Sie sind auf Seite 1von 8

J Sol-Gel Sci Technol (2011) 58:201–208

DOI 10.1007/s10971-010-2378-3

ORIGINAL PAPER

The influence of titanium dioxide phase composition on dyes


photocatalysis
Luminita Andronic • Diana Andrasi •
Alexandru Enesca • Maria Visa • Anca Duta

Received: 29 September 2010 / Accepted: 2 December 2010 / Published online: 14 December 2010
 Springer Science+Business Media, LLC 2010

Abstract A comparative study of TiO2 powders prepared 1 Introduction


by sol–gel methods is presented. Titanium tetraisopropoxide
was used as the precursor for the sol–gel processes. The Heterogeneous photocatalysis is a relatively novel method
effects of the annealing treatment on phase, crystallite size, used for removing organic pollutants (particularly, dyes) from
porosity and photodegradation of dyes (methyl orange and wastewaters resulted in the textile finishing industry [1, 2].
methylene blue) were studied. The phase structure, micro- Many photocatalysts have been reported to degrade
structure and surface properties of the films were characterized organic pollutants such as semiconductors: WO3, SnO2,
by using X-ray diffraction (XRD) and Atomic Force Micros- CdS, ZnO, Nb2O5, TiO2 [3–8], metal and nonmetal doped
copy (AFM). The X-ray diffraction was used for crystal phase semiconductors [9–14], coupled semiconductors: TiO2/
identification, for the accurate estimation of the anatase–rutile CdS, TiO2/SnO2, TiO2/ZnO, TiO2/WO3 [15–17].
ratio and for the crystallite size evaluation of each polymorph Titanium dioxide (TiO2) is frequently used for envi-
in the samples. It was found that the only TiO2 anatase phase of ronmental purification because of its many desirable
the synthesized TiO2 develops below 500 C, between 600 properties: non-toxicity, photochemical stability, and low
and 800 C the anatase coexist with rutile and above 800 C cost [18, 19]. It is usually used in two forms: particles
only the rutile phase was found in the samples. Attention has suspended in aqueous solution and thin films immobilized
been paid not only to crystal structures, but also to the porosity, on substrates. In most of the experiments concerning the
the particle size and the photocatalytic properties. However, TiO2 photocatalysis, the photocatalyst was used in the
the annealing temperature was found to have significant powder form and the recovery of powder was difficult to
influence on the photocatalytic properties. Different TiO2 execute. Thus, immobilization of the TiO2 semiconductor
doctor blade thin films were obtained mixing the sol gel particles is necessary for industrial applications.
powder (100% anatase) and TiO2 Aldrich with TiO2 Degussa The methods used for the synthesis of titanium dioxide
P25. The surfactant (Triton X100 or sodium dodecyl sulfate) powder include alkali precipitation, thermal decomposi-
affects the packing density of the particles during deposition tion, hydrothermal synthesis, sol–gel [20–22] and other
and the photocatalytic degradation efficiency of the dyes. The routes. Among all these procedures, preparation of TiO2 by
photocatalytic degradation kinetics of methyl orange and a sol–gel route remains one of the most attractive due to the
methylene blue using TiO2 thin film were investigated. less energetic conditions and the possibility of preparing
powders or thin films with controlled properties.
Keywords Titanium oxide  Sol–gel deposition  Preparing thin films from sol–gel powder can be done by
Photocatalysis  Dyes dipping or by doctor blade techniques. The doctor blade
technique requires no expensive apparatus, is flexible in
terms of substrate geometry, and can be used to obtain, in
L. Andronic (&)  D. Andrasi  A. Enesca  M. Visa  A. Duta optimized conditions, highly omogeneous surface surfaces
The Centre: Product Design for Sustainable Development,
in terms of morphology.
Transilvania University of Brasov, Eroilor 29,
500036 Brasov, Romania Titanium oxide crystallizes in three polymorphic forms:
e-mail: andronic-luminita@unitbv.ro anatase (tetragonal), rutile (tetragonal) and brookite

123
202 J Sol-Gel Sci Technol (2011) 58:201–208

(orthorhombic). The phase composition has a crucial – TiO2 Degussa P25 (with BET surface area 55 ± 15 m2/g;
impact on the photocatalytic activity. It is well known that average primary particle size around 30 nm, purity
anatase phase is found suitable for photocatalytic oxidation above 97% and a 80:20 anatase: rutile ratio),
of organic pollutants in water and air [1]. This effect has – TiO2 Aldrich (99.9% anatase, specific surface area
often been attributed to the larger surface area of anatase 48 m2 g-1 and a mean particle size of 1 lm),
compared to rutile, the higher affinity of anatase for – ethanol, acetylacetonate (C5H8O2, 99.9%, Alfa Aesar),
organic compounds, and slightly larger band gap (3.2 vs. Triton X100 (TX) (non-ionic surfactant, Scharlau
3.0 eV for rutile). Chemie), sodium dodecyl sulfate (SDS) (anionic sur-
The aim of this work was to optimize the polymorphs’ factant, Scharlau Chemie).
composition of TiO2 (anatase: rutil) efficient for dyes
degradation. The powders used were TiO2 prepare by sol–
2.2.2 Doctor blade deposition
gel and other two commercial powders: Aldrich (99.9%
anatase) and Degussa P25 (80% anatase, 20% rutile). The
For the doctor blade deposition of nanoporous TiO2 films,
effects of the annealing temperature on the crystal struc-
an aqueous colloid paste is formed by mixing TiO2
ture, morphology, porosity, and photocatalytic efficiency
powder (sol–gel or Aldrich with Degussa P25) with
towards degradation of methyl orange and methylene blue
3.6 mL of ethanol, 0.4 mL acetylacetone and 0.4 mL
are studied. Further, the effects of mixing two powders
surfactant 10 g/L (Triton X100 or SDS). The paste was
with different polymorph composition (sol–gel, Aldrich,
diluted by very slow addition of 2 mL of ethanol. Finally,
respectively Degussa P25) on the photocatalytic properties
the paste is smeared on glass substrate (1.5 9 2.5 cm2)
of the TiO2 films is investigated.
cleaned using ethanol, distilled water, acetone in succes-
sive sonication processes. After drying in air at 60 C for
about 2 min, the films were annealed in an oven at
2 Experimental
temperatures 500 C, for 3 h. The sol–gel powder was
mixed with Degussa P25 in a weight ratio 1:1, the sur-
2.1 Synthesis of TiO2 powder by sol–gel technique
factant used was Triton X100 (sample SG-P25(TX)) or
sodium dodecyl sulfate (sample SG-P25(SDS)). After
The TiO2 sol was synthesized by acid catalyzed sol–gel for-
drying in air at 60 C for about 2 min, the films were
mation method using 60 mL of 1 M HNO3 and 15 mL of
annealed in an oven at 400 C, for 3 h. The Aldrich
titanium tetra-isopropoxide following the hydrolysis reaction.
powder was mixed with Degussa P25 in a weight ratio
Titanium tetra-isopropoxide was added gradually to the
1:1, the surfactant used was Triton X100 (sample Aldrich-
aqueous solution of HNO3 solution under continuous stir-
P25(TX)).
ring for 3 h to produce a transparent sol. Subsequently, the
pH of the colloid solution was adjusted to 3 with the
addition of 1 M NaOH resulting in a turbid colloid. The pH 2.3 The catalyst characterization
adjustment was necessary to prevent the destruction of the
structure of adsorbent in the acidic media. The crystalline phase transformation temperature and
The mixed suspension was agitated by magnetically enthalpy were measured for the predominant amorphous
stirrer for another 2 h at room temperature, followed by powder dried at 100 C using a differential scanning cal-
filtration (Millipore 0.45 lm) and washings with de-ion- orimeter (PerkinElmer, DSC-2). The DSC and analysis of
ized water until the pH of the supernatant was about 6. the sample were performed from ambient temperature up to
The resulting TiO2 powder was dried in a furnace for 500 C with a heating rate of 5 C/min. During the mea-
1 h at 100 C and finally the powder was annealed at surements, pure nitrogen (N2) was used as a purging gas at
temperatures from 100 to 900 C for 3 h. a flow rate of 20 cm3/min.
The phase structure, microstructure and surface prop-
erties of the films were characterized by using X-ray
2.2 Thin films preparation
diffraction (XRD-Brucker D8 Discover diffractometer
with CuKa radiation) and Atomic Force Microscopy
2.2.1 Materials
(AFM-NT-MDT model NTGRA PRIMA EC). The images
were taken in semi-contact mode with ‘‘GOLDEN’’ silicon
The materials used for catalyst preparation were:
cantilever (NCSG10, force constant 0.15 N/m, tip radius
– TiO2 sol–gel powders annealed at 100–900 C for 3h; 10 nm). Scanning was conducted on three different places
for thin films preparation, based on the optimized (a certain area of 5 9 5 lm for each section) chosen ran-
results, the powder annealed at 500 C was used, domly at a scanning rate of 1 Hz.

123
J Sol-Gel Sci Technol (2011) 58:201–208 203

2.4 Photocatalytic experiments

The photocatalytic activity of the thin films was evaluated


using two testing dyes: methyl orange (C14H14N3NaO3S,
MO, Merck, reagent grade) and methylene blue
(C16H18N3S, MB, Fluka AG, Buchs SG, reagent grade). In
25 mL dye solution (initial concentration 0.0125 mM), the
powder (0.1 g) or film (sample of 1.5 9 2.5 cm2) was
added and then irradiated with three F18 W/T8 black light
tubes (Philips) (UVA light, typically 340–400 nm, with
kmax (emission) = 365 nm), for different periods of time.
The variation of the dyes concentration was analyzed
using UV–Vis spectrophotometer Perkin-Elmer Lambda 25
at the maximum absorption wavelength registered at
464 nm for MO, respectively 664 nm for MB using the
calibration curve (Fig. 1).
For comparison a blank sample, which contained only
MO or MB and no photocatalyst, was subjected to UV
irradiation in the photocatalytic reactor. In the absence of
the photocatalyst, the dyes solution (with and without
H2O2), was found to be stable under irradiation with UV
light. We could conclude that UV light does not directly
degrade the dye but activates the TiO2 (Fig. 1). Fig. 2 Differential scanning calorimetric curves of TiO2 sol–gel
powder

3 Results and discussions

3.1 Structural and morphological characterization The differential scanning calorimeter (DSC) results of
of the photocatalyst the sample (dried in a furnace for 1 h at 100 C) are shown
in Fig. 2. The exothermal peak in the range 240–310 C
Properties of TiO2 depend on the chemical composition, (Fig. 2a) can be attributed to the decomposition and
the crystalline structure, the phase composition, the crys- burning of organic by-product. These organics residues
tallization degree and the microstructure. may significantly affect the crystallization behavior and,
therefore, it is important their removal from the sample. It
was found that the sample treatment at 300 C in oxygen
atmosphere for 3 h reaches this goal and that the sample
still retains its predominant amorphous character.
The exothermal peak in the range 450–480 C (Fig. 2b)
is attributed to the complete phase transition of amorphous
TiO2 to anatase, which the XRD results also indicate.
Titanium oxide obtained by the sol–gel method is usu-
ally amorphous, and thermal treatments are necessary to
obtain the final materials with the desired structural char-
acteristics. The thermodynamic stability of the nanostruc-
tured phases can be tuned by controlling the crystallite size
and the energy of the exposed surfaces. Anatase is usually
more stable when the crystals are smaller than 30 nm while
rutile is the most favored phase at larger crystal size. The
annealing treatment contributes to crystal growth, trans-
forming anatase in rutile, at temperatures bellow the
regular polymorph transition (750 C). Therefore, it is
Fig. 1 The absorbance measurements of the dyes solution (methyl
orange and methylene blue), respectively, of the F18 W/T8 blacklight necessary to investigate how the phase transition is influ-
tubes enced with increasing annealing temperature.

123
204 J Sol-Gel Sci Technol (2011) 58:201–208

(‹101› peak at 2h = 25.4) and the strongest rutile (‹110›


peak at 2h = 27.5) reflections, respectively.
The average crystallite sizes of each phase were calcu-
lated from the Scherrer equation [24], at the full-width at
half-maximum (FWHM) of the peak presenting the highest
intensity and taking into account the instrument broaden-
ing; the crystallite increases with increasing the annealing
temperature (Table 1). It should be noted that the average
crystal size correspond to anatase (101) in the SG-500
powder was 15 nm, corresponding to the highest photo-
degradation efficiency.
The thin films surface morphology has a pronounced
effect on the photocatalytic activity. For a high photocat-
alytic activity, the thin film should have a porous structures
that can be can be controlled by adding surfactants (Triton
X100 or SDS). Quantitative measurements of the surface
roughness, using AFM (Figs. 4 and 5) suggest that larger
crystallites are likely to aggregate in more open structures
(SG-500, SG-900) and the polymorph composition has a
much lower influence. This is also confirmed by the films
developed using the highly regular Degussa powder, where
the lowest roughness is obtained, with a narrow voids
distribution, that can be slightly influenced by the surfac-
tant type.
By analyzing the AFM images, the distribution curves
of the voids, Fig. 5, allows to estimate the most likely
inter-particle voids size (calculated at the half width of
the peak): 160 ± 20 nm (SG-P25(TX)), 230 ± 14 nm
(SG-P25(SDS)), 260 ± 13 nm (SG-300), 350 ± 10 nm
(SG-500), 400 ± 11 nm (SG-900), 720 ± 10 nm (Aldrich-
Fig. 3 The XRD patterns of the TiO2 sol–gel powder annealing in P25(TX)) samples. As expected, the higher voids were
different temperature: a 100–600 C and b 600–900 C found for Aldrich-P25(TX) sample due to the average
primary particle size, of 1 lm, comparing to the Degussa
P25 powder (30 nm).
In Fig. 3 there are presents the XRD patterns obtained In all the cases, the voids are large, being able to
for the samples annealed at temperatures between 300 and accommodate any dye molecule with sizes lower than
900 C for 3 h. The samples annealed above 500 C 10 nm.
showed pronounced peaks indicating that the material was
fully crystalline. This is consistent with the DSC data, 3.2 Photodegradation activity of the TiO2 powders
which showed that the decomposition and crystallization and films
reactions are completed at temperatures above 470 C.
The XRD pattern of the sample annealed at 300-500 C The optimum contact time was found for methyl orange
shows only the anatase phase. Between 600 and 700 C, degradation at 30 min, using TiO2 Degussa P25 powder;
the rutile phase appears and coexists with the anatase the experimental conditions were set at: 4 g/L TiO2 pow-
phase. The single rutile phase formation was observed in der, MO initial concentration 0.0125 mM.
TiO2 sample annealed above 800 C. The photocatalytic activity of the samples annealed
The weight ratios of anatase and rutile in the sol gel between 300 and 900 C was observed by the degradation
samples were calculated using the following equation [23]: of a 0.0125 mM methyl orange solution under UV light,
1 measuring the absorption spectra for each, after 30 min
A% ¼  100 ð1Þ (Table 1).
1 þ 1:265IIA
R
As the results prove, the annealing temperature has a
where: A% is the weight fraction of the anatase phase, IA significantly affect on the methyl orange degradation, due
and IR represents the intensities of the strongest anatase to a complex of factors: phase composition, particle size,

123
J Sol-Gel Sci Technol (2011) 58:201–208 205

Table 1 The anatase and rutile percentage and crystallite sizes of rutile and anatase correlated with MO photodegradation
Sample Annealing Crystallite size [nm] Crystalline phases [%] MO degradationa
temperature [C]
Anatase Rutile Anatase Rutile

SG-300 300 8 – 100 – 35


SG-400 400 8 – 100 – 40
SG-500 500 15 – 100 – 41
SG-600 600 32 45 40 60 37
SG-700 700 34 50 35 65 23
SG-800 800 – 50 – 100 9
SG-900 900 – 34 – 100 4
a
Experimental conditions: MO initial concentration 0.0125 mM, TiO2 powder 4 g/L, irradiation time: 30 min

Fig. 4 The AFM image of TiO2

pore size, and roughness. As the annealing temperature


increase up to 700 C the photocatalytic degradation of
MO decreases as result of a different polymorph compo-
sition (increasing the rutile content), a decrease in the
crystallite size and a larger voids distribution. A slightly
higher MO photodegradation efficiency was found for the
sample annealed at 500 C and this powder was further
used for thin film preparation.
Thin-film coating of photocatalyst may solve the prob-
lems of leaching and separation. Hence, this technique may
be a viable one for the treatment of large volumes of textile
wastewater of low dyes concentration.
Firstly, tests for the photocatalytic activity of the dif-
ferent TiO2 films: P25, SG-P25(TX), SG-P25(SDS),
Fig. 5 The interparticle voids distribution Aldrich-P25(TX) were done.

123
206 J Sol-Gel Sci Technol (2011) 58:201–208

the packing density (crystallite aggregation) during depo-


sition, confirmed by the results in Fig. 5.
The surfactant at the solid–water interface could benefi-
cially affect the contaminant interaction with the catalyst
surface by enhancing the surface affinity, or adversely by
blocking the reactive sites. The enhancement of the degra-
dation efficiency of dyes on SG-P25(TX) might be due to
increased hydrophobic interactions between dye and TX100
molecules. Catalyst SG-P25(SDS) exhibited less MO deg-
radation than MB due to electrostatic repulsion between
SDS and MO molecules. Another reason could be the
amount of sodium cations resulted in the film when adding
SDS, with negative effect on the n-type TiO2 conduction.
By adding an electron acceptor (H2O2 30%–4 mL/L
dye solution) in the photocatalytic systems a 20…30%
enhancement in the photocatalytic efficiency of the TiO2
thin films was registered, the MB degradation being more
sensitive (Fig. 6).

3.3 Photodegradation kinetics

The kinetic investigations in heterogeneous photocatalysis


are developed for analyzing the reaction mechanism at
molecular level and for optimizing the process conditions.
The Langmuir–Hinshelwood (LH) kinetic is the mostly
used kinetic expression to explain the kinetics of the het-
erogeneous catalytic processes [25–28]. The Langmuir–
Fig. 6 Photodegradation efficiency of dyes (a methyl orange, Hinshelwood expression is given by equation:
b methylene blue) vs. TiO2 photocatalyst after 6 h of UV irradiation
ln c ¼ ln c0  kt ð2Þ
The photocatalytic activity of the TiO2 films was evalu- where c0 is the initial concentration of dyes aqueous
ated for the dyes (methyl orange-MO and methylene blue- solution, mM; c is the concentration of dyes, mM, mea-
MB) degradation efficiency at the initial concentration sured at the time t, minute; k is the apparent photodegra-
0.0125 mM (Fig. 6). Experiments were done at the oxygen dation rate constant, min-1.
saturation concentration for the system under stirring For the dyes photodegradation, kinetics calculations
(dye/UV/O2) and by adding an initial amount of hydrogen were done, Fig. 7 (catalyst: SG-P25(TX)) respectively
peroxide (4 ml/L) in the systems described as dye/UV/H2O2. Fig. 8 (catalyst: SG-P25(TX)).
The methyl orange dye is more difficult to degrade The rate constants in the dyes photodegradation reaction
probably due to the less adsorption/interaction of these on P25 and SG-P25(TX) are presented in Table 2. The high
molecules on the TiO2 surface. The stronger adsorption/ values of the regression coefficient (R2) for the linear fit
interaction between methylene blue (fully flat molecule) confirm that the dyes photodegradation follows the Lang-
and TiO2 surface leads to a higher photodegradation muir–Hinshelwood kinetic. These constant values are of
efficiency. same order of magnitude with those reported in literature
The highest MO photodegradation efficiency of the for organic pollutants photocatalysis [29].
mixed film SG-P25(TX) film can be mainly related to As shown in Fig. 7, the dyes degradation rate on SG-
the fact that the sol–gel (15 nm) and the Degussa P25 P25(TX) (Table 2), k1 = 0.00109 min-1 for MO, respec-
(30 nm) grains are of similar size and the average value tively, k2 = 0.001586 min-1 for MB, was higher than the
of the inter-particle voids are low (150 ± 75 nm) com- dyes degradation rate on P25 photocatalyst (Table 2), k3 =
paring to the other films. This cumulative conditions 0.00084 min-1 for MO, respectively, k4 = 0.001386
support the spatial charge separation and hindered elec- min-1 for MB, indicating that the SG-P25(TX) photocat-
tron–hole recombination. alyst could be recommendable for upscale applications,
The dyes removal efficiency was influenced by the having not only good efficiency but also convenient reac-
surfactant added during preparation. The surfactant affects tion rate.

123
J Sol-Gel Sci Technol (2011) 58:201–208 207

Fig. 7 The kinetics of dyes photodegradation (a methyl orange, Fig. 8 The kinetics of dyes photodegradation (a methyl orange,
b methylene blue) on the P25 and SG-P25(TX) films b methylene blue) on the SG-P25(SDS) film

Under the experimental conditions, the dyes photodeg- Table 2 The kinetics parameters for dyes photodegradation on P25
and SG-P25(TX)
radation on SG-P25(SDS) films follows a two step reaction
mechanism as presented in Table 3 and Fig. 8: in the first Catalyst/electron acceptor MO MB
240 min (I step) the slower reaction rates can be the result of -1
k [min ] R 2
k [min-1] R2
a slower active species generation (HO•) simultaneous with
side-reactions not leading to photo-degradation. The faster P25
step II (after 240 min) can be the result of decreasing these O2 0.00084 0.992 0.001386 0.9706
side-reactions (passivating the corners and edges on the H2O2 0.00159 0.9997 0.003049 0.9997
TiO2 surface) and/or an increased amount of HO• (co)gen- SG-P25(TX)
erated by the photo-degradation intermediates. (Fig. 8). O2 0.00109 0.9967 0.001586 0.9793
The selection of the optimum concentration may be H2O2 0.002042 0.9931 0.003707 0.9855
possible based on the rate constant in the photodegradation
reaction (k). These values, along with the regression
coefficients (R2) of the linear fittings for dyes phodegra- without oxidant addition, has the highest rate in the first
dation are shown in Table 3. The data show that with or step (0–240 min), recommending this as an optimum
without H2O2 the constant rates are in the same order of contact time for large scale applications.
magnitude thus, the amount of HO• is responsible for the For practical applications, the catalyst must conserve the
efficiency increase and not a change in mechanism. The activity over a long period of utilization. The photocata-
results also show that MB removal is faster due to lytic films were used for 3 times (18 h) in the same con-
the higher interaction between the dyes molecules and the ditions and the dyes photodegradation efficiency was
catalyst. With one exception, their removal leads in the first comparable. For the film with the best performance (SG-
240 min to slower reaction rates, probably as result of P25(TX)) the efficiency data after 6 h (first cycle), 12 h
adsorption or by-products resulted in photocatalysis that (second cycle), respectively, 18 h (third cycle) are pre-
degrade (much) slower. The exception, systems of MO sented in Fig. 9.

123
208 J Sol-Gel Sci Technol (2011) 58:201–208

Table 3 The kinetics parameters for dyes photodegradation on SG- Acknowledgments This paper is supported by the Sectoral Opera-
P25(SDS) tional Programme Human Resources Development (SOP HRD),
financed from the European Social Fund and by the Romanian
TiO2-SG-SDS Time [min] MO MB Government under the contract number POSDRU ID59323.
-1 2
k [min ] R k [min-1] R2

UV/O2 References
I 0–240 6.66E-04 0.9974 1.14E-03 0.9912
II 240–360 2.74E-04 0.9984 2.35E-03 1.0000 1. Choi HS, Al-Abed R, Dionysiou D, Stathatos E, Lianos P (2010)
In: Escobar IC, Schäfer AI (eds) Sustainable water for the future:
UV/H2O2
water recycling versus desalination. Elsevier, Amsterdam
I 0–240 7.84E-04 0.9963 2.36E-03 0.9996 2. Gaya UI, Abdullah AH (2008) J Photochem Photobiol C 9:1–12
II 240–360 3.09E-03 0.9898 3.21E-03 0.9903 3. Rehman S, Ullah R, Butt AM, Gohar ND (2009) J Hazard Mater
170:560–569
4. Torres Delgado G, Zúñiga Romero CI, Mayén Hernández SA,
Castanedo Pérez R, Zelaya Angel O (2009) Sol Energy Mater Sol
Cells 93:55–59
5. Yusta FJ, Hichman ML, Shamlian H (1997) J Mater Chem
8:1421–1427
6. Esteves A, Oliveira LCA, Ramalho TC, Goncalves M, Anastacio
AS, Carvalho HWP (2008) Catal Commun 10:330–332
7. Kim H, Senthil K, Yong K (2010) Mater Chem Phys
120:452–455
8. Janus M, Morawski AW (2007) Appl Catal. B 75:118–123
9. Li HB, Duan XC, Liu GC, Li LL (2008) Mater Res Bull 43:
1971–1981
10. Bettinelli M, Dallacasa V, Falcomer D, Fornasiero P, Gombac V,
Montini T, Roman L, Speghini A (2007) J Hazard Mater
146:529–534
11. Chen C, Wang Z, Ruan S, Zou B, Zhao M, Wu F (2008) Dyes
Pigm 77:204–209
12. Kosowska B, Mozia S, Morawski AW, Grzmil B, Janus M,
Fig. 9 The photocatalytic activity of SG-P25(TX) film after three Kalucki K (2005) Sol Energy Mater Sol Cells 88:269–280
cycles 13. Park H, Choi W (2004) J Phys Chem B 108:4086–4093
14. Li M, Zhou S, Zhang Y, Chen G, Hong Z (2008) Appl Surf Sci
254:3762–3766
4 Conclusions
15. Lin CF, Wu CH, Onn ZN (2008) J Hazard Mater 154:1033–1039
16. Wu CH (2004) Chemosphere 57:601–608
Nanocrystalline TiO2 catalysts with different contents of 17. Wang S, Zhang X, Cheng G, Jiang X, Li Y, Huang Y, Du Z
anatase and rutile phases were obtained by sol gel method. (2005) Chem Phys Lett 405:63–67
18. Fujishima A, Rao TN, Tryk DA (2000) J Photochem Photobiol. C
The sol–gel synthesis route resulted in TiO2 with good
1:1–9
structural properties such as homogeneity, controlled pore 19. Han F, Kambala VSR, Srinivasan M, Rajarathnam D, Naidu R
size distribution, and high porosity as well as enhanced (2009) Appl Catal. A 359:25–40
catalytic properties. Anatase phase was found to be the 20. Baorang L, Xiaohui W, Minyu Y, Longtu L (2003) Mater Chem
Phys 78:184–188
unique phase in the samples annealed below 500 C. Phase
21. Almquist CB, Biswas P (2002) J Catal 212:145–156
transformation of anatase into rutile was only observed in the 22. Mills A, Hill G, Bhopal S, Parkin IP, O’Neill SA (2003) J Pho-
sample synthesized by the sol–gel method when annealing tochem Photobiol. A 160:185–194
at/above 600 C. As the calcinations temperature increases, 23. Song GB, Liang JK, Liu FS, Peng TJ, Rao GH (2005) Thin Solid
Films 491:110–116
crystallite size also increased and the photocatalytic effi-
24. Burton AW, Ong K, Rea T, Chan IY (2009) Microporous Mes-
ciency decreased as expected. The use of surfactant provide oporous Mater 117:75–90
a higher pack density, increasing the available surface area. 25. Vinodgopal K, Wynkoop DE, Kamat PV (1996) Environ Sci
Our results suggest that a tailored mixture of anatase and Tech 30:1660–1666
26. Lachheb H, Puzenat E, Houas A, Ksibi M, Elaloui E, Guillard C,
rutile could be a good perspective for improving the dyes
Herrmann JM (2002) App Catal. B 39:75–90
degradation efficiency, if the mixture is adapted to the dyes 27. Tarasov VV, Barancova GS, Zaitsev NK, Dongxiang Z (2003)
composition. Process Saf Environ Prot 81:243–249
The kinetic characteristics of the photocatalytic degra- 28. Kumar KV, Porkodi K, Rocha F (2008) Catal Commun 9:82–84
29. Baran W, Adamek E, Makowski A (2008) Chem Eng J 145:
dation of dyes by a TiO2 thin film catalyst were experi-
242–248
mentally investigated. The kinetic characteristics were
ascertained to follow the Langmuir–Hinshelwood kinetic
mechanism.

123

Das könnte Ihnen auch gefallen