
phase vinyl chloride (VC) oxidation, Chem. Eng. Sci. 59 (2004) 16011609.
[27] N. Talebian, M.R. Nilforoushan, Comparative study of the structural, optical and photocatalytic properties of semiconductor metal oxides toward degradation of methylene blue, Thin Solid Films 518 (2010) 22102215.
[28] L. Borgese, E. Bontempi, M. Gel, L.E. Depero, P. Goudeau, G. Geandier, D. Thiaudiere,
Microstructure and elastic properties of atomic layer deposited TiO2 anatase thin
lms, Acta Mater. 59 (2011) 28912900.
[29] S. Balaji, Y. Djaoued, J. Robichaud, Phonon connement studies in nanocrystalline
anatase-TiO2 thin lms by micro Raman spectroscopy, J. Raman Spectrosc. 37 (2006)
14161422.
[30] R. Parra, M.S. Goes, M.S. Castro, E. Longo, P.R. Bueno, J.A. Varela, Reaction pathway to
the synthesis of anatase via the chemical modication of titanium isopropoxide
with acetic acid, Chem. Mater. 20 (2008) 143150.
[31] A. Golubovic, M. Scepanovic, A. Kremenovic, S. Askrabic, V. Berec, Z. Dohcevic-Mitrovic,
Z.V. Popovic, Raman study of the variation in anatase structure of TiO2 nanopowders
due to the changes of solgel synthesis conditions, J. Sol-Gel Sci. Technol. 49 (2009)
311319.
[32] A. Kurniawan, H. Sutiono, Y.-H. Ju, F.E. Soetaredjo, A. Ayucitra, A. Yudha, S. Ismadji,
Utilization of rarasaponin natural surfactant for organo-bentonite preparation:
application for methylene blue removal from aqueous efuent, Microporous
Mesoporous Mater. 142 (2011) 184193.

Potrebbero piacerti anche