Vous êtes sur la page 1sur 9

RSC Advances

View Article Online


PAPER View Journal | View Issue

Facile synthesis of WO3 with reduced particle size


on zeolite and enhanced photocatalytic activity†
Cite this: RSC Adv., 2014, 4, 21221
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

K. Jothivenkatachalam,*a S. Prabhu,a A. Nithyaa and K. Jeganathanb

WO3 supported on zeolite-Y (WO3-ZY) was successfully synthesized by a facile impregnation method and
well characterized by various techniques. The photocatalytic activity of the prepared catalysts was
investigated for the degradation of Rhodamine B (RhB) under visible, UV and solar light irradiation. The
enhanced photocatalytic activity was observed for the catalyst WO3-ZY, which may be due to the
presence of more active sites that can adsorb a greater number of dye molecules. The TEM, FESEM and
adsorption studies confirm that the WO3 supported on zeolite-Y has a very small particle size of about
Received 17th February 2014
Accepted 11th April 2014
8 nm compared with the bare WO3 at 97 nm. The efficient electron–hole pair separation and the role of
active species were investigated by photoluminescence spectroscopy and the test of the effect of
DOI: 10.1039/c4ra01376j
scavengers, respectively. The mechanism for the photocatalytic degradation of RhB was proposed and
www.rsc.org/advances the pathway of RhB degradation was illustrated schematically.

morphology and particle size,9–12 semiconductor coupling,13–16


Introduction noble metal deposition,7,14,15,17 and metal ion doping.18
Semiconductor photocatalysis is a promising technique for Increasing the surface area and suppressing the electron–hole
addressing environmental issues such as energy storage and pair recombination can also enhance the photocatalytic activity
decontamination of environmental pollution.1 This technique of WO3,8 but there are few reports available on tuning the
provides clean and recyclable hydrogen energy and also can physical properties of WO3 to improve the efficiency under
utilize solar energy to decompose organic and inorganic visible light irradiation.19
pollutants present in air and aqueous media.2,3 TiO2 is the most Recently, researchers have intensively focused on meso-
widely used and best photocatalytic material; however, only a porous materials such as zeolite as a support for metal oxides
small fraction of solar light (3–5%) can be utilized due to its that inuence photocatalytic activity through structural modi-
wide band gap.4,5 Even though the visible light photocatalytic cation.20–23 The zeolitic materials have gained signicant
activity of TiO2 has been reported in nitrogen-doped TiO2, the importance due to their high surface area, thermal stability and
quantum efficiency of the nitrogen-doped TiO2 is much lower their specic photophysical properties in charge and electron
than that under UV light irradiation.6 In contrast, WO3 is n-type transfer processes.24,25 The use of zeolite as a support for metal
semiconductor that has strong absorption in the solar spec- oxides improves the amount of photons absorbed by the cata-
trum, stable physicochemical properties, and is resistant to lyst and reduces the amount of metal oxides required.26 Despite
photocorrosion effects.7 The narrow band gap of about 2.4– that there are several metal oxides supported on zeolitic mate-
2.8 eV and deeper valance band (+3.1 eV) results in additional rials that proved to be better photocatalysts, there is no report
advantages for visible light driven photocatalysis.7 However, on WO3 supported on zeolitic materials for photocatalytic
due to fast recombination of electron–hole pairs and low applications.
conduction band level, pure WO3 is not an efficient photo- For the rst time, we report a facile synthesis of WO3 sup-
catalyst.8 However, it has been found that the photocatalytic ported on zeolite-Y catalyst by impregnation method and its
efficiency of the pure WO3 can be enhanced by several methods. photocatalytic efficiency for the degradation of Rhodamine B
Some examples are tuning a physical property such as (RhB) under visible, UV and solar light irradiation. The
prepared catalysts were well characterized by powder X-ray
diffraction (PXRD), X-ray photoelectron spectroscopy (XPS),
a
Department of Chemistry, Anna University, BIT Campus, Tiruchirappalli-620 024,
transmission electron microscopy (TEM), eld emission scan-
Tamil Nadu, India. E-mail: jothivenkat@yahoo.com
b
ning electron microscope (FESEM), Raman, Fourier transform
Centre for Nanoscience and Nanotechnology, School of Physics, Bharathidasan
University, Tiruchirappalli-620 024, Tamil Nadu, India infrared spectroscopy (FT-IR) and UV-vis diffuse reectance
† Electronic supplementary information (ESI) available: PXRD spectrum of spectra (UV-DRS) techniques. Greatly enhanced photocatalytic
zeolite-Y, UV-vis. Spectral changes of RhB under visible, UV and solar light activity was observed in the catalyst WO3 loaded on zeolite-Y.
irradiation without any catalyst and % degradation of RhB under UV light The TEM, FESEM and adsorption studies conrmed the
irradiation by different catalyst. See DOI: 10.1039/c4ra01376j

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 21221–21229 | 21221
View Article Online

RSC Advances Paper

reduced particle size of the catalyst. The efficient electron–hole spectra were recorded on Shimadzu UV-2450 UV-visible spec-
pair separation of the catalysts was investigated by photo- trophotometer equipped with an integrated sphere assembly
luminescence (PL) spectroscopic techniques. The photocatalytic using BaSO4 as the reectance sample. FT-IR analysis was per-
activity in the presence of different scavengers was demon- formed using Jasco FTIR 460 plus spectrophotometer. The PL
strated by testing which of the active species plays an important spectrum was recorded by a JASCO FP-6500 spectrouorometer
role in the degradation of the RhB. The mechanism for the with 300 nm excitation wavelength.
photocatalytic degradation of RhB is proposed; the pathway of
RhB degradation is also illustrated schematically.
Photocatalytic test
The photocatalytic degradation of a model pollutant RhB was
Experimental chosen to investigate the photocatalytic activity of the prepared
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

Materials catalysts. In a typical experiment, 25 mg of the catalyst was


All analytical grade reagents of Na2WO4$2H2O, H2WO4, RhB, suspended in 50 mL of aqueous solution contains 10 mg L1 of
BaSO4 and zeolite-Y were used as such without further puri- RhB dye. Prior to irradiation, the suspension was held for
cation. Deionized and double distilled water was used 30 min in the dark to achieve adsorption–desorption equilib-
throughout this work. rium by aeration. Photocatalytic activity was investigated under
visible and UV light irradiation by 300 W tungsten halogen lamp
Synthesis (8500 lumen) and 125 W medium pressure mercury lamp
emitting 365 nm (110 mW cm2, measured by a Lutron UV light
The WO3-supported zeolite-Y (WO3-ZY) catalyst was prepared by Meter) light respectively. The reaction mixture was aerated
the facile impregnation method as follows. The zeolite-Y (1 g) continuously under irradiation until the reaction mixture was
was put into an aqueous medium and stirred. Aer that, 5 g of thorough mix. At the given time interval 3.5 mL of the sample
H2WO4 was added into the above reaction mixture. Then, the was withdrawn and centrifuged to separate the catalyst. The
reaction mixture was reuxed at 100  C for 24 h. The obtained degradation of RhB was determined by measuring the
precipitate was collected by centrifugation, washed several maximum absorbance at 554 nm on a UV-vis
times using distilled water and dried. The dried sample was spectrophotometer.
calcinated at 500  C for 3 h. For comparison, pure WO3 was
prepared by the precipitation method as reported in the litera-
ture.27 The sodium tungstate was dissolved in distilled water Results and discussion
and slowly heated to 85  C. The appropriate amount of warm,
Characterization
concentrated nitric acid was added drop-wise to this solution
with vigorous stirring. The mixed solution was held at the same Crystalline structure. The PXRD pattern of the prepared
temperature for 30 min under continuous stirring. The precip- catalysts is shown in Fig. 1. The peaks obtained in the pattern
itate was allowed to settle overnight at room temperature, then are indexed with standard data (JCPDS no. 83-0950) and corre-
washed by adding a large amount of water and allowing the spond to the monoclinic phase structure. In pure WO3, the
precipitates to settle before decanting the liquid. Finally, the pattern shows very sharp peaks that represent the high crys-
precipitate was separated by centrifugation and dried. Aer tallinity and greater particle size. The diffraction peaks
drying, the pure WO3 was calcined at 500  C for 3 h. appeared in the WO3-ZY corresponds to WO3 indicating the
formation of WO3 on the zeolite-Y (please see the ESI† for the
PXRD pattern of pure zeolite-Y). On the other hand, there are no
Characterization
sharp peaks in the diffraction pattern of WO3-ZY that could be
PXRD data were collected by a PAN analytical X'pert Pro dual
goniometer diffractometer. The data were collected with a step
size of 0.008 and a scan rate of 0.5 min1. The radiation used
was Cu-Ka (1.5418 Å) with a Ni lter, and the data collection was
carried out using a at holder in Bragg–Brentango geometry.
XPS analysis was carried out by Kratos Axis-Ultra DLD instru-
ment by passing 200 eV energy. During XPS analysis the sample
was irradiated with Mg Ka and the data collected at sweep rates
of 2 eV and 5 eV in a wide and narrow range scan, respectively.
TEM analysis was done by JEOL model 2010 FasTEM instru-
ment with a 200 kV accelerating voltage. An FESEM image was
obtained by a Carl Zeiss SIGMA instrument with 1.2 nm reso-
lution. The energy-dispersive X-ray spectra (EDX) were recorded
using the INCAx-act Oxford Instrument equipped with SEM.
Raman spectra were recorded by employing a Horiba JY Lab-
RAMHR 800 Raman spectrometer coupled with a microscope in
reectance mode with a 514 nm excitation laser source. UV-DRS Fig. 1 PXRD pattern of the WO3 and WO3-ZY.

21222 | RSC Adv., 2014, 4, 21221–21229 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

attributed to the smaller size particles of WO3. The crystalline catalysts WO3 and WO3-ZY were conrmed by the EDX analysis
size of the WO3 and WO3-ZY in the (002) plane is calculated by and are shown in Fig. 3(a and b).
the Scherer formula to be around 43 and 26 nm, respectively. XPS analysis. The elements and chemical states present in
This result shows that the growth of WO3 on zeolite-Y is very the catalysts were further investigated by XPS analysis. The
limited and controlled. overview XPS spectra of the samples WO3 and WO3-ZY are
Morphology and particle size. The FESEM image of the WO3 shown in Fig. 4(a). The spectra clearly show the presence of
sample shows spherical shaped particles with the average size W4d, W4f and O1s peaks and the respective oxidation states of
of 102 nm, as shown in Fig. 2(a). The FESEM image of WO3-ZY the elements are indexed. The narrow range spectra of W4d and
shows very small particles (12 nm) as shown in Fig. 2(c). Since
there are no distinct particles, it is difficult to obtain particle
size from the image. For comparison, the FESEM image of
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

zeolite-Y is shown in Fig. 2(b). The TEM image of WO3 is given


in Fig. 2(d), which further conrms the shape of the particles.
The average particle size was calculated from the TEM image to
be around 97 nm, which is close to that calculated from FESEM
image. The highly magnied TEM image of WO3 is shown in
Fig. 2(e). The inter-planar distance from the TEM image was
calculated to be 0.39 nm, which is well matched with the lattice
spacing of (200) planes of the monoclinic WO3 structure. The
TEM image of WO3-ZY shows that the particles are very small in
size, as shown in Fig. 2(f); the particle size distribution is shown
in the inset. The calculated average particle size of WO3-ZY from
Fig. 2(f) is around 8 nm. These FESEM and TEM results shows
that the WO3 particle size was reduced from 97 nm to 8 nm
under the given reaction conditions. This decrease in particle
size may provide more active sites. The elements present in the
Fig. 3 EDX spectra of (a) WO3; and (b) WO3-ZY.

Fig. 2 FESEM images of (a) WO3, (b) zeolite-Y, (c) WO3-ZY, (d and e); TEM and HR TEM images of WO3; and (f) TEM image of WO3-ZY.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 21221–21229 | 21223
View Article Online

RSC Advances Paper


Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

Fig. 5 (a) Raman and (b) FT-IR spectra of WO3 and WO3-ZY.

Fig. 4 (a) XPS survey spectra of WO3 and WO3-ZY, and narrow range
survey spectra of (b) W4d, (c) O1s of WO3 and WO3-ZY.
were observed in the 500–900 cm1 range, which corresponds to
the n(O–W–O) stretching mode.33 The peak at around 750 cm1
is attributable to the n(O–W–O) stretching mode.34 The bands in
O1s corresponding to WO3 and WO3-ZY are shown in Fig. 4(b
the range of 3200–3550 cm1 are ascribed to the n(O–H)
and c). The binding energy observed at around 247 and 530 eV
stretching, and the band at 1625 cm1 corresponds to the d(O–
in Fig. 4(b and c) is ascribed to W4d and O1s of WO3, respec-
H) bending modes of the coordinated water.33
tively. An obvious shi of W4d and O1s peaks in WO3-ZY to a
Optical properties. The optical properties of the catalysts
higher binding energy is observed in the spectra, which is
were investigated by UV-vis DRS spectroscopy; results are shown
attributed to the strong interaction between WO3 and zeolite-Y.
in Fig. 6(a). The spectra reveal that all catalysts have absorption
The O1s peak shi from 530 eV to 538 eV is attributed to the
edges above a wavelength of 400 nm. The band gap energies of
oxygen of WO3 hydrogen bonded to the zeolite surface.28 On the
the samples can be calculated by the following equation35,36
other hand, the particle size variation and lattice variation can
also change the binding energy in XPS. The particle size
ahn ¼ A(hn  Eg)n/2 (1)
reduction will result in a higher binding energy shi, which
clearly shows the reduction of WO3 particle size in WO3-ZY.29 where a, n, Eg and A are the absorption coefficient, frequency of
This reduction in particle size is in good agreement with results the light, band gap energy and a constant, respectively; and n is
obtained from FESEM and TEM analysis. the type of optical transition of a semiconductor, which is 1. The
Raman spectral analysis. The Raman spectra of the prepared band gap energy (Eg) of the catalysts can be calculated from the
catalysts consist of three main regions, less than 200, 200–400 intercept of the tangent to the X axis from a plot of (ahn)1/2
and 600–900 cm1, as shown in Fig. 5(a), which was well versus energy (hn), as given in Fig. 6(b). The calculated band gap
matched with the literature report of the monoclinic tungsten energies of the samples are 2.52 and 2.73 eV for WO3 and WO3-
oxide phase.30,31 The Raman peaks at 135 and 184 cm1 corre- ZY, respectively. The valance band (VB) edge potential of a
spond to the (W2O2)n chains.30 The W–O–W bending modes of semiconductor at the point of zero charge can be calculated
bridging oxide ion peaks appear at 273 and 329 cm1.30 The from the following equation37
peaks at 720 and 809 cm1 are attributable to the W–O–W
stretching mode in the tungsten oxide network.32 The intensity EVB ¼ X  Ec + 0.5Eg (2)
of the Raman peaks was very low for the catalyst WO3-ZY, which
may be due to the very small particle size. where EVB is the VB edge potential; X is the electronegativity of
FT-IR analysis. The FT-IR spectral characterization was the semiconductor; Ec is the energy of free electrons on the
carried out to analyze the catalysts; the obtained spectra are hydrogen scale (about 4.5 eV); and Eg is the band gap energy of
shown in Fig. 5(b). The strong absorption peaks for pure WO3 the semiconductor. The conduction band (CB) edge potential

21224 | RSC Adv., 2014, 4, 21221–21229 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

ln(C0/C) ¼ kappt (4)

where C0 is the initial concentration of the dye solution


(mol L1); C is the concentration of the dye solution at time t
(mol L1); and kapp is the apparent rate constant (min1).
The rate constant for the degradation of RhB was calculated
from the slope of the plot, ln(C0/C), vs. irradiation time.
Under visible light. The photocatalytic activity for the
degradation of RhB by the catalysts under visible light irradia-
tion is shown in Fig. 7(a). The enhanced photocatalytic activity
was observed by the catalyst WO3-ZY, which may be due to high
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

percentage adsorption of dye on the surface of the catalyst. The


percentage adsorption of RhB on the surface of the catalyst
under dark aer 30 min is shown in the inset gure of Fig. 7(a).
The high percentage adsorption of about 30% was achieved by
the catalyst WO3-ZY, whereas WO3 shows less than 10%
adsorption. This high percentage of adsorption may be due to
decreasing the particle size of WO3 on zeolite-Y, which may
provide more active sites to adsorb dye molecules. This study
experimentally conrms the reduced particle size of the catalyst
WO3-ZY. The photocatalytic activity of zeolite-Y catalyst for the
degradation of RhB is very minimum in visible light when
compared with WO3-ZY, as shown in Fig. 7(a).
Optimization study. To effectively utilize the catalyst in a low
amount and get high efficiency at an optimum dye concentra-
tion, two sets of experiments were done. First, we changed the
Fig. 6 (a) UV-vis DRS and (b) plot of (ahn)1/2 vs. hn of WO3 and WO3-ZY. amount of WO3-ZY from 0.25 g L1 to 1 g L1 while keeping the
10 mg L1 dye concentration constant. Second, we changed the
dye concentration from 5 mg L1 to 20 mg L1 while keeping
(ECB) can be determined by ECB ¼ EVB  Eg. The X value for WO3 0.5 g L1 catalyst constant. The increased percentage degrada-
is about 6.60, and the calculated EVB and ECB of the catalysts are tion of 58% to 92% was observed when the amount of the
3.36 and 0.84 eV, respectively, for WO3 and 3.46 and 0.73 eV, catalyst was increased from 0.25 g L1 to 0.5 g L1, respectively.
respectively, for WO3-ZY. The band gap energy changes and blue This may be due to the increased availability of active sites on
shi in the UV-vis spectrum clearly suggest that the particle size the catalyst surface, thereby increasing the available amount of
of the catalyst has decreased. catalyst. Upon further increasing the amount of catalyst, the
Photocatalytic activity. The photocatalytic activity of the degradation efficiency does not increase, as shown in Fig. 7(b).
prepared catalysts for the degradation of RhB was demonstrated This may be due to the aggregation of a large amount of catalyst,
under visible, UV and solar light irradiation. Before that, the which causes a reduction in the number of active sites for the
blank experiment without any catalyst was carried out to test the adsorption of dye molecules and photons. Moreover, a high
stability of RhB under visible, UV and solar light irradiation. concentration of catalyst creates turbidity and thus reduces the
The UV-vis spectral changes of RhB under various light irradi- penetration intensity of light radiation by the scattering
ations were recorded (see Fig. S1–S3 in the ESI†). There were no effect.40,41 As the dye concentration increases, the photocatalytic
signicant changes observed in the UV-vis spectrum, which activity decreases, as shown in Fig. 7(c). This may be due to a
conrms the degradation of RhB is almost zero or negligible decreased number of photons reaching the surface of the
under the various light irradiations in the absence of any cata- catalyst because more photons are absorbed by dye molecules.42
lyst. These experimental results show RhB in aqueous medium These experimental results suggest that the optimum reaction
is highly stable under the given light irradiation. The percent condition to achieve better efficiency is 0.5 g L1 catalyst and
degradation of RhB was calculated by the following equation 10 mg L1 dye concentration.
Under UV and solar light. The photocatalytic efficiency of the
%D ¼ (A0  At)/A0  100 (3) catalyst WO3-ZY was investigated under UV and solar light
irradiation. The photocatalytic degradation of about 86% was
where %D is the percentage of degradation, and A0 and At are observed at 140 min under UV light irradiation, which is a very
initial absorption and absorption at time t, respectively. high efficiency when compared with WO3 (see Fig. S4 in the ESI†
Photocatalytic degradation of organic compounds usually for the degradation prole of different catalysts); this result is
follows the pseudo-rst order kinetics model. The rate constant consistent with that obtained under visible light irradiation.
for the degradation of RhB is calculated using the following Under solar light irradiation, 95% degradation was observed at
Langmuir–Hishelwood kinetic equation38,39 180 min. The rate constants for the degradation of RhB under

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 21221–21229 | 21225
View Article Online

RSC Advances Paper


Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

Fig. 7 Percent degradation of RhB under visible light irradiation (a) over different catalysts, over WO3-ZY (b) on varying the amount of catalyst, (c)
on varying the amount of RhB and (d) under different light irradiation.

different light irradiations are given in Fig. 7(d). The rate degradation of RhB over WO3-ZY in the presence of air,
constant increased in the order visible < UV < solar light irra- hydrogen peroxide (5 mmol) and no acceptor under visible light
diation. The high rate constant under solar irradiation sug- irradiation are shown in Fig. 8(b). The kapp of 0.0052 min1 in
gested that the catalyst can be efficiently used for the conversion the absence of an electron acceptor increased to 0.0143 min1
of solar energy. aer the addition of hydrogen peroxide. This enhanced photo-
catalytic activity in the presence of electron acceptors may be
due to a decreased recombination rate of electron–hole pairs
Reaction mechanism and degradation pathway and formation of more active radicals.
Recombination of electron–hole pair. The photocatalytic Role of active species. The holes (h+) and radical trapping
efficiency of a catalyst can be enhanced by inhibiting the experiments were carried out to investigate the role of reactive
recombination of photo-generated electron–hole pairs. The species in the photocatalytic degradation of RhB over WO3-ZY
lower PL intensity of the catalyst can have a lower recombina- under visible light irradiation. The changes of kapp for the
tion rate and a higher photocatalytic activity as the recombi- degradation of RhB in the presence of different scavengers (5
nation of electron–hole pair releases energy in the form of mmol) are shown in Fig. 9(a). When t-BuOH (as the cOH scav-
uorescence.43,44 The PL spectra of the catalysts are shown in enger) was added,45 the kapp decreased slightly to 0.0035 min1
Fig. 8(a). The PL intensity of the catalyst WO3-ZY shows the from 0.0052 min1. On the other hand, the kapp of RhB degra-
lowest intensity compared with pure WO3. It is suggested that dation decreased to 0.0023 min1 when benzoquinone (BQ) was
WO3-ZY has the lowest electron–hole pair recombination rate, added as the cO2 scavenger.46,47 However, upon the addition of
which is consistent with the high photocatalytic activity ammonium oxalate (AO) as the h+ scavenger,48 the kapp
observed. This decrease in the PL intensity may be due to the decreased radically, to 0.0005 min1. These decreases in the
small particle size of the catalyst, which can increase the surface kapp upon the addition of different scavengers suggested that h+
area and thereby increase the interfacial charge–carrier and cO2 play a major role, whereas cOH is a minor active
transfer.19 species for the degradation of RhB. The active radical cO2 can
Role of electron acceptors. Electron acceptors such as air and be formed by reacting photo-generated electrons with O2
hydrogen peroxide can create more active radicals of cO2 and molecules adsorbed on the surface of the catalyst,1 which
cOH, respectively, which are strong enough to degrade organic indicates that O2 is an efficient electron acceptor for generating
molecules. The experimental results for the photocatalytic cO2 and inhibiting electron–hole pair recombination.49,50

21226 | RSC Adv., 2014, 4, 21221–21229 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

e 
CB + O2 / cO2 + catalyst (6)

RhB + cO2 / products (7)

RhB + h+VB / products (8)

It is well known that the degradation of RhB proceeds via two


processes of N-de-ethylation and destruction of the conjugated
structure.51 The temporal UV-vis absorption spectral variation
for the photocatalytic degradation of RhB under solar light
irradiation is shown in Fig. 9(c). The absorption band at 554 nm
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

has decreased, which suggests the degradation of the xanthene


ring in RhB. The blue shi in the absorption spectrum is
ascribed to the formation of de-ethylated RhB.52 Based on the
results obtained from the UV-vis spectral changes, we propose
the degradation pathway of RhB schematically in Scheme 1: the
degradation of RhB occurred through de-ethlyation process,
followed by the destruction of the conjugated xanthane struc-
ture in RhB, which produces the benzenoid intermediates.53

Fig. 8 (a) PL spectrum of the catalysts and (b) effect of electron


acceptor for the degradation of RhB.

Possible reaction mechanism and degradation pathway


Based on the effect of scavengers, a possible mechanism for the
photocatalytic degradation of RhB over WO3-ZY is illustrated in
Fig. 9(b). As shown in Fig. 9(b), the electron–hole pair is created
by the illumination of the catalyst. Then the photoinduced
electrons in the CB react with an O2 molecule adsorbed on the
surface of the catalyst and form cO2 radicals to degrade RhB.
Meanwhile, the photoinduced holes in the VB can directly
degrade RhB. The photocatalytic degradation of RhB can be
illustrated as follows:
Scheme 1 Proposed degradation pathway for the photocatalytic
WO3 + hn / e +
CB + hVB (5) degradation of RhB.

Fig. 9(a) Effect of different scavengers; (b) possible mechanism; and (c) the temporal UV-vis absorption spectral variation for the photocatalytic
degradation of RhB over WO3-ZY.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 21221–21229 | 21227
View Article Online

RSC Advances Paper

Aer that occurs, the schematic diagram shows the ring 15 S. M. Sun, W. Z. Wang, S. Z. Zeng, M. Shang and L. Zhang,
opening of the intermediates and mineralization process. J. Hazard. Mater., 2010, 178, 427.
16 D. Su, J. Y. Wang, Y. P. Tang, C. Liu, L. F. Liu and X. J. Han,
Conclusion Chem. Commun., 2011, 47, 4231.
17 J. Kim, C. W. Lee and W. Choi, Environ. Sci. Technol., 2010,
In this study, we have successfully synthesized and thoroughly 44, 6849.
characterized both pure WO3 and WO3 supported on zeolite-Y. 18 X. F. Cheng, W. H. Leng, D. P. Liu, J. Q. Zhang and C. N. Cao,
The TEM, FESEM and adsorption studies conrm that the Chemosphere, 2007, 68, 1976.
catalyst WO3-ZY has a very small particle size: 8 nm. The pho- 19 A. Purwanto, H. Widiyandari, T. Ogi and K. Okuyama, Catal.
tocatalytic activity of the prepared catalysts was investigated for Commun., 2011, 12, 525.
the degradation of RhB under visible, UV and solar light irra- 20 W. Zhang, K. Wang, Y. Yu and H. He, Chem.–Eng. J., 2010,
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

diation. The enhanced photocatalytic activity was observed for 163, 62.
the catalyst WO3-ZY. The higher photocatalytic activity of WO3- 21 A. Nezamzadeh-Ejhieh and Z. Salimi, Desalination, 2011,
ZY under solar light irradiation compared with other light 280, 281.
irradiation sources suggests the catalyst can be utilized for solar 22 Z. M. El-Bahy, M. M. Mohamed, F. I. Zidan and M. S. Thabet,
energy conversion. The efficient electron–hole pair separation J. Hazard. Mater., 2008, 153, 364.
of the catalyst provides the condition whereby enhanced pho- 23 A. Nezamzadeh-Ejhieh and Z. Banan, Desalination, 2011,
tocatalytic activity is obtained, as conrmed by PL spectra. The 279, 146.
test of the effect of scavengers shows that the active species h+ 24 H. Chen, A. Matsumoto, N. Nishimiya and K. Tsutsumi,
and cO2 play a major role in the photocatalytic degradation of Colloids Surf., A, 1999, 157, 295.
RhB. The proposed mechanism for the photocatalysis and 25 M. Alvaro, E. Carbonell, M. Espla and H. Garcia, Appl. Catal.,
degradation pathway of RhB is illustrated: it involves N-de- B, 2005, 57, 37.
ethylation, destruction of conjugated structure, ring opening 26 M. Aleksic, H. Kusic, N. Koprivanac, D. Leszczynska and
and mineralization processes. A. L. Bozic, Desalination, 2010, 257, 22.
27 S. Supothina, P. Seeharaj, S. Yoriya and M. Sriyudthsak,
Acknowledgements Ceram. Int., 2007, 33, 931.
28 I. Gouzman, M. Dubey, M. D. Carolus, J. Schwartz and
The authors thank DST-SERC for nancial support, sanction no. S. L. Bernasek, Surf. Sci., 2006, 600, 773.
SR\FT\CS-042\2008. 29 B. Richter, H. Kuhlenbeck, H. J. Freund and P. S. Bagus,
Phys. Rev. Lett., 2004, 93, 26805.
Notes and references 30 B. Pecquenard, H. Lecacheaux, L. Livage and C. Julien,
J. Solid State Chem., 1998, 135, 159.
1 M. R. Hoffmann, S. T. Martin, W. Choi and 31 M. Boulova and G. Lucazeau, J. Solid State Chem., 2002, 167,
D. W. Bahnemann, Chem. Rev., 1995, 95, 69. 425.
2 S. C. Zhang, J. D. Shen, H. B. Fu, W. Y. Dong, Z. J. Zheng and 32 H. Habazaki, Y. Hayashi and H. Konno, Electrochim. Acta,
L. Y. Shi, J. Solid State Chem., 2007, 180, 1456. 2002, 47, 4181.
3 A. Fujishima and K. Honda, Nature, 1972, 238, 37. 33 H. I. S. Nogueira, A. M. V. Cavaleiro, J. Rocha, T. Trindade
4 W. K. Ho, J. C. Yu and S. C. Lee, Chem. Commun., 2006, 1115. and J. D. P. D. Jesus, Mater. Res. Bull., 2004, 39, 683.
5 Y. Cong, J. L. Zhang, F. Chen and M. Anpo, J. Phys. Chem. C, 34 C. Guery, C. Choquet, F. Dujeancourt, J. M. Tarascon and
2007, 111, 6976. J. C. Lassegues, J. Solid State Electrochem., 1997, 1,
6 H. Irie, Y. Watanabe and K. Hashimoto, J. Phys. Chem. B, 199.
2003, 107, 5483. 35 M. A. Butler, J. Appl. Phys., 1977, 48, 1914.
7 G. R. Bamwenda and H. Arakawa, Appl. Catal., A, 2001, 210, 36 J. Zeng, H. Wang, Y. C. Zhang, M. K. Zhu and H. Yang,
181. J. Phys. Chem. C, 2007, 111, 11879.
8 Z. G. Zhao and M. Miyauchi, Angew. Chem., Int. Ed., 2008, 47, 37 X. Zhang, L. Z. Zhang, T. F. Xie and D. J. Wang, J. Phys. Chem.
7051. C, 2009, 113, 7371.
9 D. Chen and J. H. Ye, Adv. Funct. Mater., 2008, 18, 1922. 38 J. G. Yu, H. G. Yu, B. Cheng, X. J. Zhao, J. C. Yu and W. K. Ho,
10 Y. F. Guo, X. Quan, N. Lu, H. M. Zhao and S. Chen, Environ. J. Phys. Chem. B, 2003, 107, 13871.
Sci. Technol., 2007, 41, 4422. 39 H. Kumazawa, M. Inoue and T. Kasuya, Ind. Eng. Chem. Res.,
11 Z. G. Zhao and M. Miyauchi, J. Phys. Chem. C, 2009, 113, 2003, 42, 3237.
6539. 40 A. Akyol, H. C. Yatmaz and M. Bayramoglu, Appl. Catal., B,
12 K. Z. Lv, J. Li, X. X. Qing, W. Z. Li and Q. Y. Chen, J. Hazard. 2004, 54, 19.
Mater., 2011, 189, 329. 41 B. Neppolian, H. C. Choi, S. Sakthivel, B. Arabindoo and
13 J. Papp, S. Soled, K. Dwight and A. Wold, Chem. Mater., 1994, V. Murugesan, Chemosphere, 2002, 46, 1173.
6, 496. 42 M. A. Rauf and S. S. Ashraf, Chem.–Eng. J., 2009, 151, 10.
14 H. W. Choi, E. J. Kim and S. H. Hahn, Chem.–Eng. J., 2010, 43 T. J. Cai, M. Yue, X. W. Wang and Q. Deng, Chin. J. Catal.,
161, 285. 2007, 28, 10.

21228 | RSC Adv., 2014, 4, 21221–21229 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

44 H. Tang, K. Prasad, R. Sanjines, P. E. Schmid and F. Levy, 49 Y. Y. Li, J. S. Wang, H. C. Yao, L. Y. Dang, Z. J. Li and J. Mol,
J. Appl. Phys., 1994, 75, 2042. J. Mol. Catal. A: Chem., 2011, 334, 116.
45 T. Xu, L. Zhang, H. Cheng and Y. Zhu, Appl. Catal., B, 2011, 50 Y. Q. Yang, G. K. Zhang, S. J. Yu and X. Shen, Chem.–Eng. J.,
101, 382. 2010, 162, 171.
46 J. Bandara and J. Kiwi, New J. Chem., 1999, 23, 717. 51 C. Chen, X. Li, W. Ma, J. Zhao, H. Hidaka and N. Serpone,
47 M. C. Yin, Z. S. Li, J. H. Kou and Z. G. Zou, Environ. Sci. J. Phys. Chem. B, 2002, 106, 318.
Technol., 2009, 43, 8361. 52 N. Bao, X. Feng, Z. Yang, L. Shen and X. Lu, Environ. Sci.
48 N. Zhang, S. Q. Liu, X. Z. Fu and Y. J. Xu, J. Phys. Chem. C, Technol., 2004, 38, 2729.
2011, 115, 9136. 53 Z. He, C. Sun, S. Yang, Y. Ding, H. He and Z. Wang, J. Hazard.
Mater., 2009, 162, 1477.
Published on 15 April 2014. Downloaded by University of Regina on 25/08/2014 11:32:54.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 21221–21229 | 21229

Vous aimerez peut-être aussi