Вы находитесь на странице: 1из 10

Chemical Engineering Journal 223 (2013) 737–746

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Fly ash cenospheres supported visible-light-driven BiVO4 photocatalyst:


Synthesis, characterization and photocatalytic application
Jin Zhang a,b, Hao Cui a, Bing Wang a, Chuang Li a, Jianping Zhai a, Qin Li a,⇑
a
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210046, PR China
b
School of Biochemical and Environmental Engineering, Nanjing Xiaozhuang University, Nanjing 211171, PR China

h i g h l i g h t s g r a p h i c a l a b s t r a c t

" A novel BiVO4/FACs composite was


prepared as a photocatalyst for
degradation MB.
" The catalyst is active under visible-
light irradiation.
" The catalyst is floatable and can be
recovered easily.
" The solid wastes FACs were reused
effectively.
" Two main reactive radical groups
were confirmed by the ESR spin-
trapping with DMPO.

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

Article history: A novel BiVO4/Fly ash cenospheres (FACs) composite photocatalyst was prepared by the modified metal-
Received 18 August 2012 organic decomposition (MOD) method. And the photocatalyst was characterized by X-ray diffraction
Received in revised form 17 December 2012 (XRD), scanning electron microscope (SEM), X-ray photoelectron spectra (XPS), and UV–vis diffused
Accepted 26 December 2012
reflectance spectroscopy (DRS) techniques. DRS revealed that the absorption threshold of BiVO4/FACs
Available online 7 January 2013
shifted to a longer wavelength compared with the pure BiVO4 photocatalyst. Adsorption parameters
and photocatalytical activity under visible light irradiation were evaluated using the methylene blue
Keywords:
(MB) dye as a model. FACs, the solid wastes produced in coal-firing power plants, were used as supports
BiVO4
Fly ash cenospheres (FACs)
with the advantages of low cost, and also improved the adsorption and photocatalytic activity of BiVO4.
Visible-light-responsive The maximum amount of dye adsorbed (Qmax) for BiVO4/FACs is 1.8 times more than that for pure BiVO4,
Photocatalysis and the photodegradation first-order rate constant for BiVO4/FACs is 2.5 times higher than that for pure
Composite materials BiVO4. Owing to the low density of FAC, the as-prepared BiVO4/FAC particles float in water to favor phase
separation and the recovery of the photocatalyst after the reaction. The recovery test shows that BiVO4/
FACs composite was rather stable during the MB photodegradation. Based on electron spin-resonance
spectroscopy (ESR) results, a reasonable reaction mechanism has also been proposed.
Ó 2013 Elsevier B.V. All rights reserved.

1. Introduction sponds to UV light occupying ca. 4% of the whole solar energy


due to its wide band gap energy (3.2 eV), which hinder its practical
Semiconductor photocatalysis as a green energy technology is application in water purification greatly. Therefore, many efforts
attracting extensive attention with regards environmental cleaning have been carried out to exploit visible-light-driven photocatalyst
and water splitting. In most cases, photocatalytic degradation is with high activity in the last decades [1–4].
conducted over the TiO2 photocatalyst, however, TiO2 only re- Recently, a large number of undoped, single-phase semiconduc-
tor photocatalysts with particular absorption abilities in the visible
light range have been developed, such as bismuth vanadate
⇑ Corresponding author. Tel./fax: +86 25 8359 2903. (BiVO4) [5–9], Bi2WO6 [10], Bi4TaO8I [11], InMO4 (M = V, Nb, Ta)
E-mail address: qli@nju.edu.cn (Q. Li).

1385-8947/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.12.065
738 J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746

[12] and so on. Among these photocatalysts, BiVO4, with a mono- Then, the sieved FACs were treated ultrasonically in 10% dilute ni-
clinic scheelite structure, has attracted considerable attention for tric acid for 1 h. Finally, these particles were filtered and washed
its excellent photocatalytic performance under visible light irradi- with deionized water, followed by drying in a vacuum drying oven
ation. Furthermore, monoclinic BiVO4 has band gap energy of at 120 °C for 3 h.
2.4 eV and can absorb light from the solar spectrum up to the blue
light fraction of ca. 520 nm. This is more practical than traditional 2.2.2. Preparation of KH550-grafted FACs
TiO2 in terms of the efficient utilization of visible light [13]. Never- The FAC surface was modified organofunctionally using silane
theless, there are some drawbacks in using BiVO4 in powder form coupling agents KH550, and the synthesis of KH550-grafted FACs
during the photocatalytic process: (1) the powder BiVO4 separates was discussed in paper [24]. In a typical experiment, 2 mL KH550
difficultly from water after reaction; (2) the suspended powder was first hydrolyzed in 200 mL of water/ethanol (1:9 v/v) solution,
BiVO4 tends to aggregate especially at high concentrations [14]. and then 2 g FACs were added to the solution with stirring for 6 h
One effective approach is to immobilize the catalysts on a support in a constant temperature water bath at 80 °C. These modified FACs
which can overcome these problems [15,16]. Some recent studies were filtered and rinsed with ethanol and deionized water, fol-
have reported on the use of MCM-41 [17], graphene [18], and car- lowed by drying overnight in a vacuum oven at 100 °C. The sam-
bon spheres [19] as a support to increase the photocatalytic effi- ples prepared were named KH550-grafted FACs.
ciency and immobility of the BiVO4 catalysts. However, the above
mentioned supports such as molecular sieve and carbon-based 2.2.3. Synthesis of BiVO4/FACs composite
nano-materials have high cost and difficult preparation, which re- The BiVO4 films were prepared by the MOD method [25]. In a
strict their widely application on photocatalysis. Therefore, FACs typical process, 7.27 g of Bi(NO3)55H2O was dissolved in 75 mL
are chosen as support to coat BiVO4 film on it in this paper. of acetic acid (0.2 mol L1). After magnetic stirring for 30 min, a
FACs are aluminosilicate-rich wastes produced in coal-firing transparent solution, A, was formed. Meanwhile, 3.6 mL of vana-
power plants and cause extensive environmental problems owing dium (V) tri-i-propoxy oxide was dissolved in 75 mL of acetylace-
to the considerable volumes in which they exist. One feasible tone (0.2 mol L1) to obtain a dark brown solution, B. And the
way to treat these large quantities of FACs is by ash reuse. FACs stoichiometric ratio of Bi to V was 1:1. After mixing solutions A
have been reported as substrates in many studies, such as zeolite and B, a uniform dark green sol was obtained. The dark-green sol
[20], core–shell structure [21], catalyst and geopolymeric compos- was stirred vigorously for 1 h whereafter 10.2 g of FAC was added
ites [22] due to their low cost, nontoxicity, chemical and physical to the sol with stirring for 3 h at room temperature. The mixture
stability, low thermal conductivities and hollow framework. Be- thus obtained was evaporated at 80 °C in a water bath, drying at
sides the above mentioned points, we chose FACs as supports as 110 °C for 6 h, and then annealed in air at 500 °C for 2 h. The as-
they possess two additional properties: one is that they are a waste prepared samples were labeled as BiVO4/FACs. For comparison
product for use in waste treatment while the other is that FACs purposes, the FACs without grafted KH550, were added directly
have a low density and can float in water, rendering reclamation to the sol under the same MOD process and the samples named
of the photocatalysts easy while making full use of solar energy BiVO4/no-grafted FACs. Pure BiVO4 powder without FACs was syn-
in the photocatalytic process [16,23]. thesized using the same MOD method.
In this paper, thin films of BiVO4 with monoclinic structure
were coated on FACs by the MOD method. Prior to the loading of 2.3. Characterization
the film, the FAC surface was modified organofunctionally using si-
lane coupling agents, with the purpose of creating a suitable sur- The crystal structure of the composite was investigated by XRD
face environment favoring the formation of BiVO4 films. The (ARL, Switzerland) in the region of 2h = 5–70° using Cu Ka radia-
adsorption parameters of as-prepared samples were evaluated tion (k = 0.15418 nm). The surface morphology was investigated
and the photocatalytic abilities were detected by the degradation using an S-3400NII (Hitachi, Japan) scanning electron microscope
of MB in aqueous solution under visible-light irradiation. The ben- (SEM). XPS spectra were measured on an ESCALAB 250 spectrom-
efits of this fresh photocatalyst will be discussed, and the mecha- eter (Thermo, USA) with an Al Ka X-ray source (1486.6 eV). The
nism for the enhanced photocatalytic activity of BiVO4/FACs will UV–vis diffused reflectance spectroscopy (DRS) of the samples
be also researched. was recorded on a UV-2450 PC spectrometer (Shimadzu, Japan).
The dye mineralization degree was monitored in a 5000A total or-
ganic carbon content (TOC) analyzer (Shimadzu, Japan). The Bru-
2. Experimental methods
nauer–Emmett–Teller (BET) surface area was measured using an
ASAP 2010 nitrogen adsorption apparatus (Micromeritics, USA) at
2.1. Materials and reagents
77 K. The Electron spin-resonance spectroscopy (ESR) was con-
ducted on a Bruker model EMX 10/12 spectrometer (Bruker, Ger-
FACs were obtained from Nanjing Jinling Petrochemical Com-
many) equipped with a Hg lamp for measurement of the signals
pany. Silane coupling agents KH550 (c-aminopropyltriethoxysi-
of radicals spin-trapped by 5,5-dimethyl-1-pyrroline N-oxide
lane, C9H23O3CH2NSi, purity 99%) were purchased from Nanjing
(DMPO) at a microwave frequency of 9.77 GHz.
Shuguang Chemical Group (Nanjing, China). Vanadium (V) tri-i-
propoxy oxide was purchased from J&K Scientific Limited Corpora-
2.4. Adsorption experiments
tion (Beijing, China). All other chemicals used in this study were of
analytical grade and used without further purification. Deionized
Adsorption experiments were conducted in batch equilibrium
water was used in all test work.
mode. To determine the time required for the adsorption equilib-
rium of dye, 20 mL of 10 mg L1 MB solutions were mixed with
2.2. Preparation of BiVO4/FAC composites 0.08 g different photocatalysts in 50 mL iodine flasks. The flasks
were then transferred to an incubator shaker and vibrated at
160 rpm in dark. At different time intervals, an iodine flask was ta-
2.2.1. Pretreatment of FACs ken out and the mixture in it was filtered with a syringe filter of
First, the FACs were sieved and the particles with the size range 0.22 lm. The filtrate was analyzed for dye concentration. In equi-
of 100–125 lm were chosen as following experimental material. librium experiments, 0.08 g of photocatalysts was mixed with
J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746 739

20 mL different concentration dye solutions for continuously vi- observed. This suggests that both the structures of the pure BiVO4
brated for 24 h. And the initial concentrations of the different con- powder and BiVO4 film coated on the surface of FACs match closely
centration dye solutions were 5, 7.5, 10, 15, and 20 mg L1, with the characteristic peaks of monoclinic scheelite BiVO4 struc-
respectively. ture according to the JCPDS Card No. 14-0688. New peaks with
Analysis of the dye solutions was conducted with a UV-2550 2h at 11.5°, 23.2° and 27.8° are observed as a secondary phase in
UV–vis spectroscopy (Shimadzu, Japan). The adsorbed amounts BiVO4/FAC composites, which can be indexed to a tetragonal type
were then calculated by the following equation: Bi4V2O11 (JCPDS Card No. 42-0135). The formation of this oxide is
common in the system Bi2O3–V2O5 by using similar temperatures
ðC 0  CÞ  V [26]. However, the additional signals are weak, indicating that
Q¼ ð1Þ
m the amount of the Bi4V2O11 is low.
where Q is the amount absorbed (mg g1), C0 is the initial con-
centration of MB (mg L1), C is the concentration of MB at any time 3.2. SEM observation
(mg L1), m is the mass of adsorbent used (g) and V is the volume of
dye solution (L). SEM micrographs of the FACs, KH550-grafted FACs, BiVO4/FAC
and BiVO4/no-grafted FACs samples are shown in Fig. 2. The FACs
2.5. Photocatalytic degradation of MB in Fig. 2a and b exhibit an essentially spherical shape with diame-
ter of 120 lm and relatively uniform smooth surface. Fig. 2c and d
Photocatalytic activities of the samples were determined by the present the SEM images of the KH550-grafted FACs at low and high
decolorization of MB under visible light irradiation (>420 nm). A magnifications, respectively. It can be seen that KH550 is coated on
500 W Xe-illuminator served as a light source through UV cut-off the FAC surface, but the surface is uneven and consists of some
filters to completely remove any radiation below 420 nm, and pores and cracks. The cracks on the surface create some defects
was set approximately 10 cm from the reactor. Experiments were which provide active sites for the bonding of functional groups
carried out at ambient temperature as follows: 0.2 g as-prepared or metals [27]. In contrast, Fig. 2e and f present the surface
BiVO4/FACs photocatalyst was added to 50 mL of a 10 mg L1 MB micrograph of the BiVO4/FAC sample. The surface shows that the
solution. Before illumination, the solution was stirred for 30 min KH550-grafted FAC was completely coated with BiVO4. The coating
in darkness to reach the MB adsorption–desorption equilibrium. appears to consist of nanoparticles with an average size of 200–
Samples were taken at regular intervals under continuous stirring 400 nm (Fig. 2e). In addition, some deposits were inhomogeneous
and then filtered to remove the photocatalyst particles. The with islands being observed on the surface. This illustrates that
concentrations of the remnant MB were monitored by monitored coating of BiVO4 leads to surface roughness owing to shrinkage
the solution absorbance at 664 nm with UV–vis spectroscopy in the drying and calcination process [16]. For comparison,
(UV-2550, Shimadzu, Japan). For comparison, pure BiVO4 powder Fig. 2g and h present the surface micrograph of BiVO4/no-grafted
was used in the same experiment. FACs. Besides some sporadic deposits, no BiVO4 coating forms on
the FAC surface which confirms that silane coupling favors the
formation of BiVO4 films on FACs.
3. Results and discussion BET surface areas of the raw FACs, the KH550-grafted FACs, pure
BiVO4 and BiVO4/FACs samples are shown in Table. 1. The KH550-
3.1. XRD patterns grafted FACs were used for the adsorption and the photocatalytic
degradation of the MB dyes due to their higher surface areas than
The XRD spectra of the pure BiVO4 powder, FACs and BiVO4/ raw FACs. Unless specified, FACs used in the experiment were re-
FACs composite are presented in Fig. 1. Distinctive differences ferred to the KH550-grafted FACs. The BET surface area of BiVO4 in-
are reflected by the fact that the monoclinic scheelite BiVO4 gener- creased from 3.9 to 6.7 m2 g1 with the FACs as supports.
ally shows a clear splitting of peaks at 18.5°, 35° and 46° of 2h
[9,26]. In Fig. 1a and c, the peaks at 18.5°, 35° and 46° are split, 3.3. XPS analysis
besides, the peaks with 2h at 28.6°, 30.5°, 39.7° and 53.1° are also
The overall XPS spectra of the FACs, KH550-grafted FACs and
BiVO4/FACs composite samples are shown in Fig. 3. And Fig. 3a
shows the XPS spectrum of the uncoated FACs, indicating that
the particle surfaces are composed mainly of C, Si, Al and O, but
no N was found in the uncoated FACs. The XPS spectrum of the
KH550-grafted FACs sample (Fig. 3b) appears as an obscured new
peak at 400 eV and is attributed to the nitrogen of the silane amino
groups. The inset is the corresponding high-resolution XPS of N 1s,
showing that the silane is grafted successfully onto the FACs [28].
The new peaks of Bi 4f and V 2p at a bonding energy (BE) around
160 and 520 eV, respectively (Fig. 3c), became remarkably visible
in the XPS spectra of BiVO4/FACs.
Fig. 4 shows Bi 4f, V 2p and O 1s high-resolution XPS spectra of
the as-fabricated BiVO4/FACs samples. As shown in Fig. 4a, the
sample exhibits spin–orbit splitting signals of Bi 4f7/2 and Bi 4f5/2
at BE = 158.4 and 163.7 eV, respectively, which were characteristic
of Bi3+ [29]. The V 2p region is displayed in Fig. 4b with the char-
acteristic peaks at 516.0 and 523.6 eV ascribed to the V 2p3/2 and
V 2p1/2, respectively. The binding energies of the different elements
are attributable to the monoclinic scheelite BiVO4 [30]. In Fig. 4c, O
Fig. 1. The XRD patterns of (a) BiVO4; (b) FACs; and (c) BiVO4/FACs composite. 1s spectra were fitted roughly with two peaks, representing the
(diffraction lines associated with Bi4V2O11). presence of different oxygen species on the sample surface, the
740 J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746

(a) (b)

2 µm 50 µm

(c) (d)

2 µm 50 µm

(e) (f)

2 µm 50 µm

(g) (h)

10 µm
50 µm

Fig. 2. SEM images of (a and b) the FACs; (c and d) KH550-grafted FACs; (e and f) BiVO4/FACs and (g and h) BiVO4/no-grafted FACs.

Table 1 pate actively in the photocatalytic reaction and contribute


Surface areas of different samples. significantly to catalyst activity [31,33]. Meanwhile, O2 adsorbed
Sample SBET (m2 g1) on the BiVO4/FACs surface may accept e- and form  O
2 , which leads

Raw FACs 1.6


to the formation of OH in the system. More radicals existed in the
KH550-grafted FACs 7.3 system can result in a more rapid degradation of MB [34].
Pure BiVO4 3.9
BiVO4/FACs 6.7
3.4. DRS analysis

The UV–vis diffuse reflectance spectra of pure BiVO4 and BiVO4/


components at BE = 529.3 eV are characteristic of the lattice oxide FACs composite samples are shown in Fig. 5. The BiVO4/FACs
(OI) species, while the components at BE = 531.2 eV belong to the composite display an enhanced absorption compared with pure
adsorbed oxygen (OII) species [31,32]. Generally speaking, OII spe- BiVO4 in the region of 520–800 nm. It is apparent that the diffuse
cies have a higher mobility than lattice oxygen, which can partici- reflectance spectra of BiVO4/FACs composite exhibit a red shift
J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746 741

O1s
(a) 10000 Bi4f 7/2 158.4 (a)
16000

8000 Bi4f 5/2 163.7


12000
Counts (s)

6000

Counts (s)
8000
C1s 4000
4000 Si2S Si
2p
Al2p 2000

0
0
1000 800 600 400 200 0
Binding Energy (eV) 152 154 156 158 160 162 164 166 168 170 172

16000
Binding Energy (eV)
N1s 400
320

(b) O1s
300

280
3400 V2p 3/2 516.0
12000
260
(b)
Counts (s)

240
3200
220

200 3000
Counts (s)

180
8000 160 2800
390 395 400 405 410

Counts (s)
Binding energy (eV)
2600
C1s V2p 1/2 523.6
4000 2400
N1s Si2s Si2p
2200
Al2p
2000
0
1800
1000 800 600 400 200 0
1600
Binding Energy (eV)
514 516 518 520 522 524 526
35000 Binding Energy (eV)
(c) O1s
30000 Bi4f 7/2
5000 529.3
25000 V2p 3/2 Bi4f 5/2 O
1s
(c)
4500
20000
Counts (s)

4000
531.2
15000 C1s
Counts (s)

3500
10000
3000
5000
2500
0
2000
1000 800 600 400 200 0

Binding Energy (eV) 1500


525 530 535
Fig. 3. XPS pattern of (a) FACs; (b) KH550-grafted FACs; and (c) BiVO4/FACs
composite. Binding Energy (eV)

and increased absorption in the visible-light range. This observed Fig. 4. Bi 4f, V 2p, and O 1s high-resolution XPS spectra of the BiVO4/FACs
red-shift could be attributed to a secondary phase and surface de- composite.
fects, as BiVO4 was coated on other supports [18,19].
The band gap of a semiconductor can be estimated from a plot
of (ahm)2 versus incident photon energy (hm) [35,36], where a is the 3.5. Adsorption of MB dye on BiVO4/FACs composites
absorption coefficient. The estimated band gap energy (Eg) of pure
BiVO4 and BiVO4/FACs samples from the intercept of the tangents Adsorption of the organic substrate is generally considered to
to the plots are 2.29 and 2.21 eV, respectively. They are slightly be an important parameter in determining photocatalytic degrada-
smaller than the reported values of pure BiVO4 [7,8], which could tion rates [38]. The adsorption kinetics for the MB dye onto FACs,
be ascribed to large particle sizes. The decrease in band gap energy pure BiVO4 and BiVO4/FACs composites are represented in
indicates that BiVO4/FACs has a broader optical absorption region Fig. 6a. This figure shows that the amount of MB absorbed onto
than pure BiVO4, which can be excited to produce more elec- photocatalyst from aqueous solution increases quickly with time,
tron–hole pairs under the same visible light irradiation and then and equilibrium is established within 30 min for all different
result in higher photocatalytic activity [19,37]. photocatalysts. Therefore, this time has been selected for the initial
742 J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746

Table 2
Langmuir parameters for MB adsorption onto different samples.
1.0
Adsorbent KL (L mol1) Qmax (mg g1)
FACs 1.78  105 1.55
0.8
Pure BiVO4 1.86  105 1.64
Absorbance (a.u.)

BiVO4/FACs 5.29  105 2.94

0.6
ent types of adsorbents [38,39]. In this paper, the different values
(b) of the dye MB adsorbed at steady state have been plotted in a Lang-
0.4 muirian plot 1/Qe = f (1/Ce) (Fig. 6b). And the linearity of the trans-
forms indicates that the Langmuir isotherm is correctly observed:
0.2 1 1 1 1
(a) ¼ þ 
Q e Q m Q mK L Ce
ð2Þ

0.0 where Ce (mg L1) is the equilibrium concentration of MB, Qe


200 300 400 500 600 700 800 (mg g1) is the amount adsorbed under equilibrium, Qm (mg g1)
Wavelength (nm) is the theoretical maximum adsorption capacity of the adsorbent
for MB, and KL is the adsorption constant. According to the data
Fig. 5. UV–vis diffuse reflectance spectra of (a) pure BiVO4 and (b) BiVO4/FACs
in Fig. 6b, one can determine the Qm from the intercept of the
composite.
curve, and the adsorption constants KL can be deduced from the
slope of the curve (1/Qm KL). These values are shown in Table 2.
In addition, the relation curves between removal rate of MB by
2.5 (a) adsorption process and initial concentrations of dye were depicted
in Fig. S1 (Supporting information). When the initial concentrations
of MB dye varied from 5 to 20 mg L1, the percentage of dye ad-
2.0
sorbed on the surface of the FACs ranged between 9% and 17%, the
corresponding values for the pure BiVO4 and BiVO4/FACs composites
1.5
Q (mg g-1)

were 11–21% and 30–58%, respectively. The photocatalysis process


can be improved efficiently by collecting the dye in the solution to
1.0 the surface of photocatalyst [40]. Therefore, the adsorption of MB
on the surface of the BiVO4/FACs is very important for the subse-
BiVO4 / FACs quent photocatalysis process, and the more adsorption capacity
0.5
FACs maybe lead to the higher photodegradation rate.
pure BiVO4
0.0 3.6. Photocatalytic activity
0 20 40 60 80 100
Time (min) Photocatalytic activity was evaluated by measuring the degra-
dation of the MB dye solution. In order to distinguish the portion
1.1 2 that the MB molecules were removed by adsorption and photodeg-
R = 0.9813
(b) radation, Fig. 7a shows the temporal evolution of the MB concen-
1.0 2
R = 0.9762 tration. According to the experimental conditions the curve can
be separated in two stages: (1) adsorption of MB dye in the dark-
0.9
ness and (2) irradiation of dispersion with visible light. After
0.8 30 min of dark equilibration, the percentage of MB dye adsorbed
BiVO4 / FACs
1/Qe (g mg-1)

on the surface of the pure BiVO4 and BiVO4/FACs are 13% and
pure BiVO4
0.7 42%, respectively. This indicates that BiVO4/FACs composites ab-
FACs
sorb more MB dye molecules on their surface than pure BiVO4
0.6 powder (approximately threefold more) [8]. When the lamp turns
on, the concentration of MB undergoes depletion.
0.5
Fig. 7b shows the photodegradation rates of MB on BiVO4/FACs
2

0.4
R = 0.9931 and pure BiVO4 powder. The photodegradation rates of MB reached
90% after irradiation for 300 min in the presence of the BiVO4/FACs
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 samples, while the photodegradation rate of MB over pure BiVO4
-1 was 55% after irradiation for 300 min under the same conditions.
1/Ce(L mg )
The neat photodegradation rate of MB in the absence of any photo-
catalyst was about 10% after 5 h, which is far below than that of the
Fig. 6. (a) Adsorption kinetics of MB onto different samples in dark and conditions:
C0 = 10 mg/L; m = 80 mg; V = 20 mL, T = 25 °C; (b) linear transforms of the Langmuir photocatalytic process.
isotherms. According to the report of Wang [41], the following first-order
reaction kinetics can explain the MB degradation rate in the photo-
catalytic oxidation process.
period in the dark before visible light irradiation. When the adsorp-
dC
tion reaches equilibrium, the amount of MB adsorbed onto BiVO4/  ¼ kC ð3Þ
FACs composites is the largest, which may be attributed to the syn- dt
ergistic effect between BiVO4 and FACs. where k is the first-order rate constant (min1) and C is the MB
In pioneer works, the Langmuir model has been successfully concentration (mg L1) at time t (min). And the linear relationship
employed to determine the adsorption constant of dyes on differ- of ln(C0/C) versus time for degradation of MB using different
J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746 743

10 3.0
(a) (c)
8 BiVO4 / FACs 2.5

pure BiVO4 BiVO4 / FACs


Concentration (mg L-1)

2.0 pure BiVO4


6

ln (C0/C)
1.5

4
1.0

2 0.5

0.0
0 0 60 120 180 240 300
-30 0 30 60 90 120 150 180 210 240 270 300
Time (min) Irradiation time (min)

1.0
1.0 (d)
0.8
(b) 0.8 ....MB without photocatalyst
0 min

Absorbance
0.6 30 min
0.6 60 min
C/C0

90 min
120 min
0.4 0.4 180 min
240 min
270 min
0.2 BiVO 4 / FACs 0.2 300 min
pure BiVO 4
MB photolysis
0.0 0.0
0 60 120 180 240 300 400 500 600 700 800

Irradiation time (min) Wavelength (nm)

Fig. 7. Photocatalytic degradation of MB dye solution over pure BiVO4 and BiVO4/FACs composite: (a) MB concentration profile, including dark adsorption; (b) conversion in
terms of the ratio of remaining MB concentration (C) to initial concentration (C0); (c) logarithmic plot and (d) the absorption spectral change in the concentration of MB using
BiVO4/FACs composite as photocatalyst.

Fig. 7d. It can be seen that the kmax gradually shifts from 664 to
Table 3
642 nm with the decrease in absorbance during visible light irradi-
Rate constants of MB photodecomposition and linear regression coefficients from a
plot of ln(C0/C) = kt with different samples. ation. According to the reported by Takizawa et al. [43], the blue
shifts from 664 to 642 nm are attributed to the N-demethylation
Photocatalysts Regression equation R2 K (min1)
of MB, and the formation of N,N,N0 -trimethylated azure B (AB). In
Pure BiVO4 y = 0.05793 + 0.00252x 0.9834 0.00252 order to study the mineralization of the dyes, the total organic con-
BiVO4/FACs y = 0.29037 + 0.00642x 0.9547 0.00642
tent (TOC) of MB in the sample was measured. Fig. S2 (Supporting
information) shows the TOC removal of the dye solution with
BiVO4/FACs composite as photocatalysts, and the mineralization
samples is shown in Fig. 7c. The calculated k values of BiVO4 and rate is about 53% after 5 h of visible light irradiation. Due to the
the BiVO4/FACs composites are 0.00252 and 0.00642 min1, formation of reaction intermediates, the MB dye cannot be
respectively. The rate constant for BiVO4/FACs is approximately completely degraded.
2.5 times higher than that of BiVO4 as shown in Table.3. The deg- The aspects of BiVO4/FACs samples in different experimental
radation of MB using uncoated FAC shows no obvious decrease in processes are shown in Fig. S3 (Supporting information). It can
MB concentration, confirming that this fresh substrate has no pho- be seen that the fresh sample is yellow, and it turns to yellow-
tocatalytic activity. However, the presence of FACs in the BiVO4/ green under adsorption equilibrium, but it returns to the primary
FACs composite has two important functions. One is that the colors after photocatalysis. Though 42% of MB molecules were ad-
adsorption ability of modified FAC can increase the adsorption sorbed on the surface of BiVO4/FACs, most of them were degraded
capacity of MB on the surface of this substrate in favor of degrading after photocatalysis.
contamination, and that active free radicals adhering on FAC can
react with MB. The second is to act as a dispersing support to inhi- 3.7. Stability of catalyst
bit grain growth, which contributes to making full use of light for
photocatalysis [16,42]. Owing to a synergistic effect between BiVO4 To evaluate the stability of the BiVO4/FACs catalyst, it was re-
and the FACs, the BiVO4/FACs composite exhibits high photocata- used in four successive photocatalytic experiments by calcining
lytic activity. at 500 °C for 2 h after reaction. Fig. 8 shows results from the four
The absorption spectral change in the concentration of MB successive runs for the photodegradation of MB under the same
using BiVO4/FACs composite as photocatalyst is displayed in experimental conditions. No significant loss of activity was found
744 J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746

in four successive runs and the removal of MB remains higher than


80% in each cycle, testifying that the photocorrosion of BiVO4/FACs (a)
was negligible and the as-prepared catalyst was stable during the
photodegradation of MB.
However, this regeneration method is high energy consumption

Intensity
and it needs to be improved. And we will further explore more
efficient regeneration method.

3.8. Possible reaction mechanisms of photodegradation of MB over


BiVO4/FACs

The production of reactive radical groups (mainly correspond- 3440 3460 3480 3500 3520
ing to superoxide and hydroxyl radicals) at the solid–liquid inter- Magnetic Field (G)
face was confirmed by the ESR spin-trapping by the DMPO
method with results as shown in Fig. 9. As shown in Fig. 9a, the
typical ESR spectrum of the DMPO-OH adduct with a quartet sig- (b)
nal (intensity ratio of 1:2:2:1) was observed in the light irradiated
suspension of BiVO4/FACs and six characteristic peaks of DMPO-
 
O2 can be observed (Fig. 9b) in the BiVO4/FACs methanolic disper-

Intensity
sions [44,45]. These results indicate that the photogenerated
electrons and holes have survived long enough to react with the
surface adsorbed O2 and H2O to produce  O 
2 or OH radicals [19].
  
Both O2 and OH radicals are strong oxidants which can non-
selectively degrade organic and inorganic pollutants absorbed on
the catalyst surface during the photocatalytic process, while OH
radicals produced are often considered to be the major species
responsible for the photocatalytic oxidation reaction [46–48]. 3440 3460 3480 3500 3520
From the experimental results, a possible mechanism for the Magnetic Field (G)
photodegradation of MB over a BiVO4/FACs composite catalyst
can be described by the following steps (Eqs. (4)–(8)): Fig. 9. ESR spectrum of BiVO4/FACs samples under irradiation: (a) in aqueous
dispersion for DMPO-OH and (b) in methanol dispersion for DMPO-O2.
þ
BiVO4 þ hm ! BiVO4 ðh þ e Þ ð4Þ
tured by O2 to produce a superoxide anion radical  O 2 (Eq. (5)),
BiVO4 ðe Þ þ O2 !  O2 þ BiVO4 ð5Þ
while the produced holes can react with the H2O (Eq. (7)) to form
the hydroxyl radical, OH. Moreover, the generated  O 2 can react

O2 þ H2 O !  OH þ OH ð6Þ
further with the adsorbed H2O to produce more reactive OH radi-
þ
cals (Eq. (6)). Finally, these radicals group decompose the MB dye
BiVO4 ðh Þ þ H2 O !  OH þ Hþ þ BiVO4 ð7Þ molecules adsorbed on the BiVO4/FACs composite photocatalyst
(Eq. (8)).

OH þ MBðdyeÞ ! deg radation products ð8Þ
First, under the irradiation of visible light, electrons receive en- 4. Conclusions
ergy from the photons and are excited from the valence band to the
conduction band of BiVO4, leaving positively charged holes in the In this paper, BiVO4 film was coated successfully on the surface
valence band (Eq. (4)). The photogenerated electrons are then cap- of FACs through the modified MOD method. The SEM images con-
firm that the BiVO4 film was relatively compact and the silane cou-
pling is in favor of the formation of BiVO4 films on the FAC surface.
1.0
XPS data indicate the presence of different oxygen species on the
1st run 2nd run 3rd run 4th run sample surface. These belong to the lattice oxide (OI) and adsorbed
0.8 oxygen (OII) species. The DRS diagrams demonstrate that the
BiVO4/FACs catalysts show absorption in the visible region be-
tween 550 and 800 nm. It is also found that the introduction of
0.6 FACs could effectively narrow down the catalyst band gap. The
Langmuir adsorption constant was evaluated, and the maximum
C/C0

0.4
amount of dye adsorbed (Qmax) for BiVO4/FACs is 1.8 times more
than that for pure BiVO4. The photocatalytic activity results indi-
cate that BiVO4/FACs exhibit excellent photocatalytic activity with
0.2 a first-order rate constant value 2.5 times higher than pure BiVO4
for decomposition of MB under visible light irradiation. The en-
hanced photocatalytic performance under visible light irradiation
0.0
can be attributed to the efficient separation of photogenerated
0 1 2 3 4 50 1 2 3 4 50 1 2 3 4 50 1 2 3 4 5
electron–hole pairs in the BiVO4 and FACs coupling system. The
Time (h) presence of FACs in the BiVO4/FACs composite has two important
Fig. 8. Cycling runs in the photocatalytic degradation of MB in the presence of
functions. One is that the adsorption ability of modified FAC based
BiVO4/FACs sample at initial concentration of 10 mg L1 under visible light on the pore structure can increase the adsorption capacity of MB
irradiation. on this substrate surface in favor of degrading contamination,
J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746 745

and active free radicals adhering on modified FAC can react with [15] C. Sriwong, S. Wongnawa, O. Patarapaiboolchai, Recyclable thin TiO2-
embedded rubber sheet and dye degradation, Chem. Eng. J. 191 (2012) 210–
MB. The other is to act as a dispersing support to inhibit grain
217.
growth, which contributes to making full use of light for photoca- [16] B. Wang, Q. Li, W. Wang, Y. Li, J.P. Zhai, Preparation and characterization of
talysis. Owing to the low density of FAC, the as-prepared BiVO4/ Fe3+-doped TiO2 on fly ash cenospheres for photocatalytic application, Appl.
FAC particles can float in water. This favors phase separation by Surf. Sci. 257 (2011) 3473–3479.
[17] Y. Zhang, J. Yu, K. Akihiko, X.S. Zhao, Preparation of BiVO4-MCM-41 composite
filtration to recover these photocatalysts after reaction and the catalyst and its photocatalytic activity for degradation of methylene blue,
recovery test shows that the BiVO4/FACs was rather stable during Chinese Journal of Catalysis 29 (2008) 624–628.
MB photodegradation. [18] Y.S. Fu, X. Sun, X. Wang, BiVO4-graphene catalyst and its high photocatalytic
performance under visible light irradiation, Mater. Chem. Phys. 131 (2011)
Using the as-prepared BiVO4/FACs as photocatalysts to degrade 325–330.
dye contamination makes FACs reuse possible, and FACs have the [19] W.R. Zhao, Y. Wang, Y. Yang, J. Tang, Y. Yang, Carbon spheres supported
advantages of low cost. The composite catalyst is therefore visible-light-driven CuO–BiVO4 heterojunction: preparation, characterization,
and photocatalytic properties, Appl. Catal. B: Environ. 115–116 (2012) 90–99.
promising for practical applications in water purification. [20] J. Lu, F. Xu, D. Wang, J. Huang, W.M. Cai, The application of silicalite-1/fly ash
cenosphere (S/FAC) zeolite composite for the adsorption of methyl tert-butyl
Acknowledgement ether (MTBE), J. Hazard. Mater. 165 (2009) 120–125.
[21] J.F. Pang, Q. Li, B. Wang, D.J. Tao, X.T. Xu, W. Wang, J.P. Zhai, Preparation and
characterization of electroless Ni–Fe–P alloy films on fly ash cenospheres,
The authors gratefully acknowledge financial supports from the Powder Technol. 226 (2012) 246–252.
Foundation of State Key Laboratory of Pollution Control and Re- [22] N.W. Tan, A. Riessen, C.V. Ly, D.C. Southam, Determining the reactivity of a fly
ash for production of geopolymer, J. Am. Ceram. Soc. 92 (2009) 881–887.
source Reuse of China, the China Postdoctoral Science Foundation [23] Q. Li, X.T. Xu, H. Cui, J.F. Pang, Z.B. Wei, Z.Q. Sun, J.P. Zhai, Comparison of two
funded project (No. 2012M511254), the Natural Science Founda- adsorbents for the removal of pentavalent arsenic from aqueous solutions, J.
tion of China (No. 51008154), the Jiangsu cultivate the innovative Environ. Manage. 98 (2012) 98–106.
[24] B. Wang, Q. Li, J.F. Kang, J.F. Pang, W. Wang, J.P. Zhai, Preparation and
engineering for graduate student (CXZZ12-0063), as well as the characterization of polypyrrole coating on fly ash cenospheres: role of the
Scientific Research Foundation of Graduate School of Nanjing organosilane treatment, J. Phys. D Appl. Phys. 44 (2011) 415301 (9pp).
University (No. 2012CL10). We greatly appreciate the reviewer’s [25] A. Galembeck, O.L. Alves, BiVO4 thin film preparation by metalorganic
decomposition, Thin Solid Films 365 (2000) 90–93.
useful suggestion and yours suggestion make this paper enrich.
[26] U.M. García Pérez, S. Sepúlveda-Guzmán, A. Martínez-de la Cruz, U. Ortiz
Méndez, Photocatalytic activity of BiVO4 nanospheres obtained by solution
combustion synthesis using sodium carboxymethylcellulose, J. Mol. Catal. A:
Appendix A. Supplementary material
Chem. 335 (2011) 169–175.
[27] Z.C. Kang, Z.L. Wang, On accretion of nanosize carbon spheres, J. Phys. Chem.
Supplementary data associated with this article can be found, in 100 (1996) 5163–5165.
[28] L.X. Zhang, Q. Jin, J. Huang, Y. Liu, L. Shan, X.G. Wang, Modification of
the online version, at http://dx.doi.org/10.1016/j.cej.2012.12.065.
palygorskite surface by organofunctionalization for application in
immobilization of H3PW12O40, Appl. Surf. Sci. 256 (2010) 5911–5917.
References [29] H.Y. Jiang, H.X. Dai, X. Meng, K.M. Ji, L. Zhang, J.G. Deng, Porous olive-like
BiVO4: alcoho-hydrothermal preparation and excellent visible-light-driven
photocatalytic performance for the degradation of phenol, Appl. Catal. B:
[1] H. Fan, D.J. Wang, L.L. Wang, H.Y. Li, P. Wang, T.F. Jiang, T. Xie, Hydrothermal
Environ. 105 (2011) 326–334.
synthesis and photoelectric properties of BiVO4 with different morphologies:
[30] L. Dong, S. Guo, S. Zhu, D.F. Xu, L. Zhang, M. Huo, X. Yang, Sunlight responsive
an efficient visible-light photocatalyst, Appl. Surf. Sci. 257 (2011) 7758–7762.
BiVO4 photocatalyst: effects of pH on L-cysteine-assisted hydrothermal
[2] A. Martínez-de la Cruz, U.M. García Pé rez, Photocatalytic properties of BiVO4
treatment and enhanced degradation of ofloxacin, Catal. Commun. 16 (2011)
prepared by the co-precipitation method: degradation of rhodamine B and
250–254.
possible reaction mechanisms under visible irradiation, Mater. Res. Bull. 45
[31] H. Chen, A. Sayari, A. Adnot, F. Larachi, Composition-activity effects of Mn–Ce–
(2010) 135–141.
O composites on phenol catalytic wet oxidation, Appl. Catal. B: Environ. 32
[3] D.K. Lee, I.S. Cho, S.W. Lee, S.T. Bae, J.H. Noh, D.W. Kim, K.S. Hong, Effects of
(2001) 195–204.
carbon content on the photocatalytic activity of C/BiVO4 composites under
[32] P.M. Kumar, S. Badrinarayanan, M. Sastry, Nanocrystalline TiO2 studied by
visible light irradiation, Mater. Chem. Phys. 119 (2010) 106–111.
optical, FTIR and X-ray photoelectron spectroscopy: correlation to presence of
[4] J.Q. Yu, Y. Zhang, A. Kudo, Synthesis and photocatalytic performances of BiVO4
surface states, Thin Solid Films 358 (2000) 122–130.
by ammonia co-precipitation process, J. Solid State Chem. 182 (2009) 223–228.
[33] S.X. Yang, Y. Feng, J. Wan, W. Zhu, Z. Jiang, Effect of CeO2 addition on the
[5] X.F. Zhang, X. Quan, S. Chen, Y. Zhang, Effect of Si doping on
structure and activity of RuO2/c-Al2O3 catalyst, Appl. Surf. Sci. 246 (2005)
photoelectrocatalytic decomposition of phenol of BiVO4 film under visible
222–228.
light, J. Hazard. Mater. 177 (2010) 914–917.
[34] Y. Yu, J.C. Yu, C.Y. Chan, Y.K. Che, J.C. Zhao, L. Ding, W. Kun, P.K. Wong,
[6] Y. Lu, Y.S. Luo, D.Z. Kong, D.Y. Zhang, Y.L. Jia, X.W. Zhang, Large-scale
Enhancement of adsorption and photocatalytic activity of TiO2 by using carbon
controllable synthesis of dumbbell-like BiVO4 photocatalysts with enhanced
nanotubes for the treatment of azo dye, Appl. Catal. B: Environ. 61 (2005) 1–11.
visible-light photocatalytic activity, J. Solid State Chem. 186 (2012) 255–260.
[35] B. Cheng, W. Wang, L. Shi, J. Zhang, J. Ran, H. Yu, One-pot template-free
[7] W. Sun, M.Z. Xie, L.Q. Jing, Y. Luan, H. Fu, Synthesis of large surface area nano-
hydrothermal synthesis of monoclinic BiVO4 Hollow microspheres and their
sized BiVO4 by an EDTA-modified hydrothermal process and its enhanced
enhanced visible-light photocatalytic activity, Int. J. Photoenergy (2012) 10p.
visible photocatalyticactivity, J. Solid State Chem. 184 (2011) 3050–3054.
[36] N. Serpone, D. Lawless, R. Khairutdinov, Size effects on the photophysical
[8] U.M. García-Pérez, S. Sepúlveda-Guzmán, Nanostructured BiVO4
properties of colloidal anatase TiO2 particles: size quantization or direct
photocatalysts synthesized via a polymer-assisted coprecipitation method
transitions in this indirect semiconductor, J. Phys. Chem. B. 99 (1995) 16646–
and their photocatalytic properties under visible-light irradiation, Solid State
16654.
Sci. 14 (2012) 293–298.
[37] A.P. Zhang, J.Z. Zhang, Characterization and photocatalytic properties of Au/
[9] Z.J. Zhang, W. Wang, M. Shang, W. Yin, Photocatalytic degradation of
BiVO4 composites, J. Alloy. Compd. 491 (2010) 631–635.
rhodamine B and phenol by solution combustion synthesized BiVO4
[38] A. Martínez-de la Cruz, S. Obregón Alfaro, Synthesis and characterization of
photocatalyst, Chem. Commun. 11 (2010) 982–986.
nanoparticles of a-Bi2Mo3O12 prepared by co-precipitation method: Langmuir
[10] C.Y. Wang, H. Zhang, F. Li, L. Zhu, Degradation and mineralization of Bisphenol
adsorption parameters and photocatalytic properties with rhodamine B, Solid
A by mesoporous Bi2WO6 under simulated solar light irradiation, Environ. Sci.
State Sci. 11 (2009) 829–835.
Technol. 44 (2010) 6843–6848.
[39] H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard, J.-M.
[11] J. Fan, X.Y. Hu, Z.G. Xie, K.L. Zhang, J.J. Wang, Photocatalytic degradation of azo
Herrmann, Photocatalytic degradation of various types of dyes (Alizarin S,
dye by novel Bi-based photocatalyst Bi4TaO8I under visible-light irradiation,
Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV-
Chem. Eng. J. 179 (2012) 44–51.
irradiated titania, Appl. Catal. B: Environ. 39 (2002) 75–90.
[12] G.L. Li, Z. Yin, Theoretical insight into the electronic, optical and photocatalytic
[40] C. Hu, Y.C. Tang, J.C. Yu, P.K. Wong, Photocatalytic degradation of cationic blue
properties of InMO4 (M = V, Nb, Ta) photocatalysts, Phys. Chem. Chem. Phys.
X-GRL adsorbed on TiO2/SiO2 photocatalyst, Appl. Catal. B: Environ. 40 (2003)
13 (2011) 2824–2833.
131–140.
[13] F. Lin, D. Wang, Z. Jiang, Y. Ma, J. Li, R. Li, C. Li, Photocatalytic oxidation of
[41] D.S. Wang, Y.H. Wang, X.Y. Li, Q.Z. Luo, J. An, H.X. Yue, Sunlight photocatalytic
thiophene on BiVO4 with dual co-catalysts Pt and RuO2 under visible light
activity of polypyrrole-TiO2 nanocomposites prepared by ‘in situ’ method,
irradiation using molecular oxygen as oxidant, Energy Environ. Sci. 5 (2012)
Catal. Commun. 9 (2008) 1162–1166.
6400–6406.
[42] M.V. Sharma, V.D. Kumari, A. Subrahmanyam, Photocatalytic degradation of
[14] M. Asiltürk, Sßadiye ßsener, TiO2-activated carbon photocatalysts: preparation,
isoproturon herbicide over TiO2/Al-MCM-41 composite systems using solar
characterization and photocatalytic activities, Chem. Eng. J. 180 (2012) 354–
light, Chemosphere 72 (2008) 644–651.
363.
746 J. Zhang et al. / Chemical Engineering Journal 223 (2013) 737–746

[43] T. Takizawa, T. Watanabe, K. Honda, Photocatalytic through excitation of [46] B. Zhou, X. Zhao, H.J. Liu, J.H. Qu, C.P. Huang, Synthesis of visible-light sensitive
adsorbates. 2. A comparative study of rhodamine B and methylene blue on M-BiVO4 (M = Ag, Co, and Ni) for the photocatalytic degradation of organic
cadmium sulfide, J. Phys. Chem. 82 (1978) 1391–1396. pollutants, Sep. Purif. Technol. 77 (2011) 275–282.
[44] H. Yamashita, Y. Ichihashi, S. Zhang, Y. Matsumura, Y. Souma, T. Tatsumi, M. [47] K. Ishibashi, A. Fujishima, T. Watanabe, K. Hashimoto, Detection of active
Anpo, Photocatalytic decomposition of NO at 275 K on titanium oxide catalysts oxidative species in TiO2 photocatalysis using the fluorescence technique,
anchored within zeolite cavities and framework, Appl. Surf. Sci. 121 (122) Electrochem. Commun. 2 (2000) 207–210.
(1997) 305–309. [48] J.W. Kim, C.W. Lee, W.Y. Cho, Platinized WO3 as an environmental
[45] H.B. Fu, L.W. Zhang, S.C. Zhang, Y.F. Zhu, Electron spin resonance spin-trapping photocatalyst that generates OH radicals under visible light, Environ. Sci.
detection of radical intermediates in N-Doped TiO2-assisted photodegradation Technol. 44 (2010) 6849–6854.
of 4-chlorophenol, J. Phys. Chem. B. 110 (2006) 3061–3065.

Вам также может понравиться