You are on page 1of 6

JOURNAL OF RARE EARTHS, Vol. 29, No. 9, Sep. 2011, P.

866

Photocatalytic degradation of methyl orange by polyoxometalates supported on


yttrium-doped TiO2
WANG Yajun (), LU Kecheng (), FENG Changgen ()
(State Key Laboratory of Explosion Science and Technology, Beijing Institute of Technology, Beijing 100081, China)
Received 9 March 2011; revised 7 April 2011

Abstract: A series of novel photocatalysts, H3PW12O40-Y-TiO2 nanocomposites with different H3PW12O40 loading levels (10%40%) were
prepared by impregnation method. And the Y-TiO2 support, doped with yttrium, was synthesized via sol-gel technique. The prepared catalysts
were characterized by Fourier transform infrared spectroscopy (FT-IR), powder X-ray diffraction (XRD), nitrogen adsorption-desorption
analysis and scanning electron microscopy (SEM). The processes allowed obtaining Keggin structure and crystallized anatase with large BET
surface area as well as uniform distribution. The effects of H3PW12O40 loadings, catalyst dose, initial pH and concentration of dye solution on
the degradation kinetics of methyl orange under UV light (365 nm) were discussed. Kinetics studies showed that the photocatalytic degradation of methyl orange fitted the apparent first-order reaction. Methyl orange was totally degraded in 21 min under optimum conditions: 20%
loading, 0.03 g dose and pH 1.0. The catalyst was stable and easily to be separated from reaction system for recovery.
Keywords: polyoxometalates; titanium dioxide; rare earths; photocatalytic degradation; kinetics; azo dyes

Azo dyes waste water, released into water bodies without


degradation is toxic to the ecosystem and also has significant
influence on human health[1,2]. Therefore, such pollutants
have to be treated prior to discharging into the environment.
They can be eliminated efficiently by photocatalytic degradation, an advanced oxidation process. As green and effective photocatalysts, polyoxometalates (POMs) share many
features with semiconductors and can be considered as the
analogues of the latter. The photocatalysis of POMs, which
originated from the photoexcitation of the oxygen-to-metal
charge-transfer bands of POMs, is based on the electron-hole
pair separation followed by reductive and oxidative reactions
with surrounding molecules[35].
Supporting POMs on TiO2 has proved to be an efficient
way to promote the photocatalytic performance of POMs,
attributing to the synergistic effect between POMs and
TiO2[69]. As effective electron capture agents, POMs receive
electrons generated by light irradiation of TiO2, and then extend the recombination time of electrons and holes[10].
The photocatalytic activity of TiO2 can be enhanced by
doping with rare earths[11,12]. The present research is to develop novel photocatalysts for treating the azo dye-containing contamination. H3PW12O40 (HPW) was selected as precursor and TiO2 doped with yttrium as support to prepare
photocatalysts H3PW12O40-Y-TiO2 (HPW-Y-TiO2) with different loadings, by using impregnation method and sol-gel
technique. The obtained catalysts were characterized by
various means. The photocatalytic degradation kinetics of
azo dye, methyl orange (MO), under UV light irradiation

(365 nm) was employed to investigate the photocatalytic


activity of the composites.

1 Experimental
1.1 Materials
All chemicals were purchased from Sinopharm Chemical
Reagent Co., China and employed without any further purification. Phosphotungstic acid hydrate (H3PW12O40nH2O),
yttrium nitrate hexahydrate (Y(NO3)36H2O), absolute ethyl
alcohol (EtOH), hydrochloric acid (HCl), perchloric acid
(HClO4) and methyl orange (MO) are analytical grade reagents. And tetrabutyl titanate (Ti(OBu)4) is chemical grade.
1.2 Catalysts preparation
To avoid decomposing of Keggin structure during calcination of the catalysts, HPW-Y-TiO2 composites were prepared by impregnation method while the supports Y-TiO2
were firstly synthesized via sol-gel technique.
The support was synthesized as below. A mixture of
Ti(OBu)4 (10 ml) and EtOH (30 ml) was stirred under room
temperature. HCl was added into the mixture to obtain pH
1.5, and the solution was marked A. 0.08 g Y(NO3)36H2O,
the rare earth precursor, was dissolved into the solution containing H2O (2 ml) and EtOH (4 ml), which was marked B.
Then, B was added dropwise into A during 10 min approximately. The resulting acidic mixture was stirred constantly
for about 3 h until sol was obtained. The sol was maintained

Foundation item: Project supported by Institution of Chemical Materials, China Academy of Engineering Physics
Corresponding author: WANG Yajun (E-mail: yajunwang@bit.edu.cn; Tel.: +86-10-68912941)
DOI: 10.1016/S1002-0721(10)60557-1

WANG Yajun et al., Photocatalytic degradation of methyl orange by polyoxometalates supported on yttrium-doped TiO2

for 24 h till gel formation. The gel was dried at 373 K for 4 h
and then calcinated at 823 K for 3 h with a heating rate of
5 K/min. The rare earth doping amount (calculated by Y2O3)
in the support is 1.0%. Pure TiO2 powder was also prepared
by the same process without doping.
The impregnation process is as following: 1 g Y-TiO2 support was dispersed in water solution containing a given
amount of HPW. The suspension was stirred continuously
under room conditions. After 24 h, water was removed by
evaporation at 373 K. The final product was obtained by
drying at 373 K for 2 h. The loading levels of HPW in four
prepared composites by mass are 10%, 20%, 30% and 40%
(by theoretical calculation), respectively.
1.3 Characterization
The IR spectra of the composites were investigated in the
wavenumber range of 4000400 cm1 using Thermo Nicolet
6700 Fourier transform infrared (FT-IR) spectrometer.
Crystalline phases of the prepared catalysts were characterized by powder X-ray diffraction (XRD) on a PANalytical
X' Pert PRO MPD diffractometer using Cu K radiation
which was operated at 40 kV and 40 mA. The textural properties were determined from adsorption-desorption isotherms
of nitrogen by a Beishide 3H-2000PS2 instrument using
multipoint BET and BJH methods. Surface morphologies
were observed by scanning electron microscopy (SEM) (Hitachi S-4800N) operating at 5.0 kV.

ObW), as(WOcW), respectively[10]. From Fig. 1(1) it can


be seen that the spectroscope of Y-TiO2 shows an intense,
broad, indistinct region between 1100 and 400 cm1. As a
result, some characteristic peaks of Keggin unit in HPW-YTiO2 are overlapped in this area (see Fig. 1(25)). However,
some vibration peaks of Keggin unit still can be seen clearly,
which indicates that the Keggin structure has not been destroyed.
2.1.2 XRD analysis Fig. 2 shows the XRD testing results.
It can be observed that the XRD patterns of Y-doped TiO2
and pure TiO2 are similar, both crystallized in the anatase
structure with characteristic diffraction peaks of 2 values
located at 25.3(101), 37.8(004), 48.1(200), 54.0(105),
62.7(211), respectively[14]. However, no characteristic peak
of Y oxide is found, which implies that Y oxide content is
very small and highly dispersed[15].
From Fig. 2(36) we can see that all the samples with different loading levels have anatase phase. For 40% loading, it
is easy to identify the main diffraction peaks of HPW at
2=10.4 and 25.4[16]. However, for 10%30% loading,
only anatase phase is present and no separate polyoxotungstate-related phase is observed. This may be the reason that
HPW is either in the octahedral interstitial sites or the substitutional positions of Y-TiO2[7].
The particle diameters were calculated from Scherrer for

1.4 Photocatalytic procedures


The photocatalysis experiments were performed on an
open photoreactor. The light source is provided by a PLSSXE300UV Xe lamp (300 W) with emission of 365 nm,
which is positioned above the photoreactor.
The initial concentration (C0) of MO was fixed at 10 mg/L,
except for the tests investigating the effect of C0. A given
amount of prepared catalyst was suspended into a fresh MO
solution (50 ml). Thirty-minute adsorption-desorption time
in dark condition was allowed prior to photoreaction. The
beaker was transferred to the photoreactor after the intensity
of UV light was stable. The suspension was vigorously
stirred during the whole process. At given interval of illumination, a sample of suspension (ca. 3 ml) was taken out and
filtered with microporous membrane (0.45 Pm). The absorbance of residual MO solution was analyzed by an APL
752 UV-vis spectrometer at 464 nm when pH4.0 and 510
nm when pH3.0. The decrease of absorbance was used to
indicate the degradation of MO[13].

Fig. 1 FT-IR spectra of the prepared samples


(1) Y-TiO2; (2) HPW-Y-TiO2 (10%40%)

2 Results and discussion


2.1 Characterization of the catalysts
2.1.1 FT-IR analysis Fig. 1 shows the FT-IR spectra of
Y-TiO2 and HPW-Y-TiO2 with different loadings. The
characteristic absorption peaks of Keggin unit at 1080, 982,
888, 797 cm1 attribute to as(POa), as(W=Od), as(W

867

Fig. 2 XRD patterns of prepared samples


(1) TiO2; (2) Y-TiO2; (3) HPW-Y-TiO2 (10%40%)

868

JOURNAL OF RARE EARTHS, Vol. 29, No. 9, Sep. 2011

mula and listed in Table 1. The data show that all prepared
powders are nanoparticles. The crystallite size decreases
from 25.0 to 11.5 nm, indicating that Y-doping could restrain the increase in grain size during calcination.
2.1.3 Nitrogen adsorption-desorption analysis Table 2
gives the results of nitrogen adsorption-desorption analysis
for the samples. The BET surface area of Y-TiO2 is 69.0
m2/g, which is 1.8 times as that of undoped TiO2. The increase of surface area could be explained by small crystallite
size and high dispersion of rare earth. The data also illustrate
that the BET surface area of POMs is markedly increased by
Table 1 Crystallite size of prepared materials
Materials

Crystallite size/nm

TiO2

25.0

Y-TiO2

11.5

HPW-Y-TiO2 (10%)

9.9

HPW-Y-TiO2 (20%)

9.1

HPW-Y-TiO2 (30%)

9.0

HPW-Y-TiO2 (40%)

14.5

2.2 Photocatalytic testing

Table 2 Microstructure of the prepared catalysts


SBET/

Average pore

Pore volume/

(m2g)

size/nm

(cm3g)

TiO2

37.6

10.1

0.128

Y-TiO2

69.0

10.0

0.221

Materials

immobilization, comparing with the initial POMs (<10


m2/g)[17]. As the area of the Keggin anion is ca. 1.13 nm2 [18]
and BET surface area of the support is 69.0 m2/g, a loading
of about 20% would theoretically saturate the Y-TiO2 support surface. BET surface area and pore volume decrease at
high loading, and this may be because some pores of the
support are blocked up by excess precursor. The supported
POMs obtained are typical mesoporous materials with average
pore diameter of 810 nm.
2.1.4 SEM observation The SEM images of the final
products (Fig. 3) reveal that the supported HPW forms relatively uniform nanometer particles of diameter less than 20 nm,
which is in accordance with the XRD results. The particles
obtained are regular spherical.

HPW-Y-TiO2 (10%)

67.9

9.2

0.261

HPW-Y-TiO2 (20%)

66.0

8.7

0.155

HPW-Y-TiO2 (30%)

56.0

8.2

0.123

HPW-Y-TiO2 (40%)

47.3

8.0

0.120

No detectable degradation occurred on MO solution under


UV light irradiation without catalysts and only about 5%
concentration decreased under the action of catalysts in dark,
implying that the degradation of MO originates from the
combination of the UV light and the photocatalyst. The
photocatalytic tests show that the degradation of MO follows
Langmuir-Hinshelwood (LH) apparent first-order kinetics.
2.2.1 The effect of different loadings Catalysts used in
the evaluation were 0.02 g. MO was used without pH adjustment and the pH was 5.3. The effect of different loadings
on the kinetics is shown in Fig. 4. The results indicate that
the best loading level is 20%, which is also the saturation
value of loading level. At loading levels higher than 20%,
excess HPW is easy to drop from the composite during the

Fig. 3 SEM images of HPW-Y-TiO2 composites with different loadings


(a) 10%; (b) 20%; (c) 30%; (d) 40%

WANG Yajun et al., Photocatalytic degradation of methyl orange by polyoxometalates supported on yttrium-doped TiO2

869

Fig. 4 Effect of HPW loadings on kinetics

Fig. 6 Effect of catalyst dose on kinetics

photocatalytic process, and less reactant is adsorbed, resulting in the decline of apparent first-order rate constant (kobs).
This is in agreement with the conclusions obtained from the
nitrogen adsorption-desorption analysis results.
Fig. 5 and Table 3 give the results of degradation kinetics
in the presence of different catalysts. The kobs for Y-TiO2 is
0.006 min1, which is 1.50 times as that of undoped TiO2,
indicating that the doping of Y permits an improvement of
photocatalytic activity of TiO2. This could be explained by
the fact that Y-TiO2 has smaller crystallite size and good
dispersion, which contributes to the quantum size effect[19].
In addition, Y-doping could extend the recombination time
of electron-hole pairs, resulting in better photoactivity. It is
obvious that the order of the photocatalytic performance is:
HPW-Y-TiO2(20%)>Y-TiO2>TiO2. The kobs of HPW-YTiO2 (20%) is 3.50 times the value of Y-TiO2.
2.2.2 The effect of photocatalyst dose The pH of MO
was not adjusted in the test (pH 5.3). Fig. 6 and Table 4
show the effects of HPW-Y-TiO2 (20%) dose on kinetics.
The degradation rate is very slow at 0.01 g dose, implying

Table 4 Apparent first-order kinetics equations and relative


parameters for photocatalytic degradation of methyl
orange with different HPW-Y-TiO2 (20%) doses (pH 5.3)
Dose/g

Kinetics equation

kobs/min1

0.01

ln(C0/C)=0.001t+0.005

0.001

693.1

0.991

0.02

ln(C0/C)=0.021t0.204

0.021

33.0

0.982

0.03

ln(C0/C)=0.027t0.140

0.027

25.7

0.991

0.04

ln(C0/C)=0.021t0.128

0.021

33.0

0.989

t1/2/min

that activated species is not enough at this condition. It is


clear that the optimum dose is 0.03 g. The degradation rate
decreases when dose is up to 0.04 g. This is because the turbidity of the suspension increases with high catalyst dose,
resulting in decrease in UV light penetration and photoactivated species.
2.2.3 The effect of initial pH In the test, the loading was
20% and the dose was 0.03 g. The effect of initial pH of the
MO solution, which was adjusted with HClO4, was also investigated. MO was not found degraded in the presence of
HClO4 under UV light without catalyst, suggesting that
HClO4 has no photocatalytic activity to MO. The dependence of degradation kinetics on pH is shown in Fig. 7 and
Table 5. The results demonstrate that pH has significant impact on the photocatalytic degradation rate. The reaction rate
increases with the decrease of pH. This could be explained
by the fact that HPW is a super acidic catalyst which is stable at ca. pH 1. So low pH is benefitial to the catalytic activity. Another explanation is that MO is converted into quinoid

Fig. 5 Effect of different catalysts on kinetics


Table 3 Apparent first-order kinetics equations and relative
parameters for photocatalytic degradation of methyl
orange with different catalysts (pH 5.3)
Materials

Kinetics equation

kobs/min1

t1/2/min

R2

TiO2

ln(C0/C)=0.004t0.037

0.004

173.3

0.992

Y-TiO2

ln(C0/C)=0.006t0.012

0.006

115.5

0.999

HPW-Y-TiO2 (20%)

ln(C0/C)=0.021t0.204

0.021

33.0

0.982

R2

Fig. 7 Effect of initial pH on kinetics

870

JOURNAL OF RARE EARTHS, Vol. 29, No. 9, Sep. 2011

Table 5 Apparent first-order kinetics equations and relative


parameters for photocatalytic degradation of methyl
orange with different initial pH
pH

Kinetics equation

kobs/min1

t1/2/min

R2

1.0

ln(C0/C)=0.182t0.122

0.182

3.8

0.981

2.0

ln(C0/C)=0.103t+0.004

0.103

6.7

0.999

3.0

ln(C0/C)=0.073t0.111

0.073

9.5

0.988

4.0

ln(C0/C)=0.036t+0.020

0.036

19.3

0.997

structure at low pH, while quinoid structure is easy to be destroyed.


From above discussion, the optimum conditions for photocatalytic degradation of MO solution (50 ml, C0=10 mg/L)
are: 20% loading, 0.03 g dose and pH 1.0. Fig. 8 shows the
photocatalytic behaviors of supported and starting HPW,
Y-TiO2 under optimum conditions. It can be seen that MO is
almost degraded in 21 min while just about 70% disappeared
in 120 min in the presence of HPW-Y-TiO2 and HPW, respectively. As listed in Table 6, kobs for the supported HPW
is 0.182 min1, corresponding to more than 15 times the
value of primary HPW, and 2.04 times that of Y-TiO2 under
the same conditions. The photocatalytic performance of
POMs is greatly improved by immobilization, which is attributed to the synergistic effect between Keggin-type and
anatase support as well as the increased surface area after
immobilization.
2.2.4 The effect of initial concentration Considering the
effect of initial concentration, the test conditions were: 20%
loading, 0.03 g dose and pH 1.0. When C0 is changed in the
range of 530 mg/L, the degradation is apparent first-order
reaction. It can be seen from Fig. 9 and Table 7 that the
higher the C0 is, the lower the degradation rate is. This

Fig. 9 Effect of initial concentration on kinetics


Table 7 Apparent first-order kinetics equations and relative
parameters for photocatalytic degradation of methyl
orange with different initial concentration
Kinetics equation

kobs/min1

t1/2/min

R2

ln(C0/C)=0.241t0.101

0.241

2.9

0.990

10

ln(C0/C)=0.182t0.122

0.182

3.8

0.981

20

ln(C0/C)=0.093t0.133

0.093

7.5

0.990

30

ln(C0/C)=0.036t0.232

0.036

19.3

0.989

C0/(mg/L)

may be because more MO are absorbed on the catalyst surface at high C0, resulting in decline of the free radicals photoexcited by catalyst. In addition, UV light is hard to spread
and less photons involve in the photocatalytic process.
2.3 Stability of the catalyst
The stability of HPW-Y-TiO2 (20%) catalyst was evaluated under optimum conditions. The degradation conversion
of MO keeps at ca. 96% after the catalyst was reused eight
times. The composite is easy to be separated from solution
after reaction by natural sedimentation, which is of great
importance for potential practical applications.

3 Conclusions

Fig. 8 Photocatalytic activity of supported HPW, starting HPW, and


Y-TiO2 under optimum conditions
Table 6 Apparent first-order kinetics equations and relative
parameters for photocatalytic degradation of methyl
orange with HPW, supported HPW and Y-TiO2 under optimum conditions
Materials

Kinetics equation

kobs/min1

t1/2/min

R2

HPW

ln(C0/C)=0.012t0.174

0.012

57.8

0.989

HPW-Y-TiO2 (20%)

ln(C0/C)=0.182t0.122

0.182

3.8

0.981

Y-TiO2

ln(C0/C)=0.089t0.190

0.089

7.8

0.989

Novel solid Keggin-type POMs, HPW-Y-TiO2 with different HPW loading levels were prepared by impregnation
method. Both Keggin structure and anatase existed in the
prepared materials. The optimum conditions for photocatalytic degradation of MO (50 ml, 10 mg/L) were: 20% loading,
0.03 g dose and pH 1.0. The disappearance of MO followed
Langmuir-Hinshelwood apparent first-order kinetics. The
photocatalytic activity of POMs was markedly enhanced,
owing to the synergistic effect between POMs and anatase
Y-TiO2 as well as the increased BET surface area. The
photocatalyst kept high activity after eight times recycle.

References:
[1] Feng C G, Zhuo X X, Liu X. Study on photodegradation of azo
dye by polyoxometalates/polyvinyl alcohol. Journal of Rare
Earths, 2009, 27(5): 717.
[2] Bubacz K, Choina J, Dolat D, Morawski A W. Methylene blue

WANG Yajun et al., Photocatalytic degradation of methyl orange by polyoxometalates supported on yttrium-doped TiO2
and phenol photocatalytic degradation on nanoparticles of
anatase TiO2. Polish Journal of Environmental Studies, 2010,
19(4): 685.
[3] Yamase T. Photo- and electrochromism of polyoxometalates
and related materials. Chemical Reviews (Washington, D C),
1998, 98(1): 307.
[4] Yamase T. Photoredox chemistry of polyoxometalates as a
photocatalyst. Catalysis Surveys from Asia, 2003, 7(4): 203.
[5] Guo Y H, Wang Y H, Hu C W, Wang Y H, Wang E B, Zhou
Y C, Feng S H. Microporous polyoxometalates POMs/SiO2:
synthesis and photocatalytic degradation of aqueous organocholorine pesticides. Chemistry of Materials, 2000, 12(11):
3501.
[6] Yang Y, Guo Y H, Hu C W, Jiang C J, Wang E B. Synergistic
effect of Keggin-type [Xn+W11O39](12n) and TiO2 in macroporous hybrid materials [Xn+W11O39](12n) -TiO2 for the photocatalytic degradation of textile dyes. Journal of Materials
Chemistry, 2003, 13(7): 1686.
[7] Yang Y, Wu Q Y, Guo Y H, Hu C W, Wang E B. Efficient
degradation of dye pollutants on nanoporous polyoxotungstate-anatase composite under visible-light irradiation. Journal
of Molecular Catalysis A: Chemical, 2005, 225(2): 203.
[8] Jin H X, Wu Q Y, Pang W Q. Photocatalytic degradation of
textile dye X-3B using polyoxometalate-TiO2 hybrid materials.
Journal of Hazardous Materials, 2007, 141(1): 123.
[9] Li K X, Guo Y N, Ma F Y, Li H C, Chen L, Guo Y H. Design
of ordered mesoporous H3PW12O40-titania materials and their
photocatalytic activity to dye methyl orange degradation. Catalysis Communications, 2010, 11(9): 839.
[10] Lee J, Dong X L, Dong X W. Ultrasonic synthesis and photo-

871

catalytic characterization of H3PW12O40/TiO2 (anatase). Ultrasonics Sonochemistry, 2010, 17(4): 649.


[11] Ou X Q, Meng J P, Wang Q M, Yu J M. Enhanced photoactivity of layered nanocomposite materials containing rare earths,
titanium dioxide and clay. Journal of Rare Earths, 2006, 24(1):
251.
[12] Fan C M, Xue P, Sun Y P. Preparation of nano-TiO2 doped
with cerium and its photocatalytic activity. Journal of Rare
Earths, 2006, 24(3): 309.
[13] Li F B, Gu G B, Huang G F, Gu Y L, Wan H F. TiO2-assisted
photo-catalysis degradation process of dye chemicals. Journal
of Environmental Sciences, 2001, 13(1): 64.
[14] Zhu H Y, Orthman J A, Li J Y, Zhao J C, Churchman G J,
Vansant E F. Novel composites of TiO2 (anatase) and silicate
nanoparticles. Chemistry of Materials, 2002, 14(12): 5037.
[15] Shi Z L, Liu F M, Yao S H. Preparation and photocatalytic activity of B, Y co-doped nanosized TiO2 catalyst. Journal of
Rare Earths, 2010, 28(5): 737.
[16] Dias J A, Osegovic J P, Drago R S. The solid acidity of 12tungstophosphoric acid. Journal of Catalysis, 1999, 183(1):
83.
[17] Mizuno N, Misono M. Heterogeneous catalysis. Chemical Reviews, 1998, 98(1): 199.
[18] Caliman E, Dias J A, Dias S C L, Garcia F A C, de Macedo J L,
Almeida L S. Preparation and characterization of H3PW12O40
supported on niobia. Microporous and Mesoporous Materials,
2010, 132(1-2): 103.
[19] Cai H S, Liu G G, Lv W Y, Li X X, Yu L, Li D G. Effect of
Ho-doping on photocatalytic activity of nanosized TiO2 catalyst. Journal of Rare Earths, 2008, 26(1): 71.

You might also like