You are on page 1of 6

Applied Surface Science 258 (2012) 4328–4333

Contents lists available at SciVerse ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Synthesis and characterization of natural zeolite supported Cr-doped TiO2


photocatalysts
Cheng Wang a,b , Huisheng Shi a,∗ , Yan Li b
a
Key Laboratory of Advanced Civil Engineering Materials (Tongji University), Ministry of Education, 4800 Caoan Road, Shanghai 201804, China
b
Institute of Material Science and Engineering, Shijiazhuang University of Economics, 136 Huaian Road, Hebei Shijiazhuang 050031, China

a r t i c l e i n f o a b s t r a c t

Article history: Natural zeolite supported Cr-doped TiO2 photocatalysts were synthesized for the sake of improving the
Received 16 July 2011 recovery and photocatalytic efficiency of TiO2 . The materials were characterized by X-ray diffraction
Received in revised form (XRD), Brunauer–Emmett–Teller (BET) surface areas, Fourier transform infrared spectroscopy (FT-IR),
30 November 2011
scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), ultraviolet-visible diffuse
Accepted 26 December 2011
reflection spectroscopy (UV–vis DRS) and photoluminescence (PL). The photocatalytic activity was evalu-
Available online 30 December 2011
ated by the degradation of methyl orange in aqueous solution. The results show that the Cr concentration
and the calcination temperature play important role on the microstructure and photocatalytic activity of
Keywords:
Zeolite
the samples. The 10 mol% Cr-doped TiO2 /zeolite calcined at 400 ◦ C exhibits higher photocatalytic activity
Cr than that of the other samples.
TiO2 © 2011 Elsevier B.V. All rights reserved.
Photocatalyst

1. Introduction to supported nano TiO2 [2,3,11–14]. It shows that the natural zeo-
lites supported-TiO2 display higher photocatalytic efficiency at low
Titanium dioxide is the most important photocatalyst for the initial concentration in comparison with pure TiO2 .
degradation of organic pollutants in aqueous media [1]. However, It is believed that doping TiO2 with foreign ions is one of the most
practical applications of TiO2 are limited because of the problems promising strategies for sensitizing TiO2 to visible light and also
of low recycling efficiency of TiO2 fine particles, fast recombination for forming charge traps to keep electron–hole pairs separate [15].
rate of photogenerated electron–hole pair and a low quantum yield However, so far, there is little systematic study on the supported
in the photocatalytic reactions in aqueous solutions. ions-doped TiO2 especially natural zeolite supported ions doped-
Recently, researches have been devoted in searching for the TiO2 photocatalysts. In our previous work, we have reported the
suitable substrates to support nano TiO2 in order to improve the natural zeolite supported Fe3+ -TiO2 photocatalysts. It showed that
recovery efficiency of TiO2 . Of various substrates, zeolites have zeolite and iron ion concentration played an important role on the
been the most favorable to be used in supporting TiO2 , due to microstructure and photocatalytic activity of the samples. Zeolite
their unique structures, uniform pores and channels and excellent inhibited the growth of TiO2 crystallite sizes. The iron ions could
absorption ability [2,3]. Previous works have mainly emphasized on diffuse into TiO2 lattice to form the Fe O Ti bond and gave TiO2 the
synthetic zeolites [4–10] as the substrates of TiO2 . TiO2 supported capacity to absorb light at lower energy levels. The photocatalytic
on zeolite integrates the photocatalytic activity of TiO2 with the activity of the samples could be enhanced as appropriate dosages of
adsorption properties of zeolite together, which induce a synergis- Fe3+ were doped [14]. Recently, Cr-doped TiO2 has received much
tic effect, resulting in the enhancement of photocatalytic efficiency attention because Cr ion can excellently extend the visible light
[2]. However, the complex synthesizing process and high cost of absorption and improve the photocatalytic activity of TiO2 [16–18].
the synthetic zeolites considerably limit the practical application In this paper, natural zeolite supported Cr-doped TiO2 pho-
of the as-prepared photocatalysts. Natural zeolites are cheaper and tocatalysts were synthesized and the aim of this work was
more abundant as compared with synthetic zeolites [2,3]. In recent to investigate the effect of ions concentration and the calci-
years, natural zeolites have been used instead of synthetic zeolites nation temperature on the microstructure and property of the
as-prepared photocatalysts. Based on the above goals, X-ray diffrac-
tion (XRD), Brunauer–Emmett–Teller (BET), Fourier transform
infrared spectroscopy (FT-IR), scanning electron microscopy (SEM),
∗ Corresponding author. Tel.: +86 21 69582144. X-ray photoelectron spectroscopy (XPS), ultraviolet–visible diffuse
E-mail addresses: wangchengtj@163.com (C. Wang), shs@tongji.edu.cn (H. Shi). reflection spectroscopy (UV–vis DRS) and photoluminescence (PL)

0169-4332/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2011.12.108
C. Wang et al. / Applied Surface Science 258 (2012) 4328–4333 4329

were employed to investigate the microstructure of the photocata-


lysts. The photocatalytic activity of samples for the degradation of
methyl orange (MO) was also investigated.

2. Experimental

2.1. Photocatalyst preparation

Natural zeolite (clinoptilolite) powders were mixed with doubly


distilled water and heated up to 70 ◦ C. Subsequently, TiCl4 solu-
tion (the theoretical proportion of TiO2 in the photocatalyst was
30 wt.%) and CrCl3 solution (the nominal ratios of Cr3+ to TiO2 were
1.0 to 12.0 mol%, respectively) were added drop wise into the above
mortar and the pH of the dispersion was adjusted to about 2.0. The
system was vigorously stirred at 70 ◦ C for 4 h. After aging for 12 h,
the products were repeatedly washed with doubly distilled water
and then dried at 80 ◦ C for 2 h. The above samples were finally cal-
cined at 200–600 ◦ C for 2 h in a muffle furnace to obtain the natural
zeolite Cr-doped TiO2 photocatalysts.
Fig. 1. XRD patterns of zeolite and TiO2 /zeolite doped with different dosage of Cr
ions.
2.2. Characterization

The crystal structure of the samples were characterized by pow- the decoloration rate of the solution; A0 means the absorbency of
der XRD (D/max 2550VB3+/PC, Rigaku International Corporation, the original solution; At means the absorbency of the MO solution
Japan) under the following conditions: graphite monochromatic after treated for t hours.
copper radiation (Cu K␣ ,  = 0.15418 nm); 40 kV as accelerating
voltage; 40 mA as emission current; and the 2 range of 3◦ –70◦ at a 3. Results and discussion
scan rate of 4◦ /min. The nitrogen adsorption–desorption isotherms
were determined using a Micromeritcs Tristar 3000 nitrogen 3.1. XRD analysis
adsorption apparatus. To calculate the specific surface area the BET
equation was applied on the adsorption branch of the isotherms Fig. 1 shows the XRD patterns of zeolite and TiO2 /zeolite doped
whereas the pore size was calculated from the desorption branch with different dosage of chromium ions. The peaks at 2 = 25.4◦ ,
using the Barret–Joyner–Halenda (BJH) model. The pretreatment 37.5◦ , 48.2◦ , 54.0◦ , 55.0◦ , 62.7◦ ascribed to that of anatase TiO2 are
(degassing) of the samples took place under vacuum (10−2 Torr) observed. No correspondent peaks to chromium oxides phases are
at 150 ◦ C. FT-IR spectra were recorded by a spectrometer (Nexus, identified, even at the highest Cr concentration. It can be possibly
Thermo Nicolet Corporation, America) using KBr pellets at room inferred that: (i) the concentration of Cr-doping is so low that it
temperature in the region of 400–4000 cm−1 . The morphology of cannot be detected by XRD; (ii) Cr ions can be incorporated into
the photocatalyst was examined by field-emission scanning elec- the TiO2 lattice via displacing Ti4+ sites due to the close ionic radius
tron microscope (FESEM) (Quanta 200 FEG, FEI Company, America). of Ti4+ (0.68 Å) and Cr3+ (0.64 Å), forming a chromium–titanium
XPS were recorded under ultrahigh vacuum (<10−6 Pa) at a pass solid solution [18,19].
energy of 93.90 eV on a PerkinElmer PHI 5000C ESCA system Fig. 2 shows the XRD patterns of 10%Cr/TiO2 /zeolite calcined at
equipped with a dual X-ray source by using a Mg KR (1253.6 eV) 200–600 ◦ C. With the ascent in calcination temperature, the peak
anode and a hemispherical energy analyzer. All binding energies intensity of anatase phase increases significantly, indicating the
were calibrated with contaminant carbon (C 1s = 284.6 eV) as a improvement in crystallization of anatase phase. Simultaneously,
reference. UV–vis DRS of the samples were obtained for the dry- the width of the (1 0 1) peak becomes narrower, suggesting the
pressed disk samples using a UV–vis spectrophotometer with an
integrating sphere (UV-3010, HITA-CHI, Japan). The PL spectra of
the samples were recorded with RF-540 spectrofluorophotometer
(RF-540, Shimadzu Instruments, Japan), using the 280 nm excita-
tion line of a Xe lamp.

2.3. Photocatalytic activity tests

The photocatalytic activity of the samples was evaluated by the


degradation of MO in an aqueous solution. The photocatalytic reac-
tor system consisted of a 40 W UV lamp centered at 253.7 nm (the
integrated UV intensity was about 6.7 mW/cm2 , which is about
10 cm above the dishes) and a magnetic stirrer for obtaining dis-
persion. In a typical test for MO degradation, the photocatalysts
(0.2 g) were suspended in 100 mL MO solution (10 mg/L) and then
treated under UV light for different times. The decoloration rate
of the MO solution was monitored every regular time intervals by
measuring the maximum absorbance of MO (464 nm), using 722S
spectrometer (722S, Shanghai Precision & Scientific Instrument Co.,
Ltd., China). The decoloration rate of the MO solution was calcu-
lated. The calculational formula is: P = (A0 − At )/A0 × 100%. P means Fig. 2. XRD patterns of 10%Cr/TiO2 /zeolite calcined at 200–600 ◦ C.
4330 C. Wang et al. / Applied Surface Science 258 (2012) 4328–4333

Table 1 3.3. FT-IR analysis


Average pore diameter (APD) and specific surface area (SSA) of the raw zeolite,
10%Cr/TiO2 /zeolite calcined at 200–600 ◦ C.
Fig. 3 shows the FT-IR spectra of zeolite, TiO2 /zeolite and
Samples APD (nm) SSA (m2 /g) 10%Cr/TiO2 /zeolite in the range of 400–4000 cm−1 . The board peak
Raw zeolite 8.7 56 around 3437 cm−1 and 1637 cm−1 can be assigned to the O H
Cr/TiO2 /zeolite 200 ◦ C 3.1 205 stretching and bending vibration of adsorbed water, respectively.
Cr/TiO2 /zeolite 300 ◦ C 3.5 181 As shown in the spectrum of zeolite, the peak at 1049 cm−1 cor-
Cr/TiO2 /zeolite 400 ◦ C 4.1 187
responds to the Si O Si stretching vibration. The weak peaks
Cr/TiO2 /zeolite 500 ◦ C 4.9 152
Cr/TiO2 /zeolite 600 ◦ C 6.7 92 between 800–400 cm−1 correspond to Si O Si bending vibra-
tion. In contrast to the FT-IR spectra of zeolite, the spectra
of TiO2 /zeolite and Cr/TiO2 /zeolite obviously change. The board
peak appears between 800–400 cm−1 owing to the superpo-
growth of anatase crystallites. When the calcination temperature sition of Ti O and Si O bond. The Si O Si peaks belong to
increases to 600 ◦ C, few minor but discernible peaks correspond- zeolite, TiO2 /zeolite and 10%Cr/TiO2 /zeolite locate at 1049 cm−1 ,
ing to rutile phase are detected. In addition, it can be seen that the 1059 cm−1 and 1071 cm−1 , respectively. This may be due to the
(1 0 1) peak of anatase gradually shifts to the lower angle with the interrelationship of TiO2 –zeolite, Cr–TiO2 –zeolite, respectively.
increase in calcination temperature. This result indicates that Cr
ions have entered the crystal structure and consequently induce 3.4. SEM analysis
distortion to crystal lattice [20,21]. Simultaneously, the binding
force of Cr O Ti enhances and the crystallite defect increases with Fig. 4 shows the SEM images of zeolite and 10%Cr/TiO2 /zeolite.
the increase in calcination temperature [21]. From the SEM images it is evident that the originally smooth zeo-
lite matrix (Fig. 4a) is covered with TiO2 particles (Fig. 4b) and the
TiO2 particles are either as nanoparticles or clusters attaching to
3.2. BET analysis the zeolite surface.

Table 1 shows the average pore diameter and specific sur-


face area of the raw zeolite and 10%Cr/TiO2 /zeolite calcined at 3.5. XPS analysis
200–600 ◦ C. As can be seen in Table 1, all the Cr doped samples
have smaller average pore diameter and larger specific surface area Fig. 5 shows XPS whole scanning spectrum (a); Ti 2p (b); O 1s
than that of raw zeolite. This can be attributed to the supporting of (c); and Cr 2p3/2 (d) of TiO2 /zeolite and 10%Cr/TiO2 /zeolite, respec-
TiO2 nanoparticles on/in the surface/pores of zeolite. The nanopar- tively. From Fig. 5a, it can be seen that peaks of Ti, O, C, Si exist in
ticles block the pores of the zeolite and decrease the average pore XPS spectra of the above two samples. A weak peak of Cr appears
diameter of the samples. The high specific surface area of TiO2 in the XPS spectrum of ion-doped sample. The two strong peaks in
nanoparticles on/in the surface/pores of zeolite can increase the Fig. 5b correspond to Ti4+ 2p1/2 and Ti4+ 2p3/2 , respectively, which
specific surface area of the samples. The specific surface area of the is consistent with the values of Ti4+ in the TiO2 lattices [16,17]. It
Cr doped samples decreases with the increase in calcination tem- can be seen that the Ti 2p1/2 and Ti 2p3/2 binding energies of the
perature and this can be attributed to the growth of TiO2 particles. Cr/TiO2 /zeolite (466.0 eV and 460.2 eV) shift to higher energies as
However, with the ascent in calcination temperature, the average compared with that of TiO2 /zeolite (465.6 eV and 460.0 eV). It infers
pore diameter of the Cr doped samples increases. This is proba- that the doped Cr ions may diffuse into TiO2 lattices to form the
bly because the organics or impurity components contained in the Cr O Ti bond. In order to achieve the charge balance in the TiO2
pores of raw zeolite decompose under high calcination tempera- lattice, a trace of titanium ions may transform to higher valent by
ture, and thus the average pore diameter of the samples increases releasing the plethoric electrons, which leads to the XPS peaks shift
with the increase in calcination temperature. in the Cr/TiO2 /zeolite [17]. The O 1s spectra of Cr/TiO2 /zeolite and
TiO2 /zeolite are shown in Fig. 5c. It is noted that O 1s spectrum
of Cr/TiO2 /zeolite obviously shifts from 532.9 eV to 534.1 eV com-
pared with that of TiO2 /zeolite. The results can be ascribed to an
increase in the ionic state of the oxygen bond causing the binding
energies of all electronic states of oxygen to shift [22]. The peaks of
XPS spectra of Cr 2p3/2 (Fig. 5d) at 577.2 eV and 579.6 eV should be
assigned to Cr3+ and Cr6+ , respectively [18]. The integral area under
Cr3+ and Cr6+ peaks reveals that Cr6+ is the predominant species
(81.2 at.%).

3.6. UV–vis DRS analysis

The UV–vis DRS spectra of TiO2 /zeolite and 10%Cr/TiO2 /zeolite


and their correspondent energy band gaps (Eg ) are shown in Fig. 6. It
can be seen that the UV–vis absorbance for the Cr/TiO2 /zeolite sig-
nificantly shifts to longer wavelength (the red shift) in comparison
with that of TiO2 /zeolite. The band gap energies of TiO2 /zeolite and
Cr/TiO2 /zeolite obtained by the Kubelka–Munk theory are 2.84 eV
and 1.70 eV, respectively. The reduction in the band gap and visible
response of Cr/TiO2 /zeolite has been attributed to (i) the localized
states near the conduction or valence band of the modified TiO2 ;
Fig. 3. FT-IR spectra of zeolite, TiO2 /zeolite and 10%Cr/TiO2 /zeolite. (ii) the formation of color centers, which were associated with the
C. Wang et al. / Applied Surface Science 258 (2012) 4328–4333 4331

Fig. 4. SEM images of zeolite (a) and 10%Cr/TiO2 /zeolite (b).

oxygen vacancies in TiO2 or to the radicals in the titanium dioxide doping decreases the recombination rate of electron–hole pairs.
lattice associated with the doping ions [18]. Generally speaking, the smaller the particle size, the larger the
oxygen vacancy and defects content, the higher the probability of
3.7. PL analysis exciton occurrence, the stronger the PL signal [26]. As analyzed
in the above section, Cr doping induces distortion to the crystal
Fig. 7 shows the PL spectra of TiO2 /zeolite and lattice and increases the defect of TiO2 , which may show stronger
10%Cr/TiO2 /zeolite. It can be found that all of the samples can PL signal. However, the Cr ions exist in the TiO2 crystal lattice as
exhibit obvious PL signal centered at approximately 467 nm. These 3+ and 6+, and Cr6+ can capture the photogenerated electrons and
PL signals result from the surface oxygen vacancies and defects of then maintain the charge balance. The capture effect of defect on
TiO2 [23]. A lower PL intensity may indicate a lower recombina- the photogenerated electrons is lower than that of the Cr6+ [27],
tion rate of electron–hole pairs and higher separation efficiency and this may result in the lower PL signal and recombination rate
[24,25]. As shown in Fig. 7, the PL intensity of Cr/TiO2 /zeolite is of electron–hole pairs, and thus higher photocatalytic activity of
lower than that of undoped TiO2 /zeolite which means that Cr Cr doped sample.

Fig. 5. XPS whole scanning spectrum (a); Ti 2p (b); O 1s (c); and Cr 2p3/2 (d) of TiO2 /zeolite and 10%Cr/TiO2 /zeolite.
4332 C. Wang et al. / Applied Surface Science 258 (2012) 4328–4333

Fig. 6. UV–vis DRS spectra and their corresponding calculated Eg for the TiO2 /zeolite
and 10%Cr/TiO2 /zeolite.

3.8. Photocatalytic activity

Fig. 8a shows the degradation curves of MO solution catalyzed


by the photocatalysts doped with different dosage of Cr ions cal-
cined at 400 ◦ C. It can be observed that the decoloration rates
of MO solution catalyzed by the Cr-doped samples are obviously
higher than that of the undoped sample. The decoloration rate
of MO solution catalyzed by 10%Cr/TiO2 /zeolite reaches to 41.73%
after illumination for 5 h which is 17.9% higher than that of the
undoped TiO2 /zeolite. It indicates that the photocatalytic activity
of the photocatalysts can be effectively improved after doping with Fig. 8. Decoloration rate of MO solution catalyzed by photocatalysts doped with
chromium ions. This phenomenon can be attributed to the visible different dosage of Cr ions (a); 10%Cr/TiO2 /zeolite calcined at 200–600 ◦ C (b).
response and lower recombination rate of electron–hole pairs of Cr
doped sample.
Fig. 8b shows the degradation curves of MO solution catalyzed After calcined for 200 ◦ C, the TiO2 /zeolite photocatalysts have
by the 10%Cr/TiO2 /zeolite calcined at 200–600 ◦ C. It can be observed high specific surface area and excellent adsorption ability, and thus
that the decoloration rate of MO solution catalyzed by the sample exhibit excellent decoloration effect on the MO solution due to its
calcined at 200 ◦ C reaches to 40.68% after illumination for 5 h. As remarkable ability in gathering the organic molecules onto/near
the calcination temperature increases to 300 ◦ C, the decoloration the surface of TiO2 particles. The specific surface area decreases due
rate rapidly reduces to 9.58%. However, as the calcination temper- to the increase in calcination temperature. This result may further
ature further increases to 400 ◦ C, the decoloration rate drastically lead to the decrease of the adsorption and decoloration abilities
increases to 41.73%. The decoloration rates reduce to 35.83% and of the photocatalysts. Simultaneously, the photocatalytic activity
6.69% when the calcination temperatures are 500 ◦ C and 600 ◦ C, of TiO2 increases as the calcination temperature increases from
respectively. 200 ◦ C to 400 ◦ C. However, when the calcination temperature rises
up to 500 ◦ C, the photocatalytic activity of TiO2 decreases again. In
conclusion, the absorption and photocatalytic synergetic effect of
Cr-doped TiO2 /zeolite can lead to the above experiential results.

4. Conclusions

Natural zeolite supported Cr-doped TiO2 photocatalysts were


successfully prepared in this study. Cr ions can be incorporated
into the TiO2 lattice via displacing Ti4+ sites due to the close ionic
radius of Ti4+ and Cr3+ . The binding force of Cr O Ti enhances with
the increase in calcination temperature. Cr dopant is present as
Cr3+ (19.8 at.%) and Cr6+ (81.2 at.%) species. The band gap energy
of 10 mol%Cr/TiO2 /zeolite significantly reduces from 2.84 eV to
1.70 eV in comparison with that of the undoped TiO2 /zeolite. The
photocatalytic activities of the doped samples strongly depend
on the concentration of chromium ions and the calcination tem-
perature. The decoloration rate of MO solution catalyzed by
10%Cr/TiO2 /zeolite calcined at 400 ◦ C reaches to 41.73% after illu-
mination for 5 h which is 17.9% higher than that of the undoped
Fig. 7. PL spectra of TiO2 /zeolite and 10%Cr/TiO2 /zeolite. TiO2 /zeolite.
C. Wang et al. / Applied Surface Science 258 (2012) 4328–4333 4333

Acknowledgments [12] M. Nikazar, K. Gholivand, K. Mahanpoor, Photocatalytic degradation of azo


dye Acid Red 114 in water with TiO2 supported on clinoptilolite as a catalyst,
Desalination 219 (2008) 293–300.
This work was financially supported by the Hebei Natural Sci- [13] S. Ko, P.D. Fleming, M. Joyce, P. Ari-Gur, High performance nano-titania photo-
ence Foundation, China (grant no. E2008000537), Hebei Foundation catalytic paper composite. Part II: preparation and characterization of natural
for Development of Science and Technology, China (grant no. zeolite-based nano-titania composite sheets and study of their photocatalytic
activity, Mater. Sci. Eng. B 164 (2009) 135–139.
07215156) and Open Research Foundation of Key Laboratory of [14] C. Wang, H.S. Shi, Y. Li, Synthesis and characteristics of natural zeolite supported
Advanced Civil Engineering Materials (Tongji University), Ministry Fe3+ -TiO2 photocatalysts, Appl. Surf. Sci. 257 (2011) 6873–6877.
of Education, China (grant no. 2010412). [15] A. Fujishima, X. Zhang, D.A Tryk, TiO2 photocatalysis and related surface phe-
nomena, Surf. Sci. Rep. 63 (2008) 515–582.
[16] X.X. Fan, X.Y. Chen, S.P. Zhu, Z.S. Li, T. Yu, J.H. Ye, Z.G. Zou, The structural, physical
References and photocatalytic properties of the mesoporous Cr-doped TiO2 , J. Mol. Catal.
A: Chem. 284 (2008) 155–160.
[1] M. Alvaro, E. Carbonell, V. Fornés, H. García, Enhanced photocatalytic activity [17] J.F. Zhu, Z.G. Deng, F. Chen, J.L. Zhang, H.J. Chen, M. Anpo, J.Z. Huang, L.Z.
of zeolite encapsulated TiO2 clusters by complexation with organic additives Zhang, Hydrothermal doping method for preparation of Cr3+ -TiO2 photocat-
and N-doping, ChemPhysChem 7 (2006) 200–205. alysts with concentration gradient distribution of Cr3+ , Appl. Catal. B 62 (2006)
[2] M.L. Huang, C.F. Xu, Z.B. Wu, Y.F. Huang, J.M. Lin, J.H. Wu, Photocatalytic dis- 329–335.
colorization of methyl orange solution by Pt modified TiO2 loaded on natural [18] R. López, R. Gómez, S. Oros-Ruiz, Photophysical and photocatalytic properties
zeolite, Dyes Pigm. 77 (2008) 327–334. of TiO2 -Cr sol–gel prepared semiconductors, Catal. Today 166 (2011) 159–165.
[3] F.F. Li, S.M. Sun, Y.S. Jiang, M.S. Xia, M.M. Sun, B. Xue, Photodegradation of an [19] J.G. Yu, Q.J. Xiang, M.H. Zhou, Preparation, characterization and visible-light-
azo dye using immobilized nanoparticles of TiO2 supported by natural porous driven photocatalytic activity of Fe-doped titania nanorods and first-principles
mineral, J. Hazard. Mater. 152 (2008) 1037–1044. study for electronic structures, Appl. Catal. B 90 (2009) 595–602.
[4] E. Amereh, S. Afshar, Photodegradation of acetophenone and toluene in water [20] C.J. Yang, C.J. Li, C.X. Li, Y.Y. Wang, X.C. Huang, Effect of Cu2+ doping on pho-
by nano-TiO2 powder supported on NaX zeolite, Mater. Chem. Phys. 120 (2010) tocatalytic performance of liquid flame sprayed TiO2 coatings, J. Therm. Spray
356–360. Technol. 15 (2006) 582–586.
[5] H. Yahiro, T. Miyamoto, N. Watanabe, H. Yamaura, Photocatalytic partial oxi- [21] B. Zhang, C.Z. Sun, L. Qi, L. Dong, Effect of Zr4+ doping on structures and per-
dation of ␣-methylstyrene over TiO2 supported on zeolites, Catal. Today 120 formance of nano-structured Au/TiO2 catalysts, Chin. J. Inorg. Chem. 27 (2011)
(2007) 158–162. 1798–1804.
[6] M. Lafjah, F. Djafri, A. Bengueddach, N. Kellera, V. Keller, Beta zeolite supported [22] W.C. Hung, Y.C. Chen, H. Chu, T.K. Tseng, Synthesis and characterization of TiO2
sol–gel TiO2 materials for gas phase photocatalytic applications, J. Hazard. and Fe/TiO2 nanoparticles and their performance for photocatalytic degrada-
Mater. 186 (2011) 1218–1225. tion of 1,2-dichloroethane, Appl. Surf. Sci. 255 (2008) 2205–2213.
[7] M. Khatamian, S. Hashemian, S. Sabaee, Preparation and photo-catalytic activ- [23] B.Q.Wang, L.Q. Jing, Y.C. Qu, S.D. Li, B.J. Jiang, L.B. Yang, B.F. Xin, H.G. Fu, Enhance-
ity of nano-TiO2 -ZSM-5 composite, Mater. Sci. Semicond. Process. 13 (2010) ment of the photocatalytic activity of TiO2 nanoparticles by surface-capping
156–161. DBS groups, Appl. Surf. Sci. 252 (2006) 2817–2825.
[8] W.J. Zhang, K.L. Wang, Y. Yu, H.B. He, TiO2 /HZSM-5 nano-composite photocata- [24] H.S. Cai, G.G. Lu, W.Y. Lü, X.X. Li, L. Yu, D.G. Li, Effect of Ho-doping on photocat-
lyst: HCl treatment of NaZSM-5 promotes photocatalytic degradation of methyl alytic activity of nanosized TiO2 catalyst, J. Rare Earths 26 (2008) 71–75.
orange, Chem. Eng. J. (Amst., Netherland) 163 (2010) 62–67. [25] J.C. Yu, J.G. Yu, W.K. Ho, Z.T. Jiang, L.Z. Zhang, Effects of F-doping on the photo-
[9] M. Signoretto, E. Ghedini, V. Trevisan, C.L. Bianchi, M. Ongaro, G. Cruciani, TiO2 - catalytic activity and microstructures of nanocrystalline TiO2 powders, Chem.
MCM-41 for the photocatalytic abatement of NOx in gas phase, Appl. Catal. B Mater. 14 (2002) 3808–3816.
95 (2010) 130–136. [26] L.Q. Jing, X.J. Sun, B.F. Xin, B.Q. Wang, W.M. Cai, H.G. Fu, The preparation and
[10] M. Takeuchi, J. Deguchi, M. Hidaka, S. Sakai, K. Woo, P.P. Choi, J.K. Park, M. Anpo, characterization of La doped TiO2 nanoparticles and their photocatalytic activ-
Enhancement of the photocatalytic reactivity of TiO2 nano-particles by a simple ity, J. Solid State Chem. 177 (2004) 3375–3382.
mechanical blending with hydrophobic mordenite (MOR) zeolite, Appl. Catal. [27] B.F. Xin, L.Q. Jing, G.H. Fu, Z.H. Sun, Z.Y. Ren, B.Q. Wang, W.M. Cai, Preparation
B 89 (2009) 406–410. and characterization of Cu doped TiO2 nanoparticles and their photocatalytic
[11] F.F. Li, Y.S. Jiang, L.X. Yu, Z.W. Yang, T.Y. Hou, S.S. Sun, Surface effect of natural activity, Chem. J. Chin. Univ. 25 (2004) 1076–1080.
zeolite (clinoptilolite) on the photocatalytic activity of TiO2 , Appl. Surf. Sci. 252
(2005) 1410–1416.

You might also like